Next Article in Journal
Palaeostomate Bryozoans from Glacial Erratics in the Tvären Region, Sweden
Next Article in Special Issue
Alteration Lithogeochemistry of an Archean Porphyry-Type Au(-Cu) Setting: The World-Class Côté Gold Deposit, Canada
Previous Article in Journal
Selective Flotation Separation of Chalcopyrite from Copper-Activated Pyrite and Pyrrhotite Using Oxidized Starch as Depressant
Previous Article in Special Issue
Shortwave Infrared Spectroscopy and Geochemical Characteristics of White Mica-Group Minerals in the Sinongduo Low-Sulfide Epithermal Deposit, Tibet, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interpreting the Complexity of Sulfur, Carbon, and Oxygen Isotopes from Sulfides and Carbonates in a Precious Metal Epithermal Field: Insights from the Permian Drake Epithermal Au-Ag Field of Northern New South Wales, Australia

1
BGRIMM Technology Group, Beijing 100160, China
2
Earth and Sustainability Science Research Centre, School of Biological, Earth and Environmental Sciences, UNSW Sydney, Kensington, NSW 2052, Australia
3
White Rock Minerals Ltd., 12 Anderson Street, West Ballarat, VIC 3350, Australia
4
Cipango Limited, Suite 506 Level 5, 50 Clarence Street, Sydney, NSW 2000, Australia
5
Bioanalytical Mass Spectrometry Facility, UNSW Sydney, Kensington, NSW 2052, Australia
6
Central Science Laboratory, University of Tasmania, Hobart, TAS 7001, Australia
7
Bowdens Silver Pty Ltd., 68 Maloneys Road, Lue, NSW 2850, Australia
8
Solid State & Elemental Analysis Unit, Mark Wainwright Analytical Centre, UNSW Sydney, Kensington, NSW 2052, Australia
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(2), 134; https://doi.org/10.3390/min15020134
Submission received: 24 November 2024 / Revised: 20 January 2025 / Accepted: 23 January 2025 / Published: 29 January 2025

Abstract

:
The Drake Goldfield, also known as Mount Carrington, is located in north-eastern New South Wales, Australia. It contains a number of low–intermediate-sulfidation epithermal precious metal deposits with a current total resource of 724.51 metric tons of Ag and 10.95 metric tons of Au. These deposits occur exclusively within the Drake Volcanics, a 60 × 20 km NW-SE trending sequence of Late Permian volcanics and related epiclastics. Drilling of the Copper Deeps geochemical anomaly suggests that the volcanics are over 600 m thick. The Drake Volcanics are centered upon a geophysical anomaly called “the Drake Quiet Zone” (DQZ), interpreted to be a collapsed volcanic caldera structure. A total of 105 fresh carbonate samples were micro-drilled from diamond drillcores from across the field and at various depths. A pXRD analysis of these carbonates identified five types as follows: ankerite, calcite, dolomite, magnesite, and siderite. Except for three outlier values (i.e., −21.32, −19.48, and 1.42‰), the δ13CVPDB generally ranges from−15.06 to −5.00‰, which is less variable compared to the δ18OVSMOW, which varies from −0.92 to 17.94‰. μ-XRF was used to analyze the elemental distribution, which indicated both syngenetic/epigenetic relationships between calcite and magnesite. In addition, a total of 53 sulfide samples (primarily sphalerite and pyrite) from diamond drillcores from across the Drake Goldfield were micro-drilled for S isotope analysis. Overall, these have a wide range in δ34SCDT values from −16.54 to 2.10‰. The carbon and oxygen isotope results indicate that the fluids responsible for the precipitation of carbonates from across the Drake Goldfield had complex origins, involving extensive mixing of hydrothermal fluids from several sources including those of magmatic origin, meteoric fluids and fluids associated with low-temperature alteration processes. Sulfur isotope ratios of sulfide minerals indicate that although the sulfur was most likely derived from at least two different sources; magmatic sulfur was the dominant source while sedimentary-derived sulfur was more significant for the deposits distal from the DQZ, with the relative importance of each varying from one deposit to another. Our findings contribute to a greater understanding of Au-Ag formation in epithermal environments, particularly in collapsed calderas, enhancing exploration strategies and models for ore deposition.

1. Introduction

Epithermal deposits, which form at shallow depths and low temperatures, are significant sources of precious metals, such as Au and Ag [1]. Collapsed calderas are considered ideal locations for these types of deposits, as they provide the necessary conditions for the formation of mineralization from ascending thermal waters. These deposits typically occur in the shallow parts of magmatic systems, at depths ranging from ~50 to 1500 m, with ore minerals precipitating at temperatures between <150 °C and ~300 °C, often within associated volcano-sedimentary basins [2,3,4,5].
The Drake Goldfield, also known as Mount Carrington, is located in north-eastern New South Wales (NSW) Australia, ~5 km north-east of Drake Village, 44 km east of Tenterfield, and ~800 km north of Sydney (Figure 1). It contains a number of low–intermediate-sulfidation epithermal precious metal deposits (Figure 2), including Kylo, Strauss, Red Rock, Lady Hampden, Silver King, White Rock, and White Rock North [6], along with numerous other small deposits and prospects. Lady Hampden is unusual for the Drake Goldfield in being a Au-Ag deposit; All Nations, Gladstone Hill, and West Copper Deeps are all Cu prospects. The Strauss, Kylo, Silver King, and Lady Hampden deposits are collectively known as the Mount Carrington Group, as they are in close proximity to the historical Mount Carrington Au-Ag deposit. These deposits occur exclusively within the Drake Volcanics, which comprise a 60 × 20 km NW-SE trending sequence of Late Permian shallow volcanics and related epiclastics.
The first significant precious metal deposit discovered in the Drake Volcanics was the White Rock silver deposit which was discovered in 1886, and its geology was first described by Andrews [9]. Subsequently, many deposits of differing styles were discovered and mined. Although being known and mined for over 100 years, there has been no detailed study on the deposits associated with the Drake Volcanics as a whole, apart from that of Perkins [10] on the deposits of the Red Rock Field, which is one of the smallest of the fields within the larger Drake Goldfield. Recent work by White Rocks Minerals suggests that most of the economic precious metal deposits within the Drake Volcanics are centered upon a geophysical anomaly called “the Drake Quiet Zone” (DQZ), interpreted to be a collapsed volcanic caldera structure (Figure 3). The Drake Goldfield has been mined since the late 1800s for precious metals, particularly Au and Ag [6,9] with the most recent mining activity by Mt. Carrington Mines (1988 to 1990) producing 0.714 metric tons of Au and 434,870 oz of Ag [11]. Recent estimates suggest that the Drake Goldfield contains an Indicated Resource of 724.51 metric tons of Ag and 10.95 metric tons of Au [6]. The present study builds upon more recent unpublished studies on the relationships between mineralization, grade, and alteration assemblages/intensity on the White Rock and the Lady Hampden deposits [12,13]; the relationship between primary volcanic facies and mineralization and correlation between the White Rock and White Rock North deposits [14]; the structural controls and relationship between mineralization and alteration for the Strauss deposit [15]; and a detailed mineralogical and geochemical study on selected Ag-rich deposits within the Drake Goldfield [16].
As there were no comprehensive, published studies across the Drake Goldfield itself and on mineralization within the Drake Volcanics, the main objective of this study was to document the isotopic compositional variations of sulfides and carbonates that precipitated during different stages across the Drake Goldfield. A number of vein-like carbonates and sulfides from mineralized horizons from diamond drillcores from across the Drake Goldfield and at various depths were systematically studied for their carbon, oxygen, and sulfur isotopes in order to better understand the source of the mineralizing and alteration fluids and to provide information about the genesis of the Drake Goldfield.

2. Geological Setting

2.1. The Southern New England Orogen

The Drake Goldfield is near the north-eastern boundary of the Southern New England Orogen (SNEO) or Southern New England Fold Belt [6,18] (Figure 1). The Southern New England Orogen is a sub-province of the New England Orogen (NEO), the youngest sub-province of the Tasmanides, whose formation is widely accepted to comprise two main cycles of compression and extension since the early Cambrian [7,19,20]. The NEO extends ~1600 km along the east coast of Australia from Townsville in northern Queensland to Newcastle in New South Wales (Figure 1). It is divided by the Clarence–Moreton Basin into northern and southern provinces, which are the Yarrol–Gympie Province and SNEO, respectively [19,21]. The SNEO is further divided into two regions, the Tamworth Belt and the Tablelands Complex, divided by the Peel-Manning Fault System [7,22]. The Tamworth Belt represents a relatively weakly deformed fore arc environment, while the Tablelands Complex comprises a series of accretionary wedge/subduction complexes [19,21]. During the uppermost Carboniferous to Lower/Middle Permian, the SNEO had been through an extensional cycle followed by widespread orogenic deformation, granitoid intrusion and volcanism [18,20]. The host rocks of the Drake Goldfield are mostly volcanics which formed during this period [6,23].

2.2. The Drake Volcanics

The Drake Volcanics were first briefly described by Andrews [9] and much later as a sequence of volcanic rocks some 20 km wide and 60 km long trending NNW, close to the north-eastern margin of the Tablelands Complex within the SNEO and bounded to the west by the Demon Fault (Figure 2 and Figure 3) [7]. It has been shown through U-Pb zircon dating that the Drake Volcanics comprise middle Permian (~265 Ma [23]) acid to intermediate volcanics dominated by volcaniclastic andesitic units intruded by sub-volcanic andesite and rhyolite porphyries [18,24,25]. The Drake Volcanics have an estimated thickness of ~600 m in the Red Rock–White Rock–Lady Hampden area [18], while Thomson [8] estimated a maximum thickness of 900 m elsewhere.
The Drake Volcanics are calc-alkaline, conformably overlie the Razorback Creek Mudstone, and are conformably overlain by the Gilgurry Mudstone (Figure 2 and Figure 3) [6,8,25,26,27]. The Carboniferous to Early Permian sedimentary Emu Creek Formations both underlie the Drake Volcanics and are fault-bounded on its eastern boundary, hosting the gold mineralization of the Tooloom and Lunatic Goldfields, while to the west, the Drake Volcanics are intruded by the Early Triassic Stanthorpe Monzonite pluton and structurally bound by the Demon Fault (Figure 2 and Figure 3) [6]. The volcanics are relatively undeformed, with only slightly inclined bedding (~20°). Some researchers have suggested that the Drake Volcanics formed during rifting or graben formation [28,29], while others have suggested that they formed due to transform/strike slip fault movement [30].
Airborne magnetic imagery revealed a 250 km2 roughly circular area of low magnetic response (the Drake Quiet Zone or DQZ) (Figure 3). This region with a diameter of approximately 20 km has been interpreted to represent a collapsed volcanic caldera structure with the circular shape and the quiet magnetic signature explained by the presence of a caldera filled with less magnetic volcanics than the surroundings [11,31]. This type of structure is a relatively common feature in a range of other epithermal Au-Ag deposits, including Creede (2488.28 metric tons Ag), Colorado, USA [32,33,34] and Round Mountain (513.21 metric tons Au), Nevada, USA [35,36,37]. As a collapsed caldera, the structure has been proposed to control the location of post-collapse intrusives, mineralization, and the extensive alteration within the region. Although Craighead and Gordon [11] believed that the flat to moderately dipping bedding planes acted as major fluid feeder structures, both field relationships and drill holes conclusively show that the mineralized veins are mostly subvertical in orientation, as expected, following epigenetic extensional faulting [37]. This is supported by the conclusions of Herbert [26], Bottomer [18], Perkins [27], and Houston [31] who all believed that structure played an important role in ore localization. Thick lobes and layers of dacitic–lithic and rhyolitic quartz–feldspar fiamme breccias typically observed throughout the area as a thick sequence have been interpreted as representing the caldera-forming eruption products [38].

3. Samples and Methodology

Prior to sample preparation for further analysis, samples were described, imaged, then made into thin sections or polished blocks, and analyzed using transmitted and reflected-light microscopy.

3.1. Sample Description

Carbonate deposition is mostly very late stage and is unrelated to the main mineralization event. The deposition was a multi-stage process involving a Ca-Mg carbonate stage, calcite forming cross-cutting veins, euhedral calcite lining cavities, late-stage calcite filling fractures, and siderite of meteoric origin on fracture surfaces of the Strauss open cut, comprising drusy euhedral crystals (to a few mm) of dark brown siderite lining fractures on altered andesite.
The carbonate samples were selected from carbonate veins from throughout the Drake Goldfield, covering a diverse range in terms of shape, thickness, and lateral continuity. The thickness of the veins ranged widely from <5 mm to >10 cm. The wide range in vein morphologies included one massive vein with straight and sharp boundaries (Figure 4b,e), stockwork veins (Figure 4a), breccia with carbonate-fill (Figure 4d) and sub-parallel veins (Figure 4b), vuggy infill with euhedral carbonate crystals (mostly calcite, Figure 4f), carbonate spots (Figure 4c), and zonation within individual veins occurred in some places (Figure 4e).
The principal occurrence of carbonates within the Drake Goldfield are as follows: (1) replacement/overprinting of earlier rock-forming minerals (e.g., plagioclase) in volcanic tuffs (Figure 5a,d); (2) vein infill as a main component of the carbonate ± quartz stage (Figure 5b,c,f). The vein carbonates are dominantly composed of fine to coarsely crystalline calcite or dolomite (Figure 5b,c,f). Three calcite generations were found within the veins: (1) fine-grained granular-textured calcite veinlets; (2) calcite replacing earlier coarse-grained dolomite; and (3) massive veins comprising coarse euhedral calcite crystals. Ankerite is rare, occurring as fine-grained veinlets, mostly replacing/overprinting calcite (Figure 5e). Dolomite occurs as massive veins and is locally banded with magnesite (Figure 4e) or calcite. Magnesite is more massive and coarsely crystalline, and interbanded with calcite/dolomite (Figure 4a,b). Siderite was observed as euhedral crystals on fracture surfaces within the Strauss open-cut or rimming/cross-cutting earlier magnesite in drill cores.
Sulfides were chosen from the vein and veinlet ores of the Drake Goldfield. Pyrite, sphalerite, and chalcopyrite are by far the most dominant sulfide species and were thus chosen for the sulfur isotope analyses, mainly occurring as sulfide ± quartz veins (Figure 6). Figure 6a–i show the main sulfide assemblages from the Drake Goldfield, including Ccp + Py + Qtz veins (Figure 6a), Py + Sph + Qtz veins (Figure 6b), Py veins, low-Fe Sph + Ccp + Qtz veins (Figure 6d,f), and low-Fe Sph + Ccp + Gn + Py + Qtz veins (Figure 6e).
The overall mineralogy for the deposits within the Drake Goldfield is relatively simple with the main ore comprising pyrite (some with micro to submicron inclusions of gold), with lesser sphalerite, chalcopyrite, and galena; minor electrum/gold and Ag minerals; along with supergene covellite and chalcocite–digenite. A variety of Ag minerals were identified, primarily of the tetrahedrite–tennantite group, along with minor occurrences of polybasite–pearceite species, and trace amounts of acanthite, native Ag, stephanite, pyrargyrite, and jalpaite. There are at least three generations of pyrite in the Drake Goldfield (Figure 6a–i): as small subhedral–euhedral disseminated pyrite in the host rocks (Figure 6c,h) (pyrite I); as euhedral coarse-grained pyrite (pyrite II, Figure 6c,g,i), and as framboidal-like or spheroidal textured pyrite (pyrite III, Figure 6h). Pyrite I formed before the main mineralization stage, while pyrites II and III formed during the main mineralization stage. Chalcopyrite occurs in two different generations, as exsolutions within sphalerite (Figure 6i) or as coarse grains rimming sphalerite or pyrite (Figure 6d,g). Sphalerite is often rimming pyrite (Figure 6g,i). Galena commonly exhibits mutually embayed grain boundaries with sphalerite suggesting co-crystallization (Figure 6i). For a detailed paragenetic sequence, please refer to the electronic Supplementary Material Figures S5–S7.

3.2. Sample Preparation

Fresh carbonate vein minerals were sampled from diamond drillcores covering a wide geographical and vertical distribution from across the Drake Volcanics and selected according to drill logs [6], with all either associated or spatially close to mineralization. A total of 105 samples were selected from 33 diamond drill holes from the Kylo, Lady Hampden, Silver King, All Nations, Mozart, Gladstone Hill, West Copper Deeps, Red Rock, Strauss, White Rock, and White Rock North deposits and one from the Strauss open cut. These deposits are distributed across the entire Drake Goldfield (Figure 2 and Figures S1–S4 in the Supplementary Material).
A further 53 sulfide samples from the main mineralization stage (mainly coarse-grained pyrite and sphalerite along with minor chalcopyrite and galena) were selected from across the Drake Goldfield (Figures S1–S4 in Supplementary Material) to investigate any spatial variation, and samples were also selected from various depths to investigate any vertical variation. Where possible, only coarse-grained and pure separates of sphalerite and pyrite were chosen.
The samples were then collected as ground powders using a tungsten-tipped micro-drill. They were then placed in small sample holders before being ground to <50 µm using an agate mortar and pestle to ensure compositional and grain-size homogeneity. After each sample was collected, the micro drill head was thoroughly cleaned with 85% acetone solution.

3.3. Portable X-Ray Diffraction

Carbonate species identification was made using an Olympus Terra 6400 portable X-ray diffraction (pXRD) (Olympus, Melburne, Australia) analyzer at the University of New South Wales, Sydney, Australia. In total, 15 mg of the sample with a uniform grain size <150 µm were loaded into the vibrating sample holder (VSH), which vibrated the sample without macroscopic movement of the holder [39]. This exposed crystallites in each sample to the X-ray beam at random orientations, thus helping to reduce orientation effects and allowing for superior analytical statistics [39]. The VSH was then introduced into the Olympus Terra pXRD, and the sample was subjected to XRD analysis using CoKα radiation, with a data collection range of 5 to 55° 2θ and an increment of 0.25°. Collected data were then analyzed and interpreted using HighScore Plus and SiroquantTM (v.4) commercial interpretation software. Phase identification was conducted using HighScore Plus, while SiroquantTM was used to estimate the quantitative abundance of mineral phases within the sample [40]. For full details on the use of pXRD for quantitative analysis of epithermal assemblages, see Burkett et al. [41].

3.4. Micro X-Ray Fluorescence

Micro X-ray fluorescence (μ-XRF) mapping of two carbonate veins, one from each of the All Nations and Kylo deposits, containing magnesite and calcite, was conducted using a Bruker M4 Tornado μ-XRF at the X-ray Fluorescence Laboratory, University of New South Wales, Sydney, Australia. The instrument is equipped with a Rh tube with a maximum power of 30 W, a tube voltage of 45 kV, a current of 600 μA, and a 25 μm spot size in order to achieve a high spatial resolution. One silicon drift detector was used with an active area of 30 mm2. For element-mapping analysis, spectra were acquired every 50 μm with an acquisition time set at 120 ms per pixel. All analyses were performed under vacuum (20 mbar) to facilitate the detection of light elements. Data analysis was performed in standardless mode using the Bruker M4 TORNADO built-in software FP MQuant.

3.5. Stable Isotope Measurements

Carbon and oxygen isotope analyses were carried out in the Bioanalytical Mass Spectrometry Facility, Mark Wainwright Analytical Centre, University of New South Wales, while sulfur isotopes were determined at the Isotope Ratio Mass Spectrometry facility, Central Science Laboratory, University of Tasmania.
For carbon and oxygen isotope analysis, 40 to 60 µg of carbonates were weighed out for each sample and then analyzed using a MAT 253 isotope ratio mass spectrometer equipped with a Kiel IV carbonate device (Thermo Fisher Scientific, Bremen, Germany). The carbonates were reacted with 2 drops of phosphoric acid at 70 °C to release CO2. The reacted gasses were then subjected to cryogenic removal of water vapor and other non-condensable gasses, and the pure CO2 was measured against a reference gas. For sulfur isotope determination, between 0.4 and 1.0 mg of finely powdered mineral sample was weighed in tin capsules and analyzed using flash combustion isotope ratio mass spectrometry on a varioPYRO cube coupled to an Isoprime100 mass spectrometer (Elementar Analysensysteme, Langenselbold, Germany). The SO2 produced during combustion is separated from other reaction products in a trap, then fed into the mass spectrometer and measured against a reference gas.
Stable isotope abundances are reported in delta (δ) values as the deviations from conventional standards in parts per mil (‰) using the following equations:
δHI (‰) = [(Rsample/Rstandard − 1) × 1000],
with HI = 13C, 18O, or 34S and R = 13C/12C, 18O/16O, or 34S/32S
δ13C values are expressed as relative to VPDB (Vienna Pee Dee Belemnite). δ18O can be standardized against either the VPDB or VSMOW (Vienna Standard Mean Ocean Water) scale. For the present work, the VPDB scale was adopted. The following equation was used for conversion [42] between VPDB and VSMOW scales:
δ18OVSMOW = 1.3086 δ18OVPDB + 30.86
δ 34S values are reported relative to CDT (Canyon Diablo Troilite, a meteorite which created the Barringer crater, Arizona, USA).
International reference standards with known isotopic compositions, sulfur: IAEA-S-1 (−0.3‰), IAEA-S-2 (22.66‰), IAEA-S-3 (−32.3‰), IAEA-SO5 (0.49‰), NBS 123 (17.44‰), and NBS 127 (20.32‰); carbon and oxygen: NBS18 (−5.01 and −23.2‰ for carbon and oxygen, respectively) and NBS 19 (1.95 and 2.2‰ for carbon and oxygen, respectively), were measured for instrument calibration and quality assurance. The analytical performance of the instrumentation, drift correction, and linearity performance were calculated from the repetitive analysis of these standards. The analytical precision was 0.1‰ for both δ13C and δ18O and 0.2‰ for δ34S.

4. Results

4.1. Carbonate Species

The results of the pXRD analysis for the carbonates from the Drake Goldfield are presented in Supplementary Material Table S1. Five species of carbonates were identified, including ankerite, calcite, dolomite, magnesite, and siderite. Although commonly found in many epithermal deposits (e.g., Eroy et al. [43]), no manganese carbonates, such as kutnohorite or rhodochrosite, were found in the Drake Goldfield. The quantitative Siroquant results are also shown in Table S1.

4.2. μ-XRF

μ-XRF geochemical maps for two carbonate veins from the Drake Goldfield (Figure 7) show the relative variations in chemistry of the veins. There are relatively pure granular magnesite veins in ANDD012 123.55 (Figure 7a,b), which contain a thin prominent pure calcite vein enclosed within the magnesite (Figure 7b), indicating a clear epigenetic origin, and the calcite appears to be a late-stage cavity infill. In contrast, the carbonate veins in KYDD003 188 have a relatively complex texture (Figure 7c,d). The Mg and Ca distributions (Figure 7d) collectively show that there was a regularly changing Mg/Ca ratio during precipitation with coexisting magnesite, calcite, and magnesium-bearing calcite, indicating a syngenetic relationship. The pXRD confirmed the presence of magnesite, calcite, and magnesium-bearing calcite in these samples, consistent with the μ-XRF analysis.

4.3. Stable Isotope Results

4.3.1. Carbon and Oxygen Isotopes

The detailed carbon and oxygen isotope analyses from across the Drake Goldfield are presented in Supplementary Material Table S1. The range of C and O isotopic compositions for these carbonates are presented in Table 1.

4.3.2. Sulfur Isotopes

The results for sulfur isotopes are presented in Supplementary Material Table S2. A total of 53 sulfide samples (primarily sphalerite and pyrite) were analyzed, and these have a wide range of δ34SCDT values from −16.54 to 2.10‰. Specifically, δ34SCDT in pyrite ranges between −8.77 and +2.4‰ (mean −2.86‰—12 analyses) with one outlier of −16.33‰ from White Rock; in sphalerite between −11.50 and −0.53‰ (mean −3.62‰—23 analyses); and in a mixture of sulfides ranging from 8.38‰ to −0.64‰ (mean −4.03‰—16 analyses). To this, a single analysis each for chalcopyrite (−2.64‰) and galena (−11.57‰) are added.

5. Discussion

The study of stable isotopes can provide information regarding the diverse origins of fluids, metal sources, chemical conditions, as well as temperatures of ore formation (e.g., [Ohmoto et al. [44]; Huston [45]; Hoefs [46]). In this discussion, we compare the differences within the Drake Goldfield and attempt to explain the spatial distribution of the gold (copper) and silver-dominant deposits.

5.1. Sources of Fluid for the Carbonates

Late-stage carbonate samples from the Drake Goldfield have δ13CVPDB values ranging from −21.32 to 1.42‰, with 102 samples having a relatively limited variation between −15.06 to −5.00‰. This range for most of the analyses is consistent with carbonates from other epithermal deposits, such as the Kushikino Au-Ag deposit, Japan, and the Qiucun Au deposit, China [47,48,49], while the mean value of −8.65‰ is close to the mantle range (Figure 8) [50]. The δ13CVPDB value range for ankerite (−9.33 to −5.92‰), dolomite (13.02 to −5.15‰) and magnesite (−10.44 to −5.37‰) are relatively limited to the mantle range, while calcite (−21.32 to 1.42‰) has a much wider range, and siderite has a more negative value (−13.71‰). This suggests that carbon in ankerite, dolomite, and magnesite has a magmatic dominant source, while the carbon in calcite more likely results from significant mixing of fluids from a predominant magmatic source with those from a sedimentary organic carbon source. As the siderite occurs, lining fractures near the surface in the open cut, the negative value most likely indicates a groundwater source, which could be interpreted as meteoric water that has undergone interaction with the host andesites and organic matter as the waters percolated downwards via fracture networks from the surface. In the δ18O–δ13C diagram (Figure 9 and Supplementary Material Figures S8–S10), the analyses collectively form an overall linear trend centered on an initial magmatic source trending to a more meteoric influence on one side and a low-temperature alteration influence on the other. Moreover, no data plot along any clear trend, indicating mixing of carbon derived from contaminated sedimentary rocks or high-temperature effects. The Razorback Creek Mudstone is an underlying sedimentary host rock; however, the effects of mixing with this potential carbon source are limited.
The carbon isotopes do not show any significant variation between the Au-Cu and Ag deposits, while the oxygen isotopes are slightly different (Figure 9). The δ18O for carbonates from the Ag deposits trend to heavier values (concentrated in the range of 10 to 15‰), compared to those for the Au (Cu) deposits (concentrated in the range of 3 to 7‰). This difference also indicates that the carbon for the carbonates in the Au (Cu) deposits was dominantly from a magmatic source compared to that for the Ag deposits.
For these samples, the carbon and oxygen isotope trends are generally near horizontal in the δ18O–δ13C diagram (Figure 9). This near-horizontal distribution of carbon and oxygen isotope values may be due to two processes: (1) degassing of CO2; (2) reaction between the fluid and surrounding host rocks [52,53,54]. If the distribution of carbon and oxygen isotopes is caused by the degassing of CO2, then the hydrothermal fluid is generally dominated by H2O, and the degassing of CO2 has no significant effects on the oxygen isotope composition but has a significant effect on the carbon isotopes within the fluid [55]. Though the number of samples is limited, the silver-rich deposits have a relatively wider range of δ13C, e.g., Lady Hampden (−21.32 to −6.16‰), Silver King (−13.72 to 1.4‰), and White Rock (North) (−19.48 to −6.21‰). This likely indicates that the carbonates in the silver-rich deposits have precipitated due to hydrothermal degassing. The observation of coexisting vapor-rich and liquid-rich fluid inclusions from Red Rock [27], and the presence of direct deposition of quartz (throughout the Drake Goldfield), adularia (mostly in the silver-rich deposits), and calcite (usually platy, present in White Rock and West Copper Deeps) into open spaces or colloform banded quartz (Red Rock and Kylo) could be the result of boiling in the Drake Goldfield [54,56]. However, the gold and copper-rich deposits have a narrower range of δ13C, e.g., Kylo (−7.83 to −6.79‰), Red Rock (−14.82 to −6.05‰), Strauss (−13.72 to −5.15‰), and West Copper Deeps (−15.06 to −8.84‰). Thus, the degassing of CO2 might not be the only factor influencing the precipitation of carbonates across the Drake Goldfield. In hydrothermal fluids, the solubility of carbonates increases with temperature, decreases with pressure, and increases with decreasing solution pH [57,58]. Cooling in a closed system is unlikely to induce precipitation of carbonate from the hydrothermal fluid [51]. Although water–rock reactions involving magmatic-derived fluids and country rocks are common in epithermal deposits [59], they may have no significant effects on the carbon isotopes within the fluids [55], due to minor quantities of carbonates in the host Drake Volcanics. Therefore, the precipitation mechanism of carbonate in the gold- and copper-dominant deposits is most likely dominated by the boiling process with minor effects due to water–rock reactions.
On the δ18O versus δ13C diagram (Figure 9), data from the Gladstone Hill and Mozart deposits plot in the magma–mantle field. For the Silver King, White Rock, and White Rock North deposits, most of the points plot in the magma–mantle field with a slight shift to the negative δ18O direction, suggesting that the fluids from which these carbonates precipitated from were partly meteoric in origin. For the All Nations, Kylo, and West Copper Deeps deposits, the data suggest that the carbon was derived from a magma–mantle source with a significant influence of low-temperature alteration fluids and also exhibits some characteristics indicative of recrystallization (Figure 5a). Only a few of the analyses for the Lady Hampden and Strauss deposits plot in the magma–mantle zone, with most trending towards higher δ18O values and a few to lower values, presumably due to fluid reactions with the surrounding host rocks and minor localized mixing with meteoric water. For the Red Rock deposit, the carbon appears to have been mainly derived from a magma–mantle source but with a strong meteoric water influence.
The carbon in the carbonate veins throughout the Drake Goldfield likely originated from a magma–mantle source, subsequently mixed to varying extents with fluids linked to low-temperature hydrothermal alteration and meteoric water. Additionally, a few carbonates exhibit characteristics indicative of recrystallization. The spatial distribution of the Drake Goldfield deposits in the DQZ is shown in Figure 2, from which it can be seen that, except for the Red Rock field, all the gold- and copper-rich deposits (including Kylo, Strauss, Lady Hampden, All Nations, Gladstone Hill, and West Copper Deeps) are closer to the center of the DQZ than the silver-rich deposits (White Rock, White Rock North, Silver King, and Mozart). All the deposits in the Drake Goldfield are hosted within rhyolitic to andesitic volcanic tuffs. However, for the silver-rich deposits and the Red Rock deposit, mineralization is associated with intensive brecciation and adularia ± illite ± quartz ± pyrite alteration, characteristic of a typical rock-buffered low-sulfidation epithermal system [5,16]. In addition, the White Rock (North) deposits are hosted within a magmatic breccia, whereas the other deposits are hosted within various tuff and tuff–breccia units. On the other hand, the gold (copper)-rich deposits show characteristics of an intermediate-sulfidation epithermal system, having a common alteration assemblage comprising quartz ± kaolinite ± albite ± illite ± muscovite ± chlorite ± calcite ± siderite ± pyrite, associated with pyrite + chalcopyrite + low-Fe sphalerite + galena ± minor tennantite group minerals, tetrahedrite, and polybasite mineralization [15]. As we have discussed above, the fluids responsible for at least the carbonates in the gold and copper deposits (except for the Red Rock field) and Lady Hampden (δ18O shift towards positive values in Figure 9) show an initial magmatic origin but were then more greatly affected by fluids responsible for low-temperature alteration. The carbonate-forming fluids for the silver-rich deposits show an initial magmatic fluid origin followed by later mixing with minor meteoric water during their evolution. This is supported by a δ18O shift towards negative values, as illustrated in Figure 9 and ankerite overprinting calcite (Figure 5e). The metal zonation of the deposits within the central part of DQZ may relate to the localized distance from the fluid and/or heat source. For the Red Rock area, the carbonate-forming fluids also show a magmatic origin but have a much greater meteoric water influence. This may indicate a separate fluid and/or heat source for the Red Rock area. Although the carbon and oxygen isotope data for the carbonate veins from the White Rock and White Rock North deposits overlap with those of the Mt Carrington group silver-rich deposits, field relationships and other data (e.g., petrographic analysis, whole rock geochemistry) strongly suggest that these silver-rich deposits are from separate shallow level volcanic centers.

5.2. Sources of Sulfur

Sulfur in ore-forming fluids can originate from three main sources: (I) mantle-derived sulfur (0 ± 3‰ [60]), (II) seawater sulfur (>20‰), and (III) strongly reduced (sedimentary) sulfur (<0‰ [50]). The δ34S values for the Drake Goldfield epithermal deposits are summarized in Figure 10a and Supplementary Material Table S2, from which it can be seen that the δ34S values for the Drake Goldfield all cluster around −6 to 0‰. Pyrite is the most abundant sulfide and has a wider range in δ34S values (−16.33 to 2.4‰), while sphalerite has a narrower range in composition (−6 to 0‰). Potential fractionation of δ34S within the sulfides in the main stage is unlikely, as supported by the mutually embayed grain boundaries between galena and sphalerite (Figure 6i); similar δ34S values were found for the galena and sphalerite in the same drill hole (WRDD019 in Supplementary Materials Table S2) and in the similar values for coexisting pyrite and sphalerite (SRDD020 10 and WRDD026 78.5 in Supplementary Materials Table S2).
The similar δ³⁴S values observed for both galena and sphalerite within the same drill hole (WRDD019, Supplementary Material Table S2), as well as for coexisting pyrite and sphalerite (SRDD020 10 and WRDD026 78.5 in Supplementary Material Table S2), suggest that isotopic reversals may have occurred. Sulfide minerals generally show isotopic fractionation in the following sequence: δ³⁴S pyrite > δ³⁴S sphalerite > δ³⁴S chalcopyrite > δ³⁴S galena [44]. The slight isotopic disequilibrium observed in our study is also supported by petrographic evidence, which shows early sulfides being replaced by later generations (Figure 6g,i).
Some deposits have a narrow compositional range and therefore suggest a single source of sulfur, while other deposits (e.g., White Rock) have a wide range that extends to more negative values. The broadening of the ranges may suggest mixing with other isotopic reservoirs or the operation of redox processes during epithermal mineralization [64]. However, there is no evidence for a significant change in oxidation state observed between the sulfides. The wide range of sulfur isotopes may in part be due to thermochemical sulfate reduction, as the reduced Razorback Creek Mudstone underlies the Drake Volcanics in the study area and could act as a reductant.
Herbert and Smith [63] reported δ34S values for sulfides from 17 deposits (host rock included) from the Drake Volcanics ranging from −29.5 to 3‰ (Golden Age, Addisons, Red Rock ‘K’, Emu Creek, Lady Hampden, Mt Carrington, Kylo, Guy Bell, Pioneer, White Rock, Adeline, Mascotte, Lady Jersey, Golden Drake, Just In Time, Just In Time North, and the Drake Volcanics), most clustering between −4 and 2‰. They believed that this sulfur was derived from the local country rocks. Perkins [27] concluded that the sulfur for the Red Rock deposits was dominantly magmatic sulfur, while sulfur for the other deposits in the Drake Goldfield and the Drake Volcanics may be the result of seawater sulfur mixing with magmatic sulfur.
In this study, sulfides from the All Nations, Gladstone Hill, Kylo, Lady Hampden, Red Rock, Strauss, and West Copper Deeps have δ34S values ranging from −6.96 to +2.4‰, with 89% of these close to 0 ± 4‰ (Figure 10a,b), similar to the range reported from other low-sulfidation epithermal deposits (−6 to 5‰ [61]). These sulfides have a relatively narrow range in δ34S compositions, suggesting that they formed under stable physical and chemical conditions and were derived from a relatively homogeneous source [44]. The δ34S values for these deposits are clustered near to 0‰, suggesting that the sulfur has a magmatic origin or has been leached from the host volcanic sequence (~1‰ [63]).
In contrast, the δ34S data for Mozart, Silver King, White Rock, and White Rock North deposits show a wider and more negative range (Figure 10b) from −16.54 to −3.98‰, from which White Rock and White Rock North have a range from −16.54 to −4.4‰, similar to previously reported results (~ −15 to ~ −4‰ [63]). Variation in δ34S values is common in epithermal deposits [45,46,47], and Richards [59] summarized that the δ34S values in epithermal systems range widely from −15 to +8‰ (though mostly <0‰). Sulfur isotope values as high as −3.98‰ (close to 0 ± 3‰) suggest a magmatic sulfur source. The presence of a −16.54‰ value implies that a second source contributed minor amounts of sulfur to the Drake Goldfield and was similar to that of reduced sulfur in sedimentary rocks [50], suggesting that remobilized sulfur from a sedimentary source may have been incorporated into the hydrothermal fluids. Furthermore, remobilization of sulfur from a sedimentary source would be expected to produce a wider spread of δ34S values [65]. Given that the Razorback Creek Mudstone underlies the Drake Volcanics in the study area, magmatic and sedimentary-derived sulfur mixing is certainly possible.
Furthermore, most of the deposits which have a magmatic dominant δ34S source are gold- and copper-dominant, while the deposits that have a mixed δ34S source are silver-dominant. Importantly, there is no noticeable δ34S variation with depth either for individual deposits or across the Drake Goldfield as a whole.

6. Conclusions

Mineralization within the Drake Goldfield mainly occurs in the felsic intrusive units and andesitic/dacitic tuffs. There is a wide range of carbonate species within the Drake Goldfield, including ankerite, calcite, dolomite, magnesite, and siderite. Most of the carbonates having a vein or stockwork structure are of very late-stage origin and appear to have been the last phases to have precipitated from hydrothermal fluids across the Drake Goldfield. These carbonates are likely not related to the main mineralization event. The geochemistry and morphology show both syngenetic/epigenetic relationships between calcite, magnesium-bearing calcite, and magnesite, providing strong evidence for fluctuating Mg/Ca ratios in the fluids during carbonate precipitation and disequilibrium assemblages.
Apart from three outlier values, the carbonates are characterized by a wide range in δ18OVSMOW but a relatively narrow range of δ13CVPDB. Carbon and oxygen isotopes of vein carbonates from the Drake Goldfield deposits indicate that the carbon had a mantle-derived source and that this then underwent limited mixing with meteoric waters and fluids associated with low-temperature alteration. Deposits dominated by gold and copper (except for the Red Rock field) show an initial magmatic origin for the late-stage fluid, which then mixed with fluids associated with low-temperature alteration processes during their evolution. In contrast, for the silver dominant deposits and the Red Rock deposit, although also having an initial magmatic origin, the fluids responsible for carbonate formation were more greatly affected by mixing with meteoric water.
Sulfur isotope ratios of sulfide minerals indicate that the sulfur was most likely derived from at least two different sources: a predominantly magmatic source (especially for the gold and copper dominant deposits) and a minor reduced sedimentary source (only for the silver-dominant deposits), the relative importance of each varying between deposits.
The combined carbon, oxygen, and sulfur isotopes clearly show that although the carbonate veins were derived from a wide range of fluids, these were different to those responsible for sulfide deposition. The different carbon, oxygen, and sulfur isotope characteristics from different deposits located across the Drake Goldfield collectively suggest a major heat and metal source below the center of the DQZ, which best explains the spatial distribution of the gold (copper) and silver-dominant deposits. The data also suggest that the Red Rock area may have had a second localized heat source, though this requires further work.
This study also has implications for further exploration for Ag-Au (Cu) and low–intermediate-sulfidation epithermal deposits within the Drake Goldfield. Au (Cu) intermediate-sulfidation deposits mainly occur within the center of the DQZ, while Ag and Au low-sulfidation epithermal deposits mainly occur within the periphery. In general, the Au (Cu) deposits within the Drake Goldfield show a limited variation in values for sulfur isotopes of the main mineralization stage sulfides (mostly close to 0‰ of δ34S) and oxygen isotopes of the carbonates, indicating magmatic dominant fluid sources. In contrast, the sulfur and oxygen isotopes for the Ag deposits have larger variations, suggesting substantial mixing processes during the fluid’s evolution. The carbon and oxygen isotopes for the low-sulfidation epithermal deposits within the Drake Goldfield show an initial magmatic source which then mixed with meteoric fluids, while the intermediate-sulfidation epithermal deposits show an initial magmatic source, which then mixed with fluids associated with low-temperature alteration. In addition, the Ag low-sulfidation epithermal deposits show a greater variation in carbonate isotope values compared to the Au low-sulfidation epithermal deposits, reflecting increased fluid-mixing processes within the Ag low-sulfidation epithermal deposits.

Supplementary Materials

The following supporting information can be downloaded at: www.mdpi.com/article/10.3390/min15020134/s1, Figure S1: Location of diamond drill holes and open-cuts which were sampled for this study across the Drake Goldfield—overall distribution of the diamond drill holes in the Drake Goldfield. Note: For drill hole numbers please refer to Supplementary Tables S1 and S2 [66]; Figure S2: Location of diamond drill holes and open-cuts which were sampled for this study—zoom-in map of the White Rock and White Rock North deposits. Note: For drill hole numbers please refer to Tables S1 and S2 [66]; Figure S3: Location of diamond drill holes and open-cuts which were sampled for this study—zoom-in map of the Kylo, Strauss and All Nations deposits. Note: For drill hole numbers please refer to Tables S1 and S2 [66]; Figure S4: Location of diamond drill holes and open-cuts which were sampled for this study—zoom-in map of the Lady Hampden. Silver King and Mozart deposits. Note: For drill hole numbers please refer to Tables S1 and S2 [66]; Figure S5: Ore paragenesis for the Cu prospects within the Drake Goldfield; Figure S6: Ore paragenesis for the Au-dominant deposits within the Drake Goldfield; Figure S7: Ore paragenesis for the Ag-dominant deposits within the Drake Goldfield; Figure S8: δ18O versus δ13C diagram for carbonates from the Cu prospects [51]; Figure S9: δ18O versus δ13C diagram for carbonates from the Au-dominant deposits [51]; Figure S10: δ18O versus δ13C diagram for carbonates from the Ag-dominant deposits [51]; Table S1: Carbon and oxygen isotopic compositions of carbonates from across the Drake Goldfield. Note: Coordinates are documented using AMG AGD66 coordinate system. Abbreviations: Ank-ankerite; Dol-dolomite; Cal-calcite; Mgs-magnesite; Sd-siderite; Brt-barite; Ccp-chalcopyrite; Py-pyrite; Sp-sphalerite; Dkt- dickite; Ab-albite; Table S2: Sulfur isotope compositions of sulfides from the Drake Goldfield. Note: Coordinates are documented using AMG AGD66 coordinate system. Abbreviations: Ccp-chalcopyrite; Cv-covellite; Cc-chalcocite; Py-pyrite; Gn-galena; Sp-sphalerite.

Author Contributions

Writing—original draft preparation, H.Q.; writing—review and editing, I.G. and D.F.; formal analysis, H.Q., L.A., C.D. and H.W.; investigation, H.Q., E.M. and I.G.; supervision and writing input, R.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the PANGEA Research Centre, School of Biological, Earth and Environmental Sciences, UNSW Sydney, and White Rock Minerals Ltd. Hongyan Quan was supported in part by the China Scholarship Council (No. 201706170039).

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

We would like to thank Brett Nagel from White Rock Minerals for all his help during fieldwork and sample collection. Constructive and thorough reviews by Stuart Simmons greatly improved this manuscript. We would also like to thank all three anonymous reviewers for their insightful comments which greatly improved the revised manuscript.

Conflicts of Interest

The authors Hongyan Quan and Rohan Worland are respectively employees of BGRIMM Technology Group and White Rock Minerals Ltd. The paper reflects the views of the scientists and not the company.

References

  1. Lindgren, W. Mineral Deposits; McGraw-Hill Book Company Inc.: New York, NY, USA, 1933; p. 930. [Google Scholar]
  2. White, N.C.; Hedenquist, J. Epithermal gold deposits. Styles, characteristics and exploration. SEG Newsl. 1995, 27, 1–13. [Google Scholar] [CrossRef]
  3. Simmons, S.F.; White, N.C.; John, D.A. Geological characteristics of epithermal precious and base metal deposits. In 100th Anniversary Volume; Society of Economic Geologists: Littleton, CO, USA, 2005; pp. 485–522. [Google Scholar]
  4. Sillitoe, R.H. Epithermal paleosurfaces. Miner. Depos. 2015, 50, 767–793. [Google Scholar] [CrossRef]
  5. Hedenquist, J.W.; Arribas, A.; Gonzalez-Urien, E. Exploration for epithermal gold deposits. In Reviews in Economic Geology; Society of Economic Geologists: Littleton, CO, USA, 2000; Volume 13, pp. 245–277. [Google Scholar]
  6. White Rock Minerals. Mt Carrington Project: Overview, Geological Setting, Resources, Development, Exploration Gold and Silver, and Exploration Copper. Available online: https://www.whiterockminerals.com.au/mt-carrington-overview (accessed on 7 August 2019).
  7. Murray, C. Tectonic evolution and metallogenesis of the New England Orogen. In New England Orogen-Tectonics and Metallogenesis; Kleeman, J.D., Ed.; University of New England: Armidale, Australia, 1988; pp. 204–210. [Google Scholar]
  8. Thomson, J. Geology of the Drake 1: 100,000 Sheet; 0724006699; Geological Survey of New South Wales, Department of Mines: Sydney, NSW, Australia, 1976; p. 202. [Google Scholar]
  9. Andrews, E.C. Report on the Drake Gold and Copper Field by EC Andrews; Geological Survey; Department of Mines and Agriculture: Perth, WA, Australia, 1908; p. 46. [Google Scholar]
  10. Perkins, C. Mineralization in the Drake Volcanics. New England Orogen Tectonics and Metallogenesis; University of New England: Armidale, NSW, Australia, 1988; p. 275. [Google Scholar]
  11. Craighead, G.; Gordon, M. White Rock Minerals (WRM): Low Cost Gold/Silver Start-Up Opportunity, Sydney; Breakaway Investment Group: St. Augustine, FL, USA, 2016; pp. 6–19. [Google Scholar]
  12. Chomiszak, G. Relationship Between Alteration Assemblages, Their Intensity and Mineralisation and Grade at the White Rock Epithermal Ag Deposit, Drake, Northeastern NSW. Unpublished Honours Thesis, University of New South Wales, Sydney, Australia, 2016. [Google Scholar]
  13. Zhang, H. Relationship Between Alteration Assemblages, Alteration Intensity and Mineralisation and Grade for the Lady Hampden Epithermal Au-Ag Deposit, Drake, North-Eastern NSW. Unpublished Honours Thesis, University of New South Wales, Sydney, Australia, 2016. [Google Scholar]
  14. White, V. Volcanic Facies and Its Relationship to Silver Mineralisation, White Rock and White Rock North Epithermal Deposits, Drake Goldfield, NE NSW. Unpublished Honours Thesis, University of New South Wales, Sydney, Australia, 2017. [Google Scholar]
  15. Madayag, E. Controls on Mineralisation and Alteration in the Epithermal Strauss Au-Ag Deposit, Drake Goldfield, Northern NSW. Unpublished Honours Thesis, University of New South Wales, Sydney, Australia, 2020. [Google Scholar]
  16. Lay, A. A Comparative Study of the Mineralogy and Geochemistry of Ore Minerals from Silver-Rich Polymetallic Deposits of the Lachlan and Southern New England Orogens, New South Wales, Australia. Unpublished Ph.D. Thesis, University of New South Wales, Sydney, Australia, 2019. [Google Scholar]
  17. Davies, B. Mt Carrington District: Structural Framework & Mineralisation; Confidential company report prepared; Rex Minerals Ltd.: Pine Point, Australia, 2010; p. 44. [Google Scholar]
  18. Bottomer, L. Epithermal silver-gold mineralization in the Drake area, northeastern New South Wales. Aust. J. Earth Sci. 1986, 33, 457–473. [Google Scholar] [CrossRef]
  19. Leitch, E. The geological development of the southern part of the New England Fold Belt. J. Geol. Soc. Aust. 1974, 21, 133–156. [Google Scholar] [CrossRef]
  20. Jessop, K.; Daczko, N.; Piazolo, S. Tectonic cycles of the New England Orogen, eastern Australia: A review. Aust. J. Earth Sci. 2019, 66, 459–496. [Google Scholar] [CrossRef]
  21. Cawood, P.; Pisarevsky, S.A.; Leitch, E. Unraveling the New England orocline, east Gondwana accretionary margin. Tectonics 2011, 30, 1–15. [Google Scholar] [CrossRef]
  22. Shaw, S.; Flood, R. The New England Batholith, eastern Australia: Geochemical variations in time and space. J. Geophys. Res. Solid Earth 1981, 86, 10530–10544. [Google Scholar] [CrossRef]
  23. Waltenberg, K.; Blevin, P.; Bull, K.; Cronin, D.; Armistead, S. New SHRIMP U–Pb zircon Ages from the Lachlan Orogen and the New England Orogen; Geoscience Australia: Canberra, Australia, 2015; pp. 64–75. [Google Scholar]
  24. Herbert, H. The Drake mineral field—An unique entity in eastern Australia. In Proceedings of the In Permian Geology of Queensland, Geological Society of Autralia, Queesland Diveision, Brisbane, Australia, 14–16 July 1982; pp. 367–377. [Google Scholar]
  25. Cumming, G. Geochemistry of the Drake Volcanics, Drake Region, Northern NSW: Preliminary Data Analysis; unpublished report; White Rock Minerals Ltd.: Ballarat, Australia, 2011; p. 25. [Google Scholar]
  26. Herbert, H.K. Gold-silver mineralisation within the Drake Volcanics of northeastern N.S.W. In Proceedings of the Permian Geology fo Queensland, Geological Society of Autralia, Queesland Diveision, Brisbane, Australia, 14–16 July 1982; pp. 401–412. [Google Scholar]
  27. Perkins, C. The red rock deposit: A late Permian submarine epithermal precious metal system in Northastern New South Wales. In Proceedings of the Pacific Rim 87. International Congress on the Geology, Structure, Mineralisation and Economics of Pacific Rim, Gold Coast, Australia, 26–29 August 1987; pp. 895–898. [Google Scholar]
  28. Brownlow, J. Diapirism and the development of four thermal provinces in north-eastern New South Wales. In Proceedings of the New England Geology, Voisey Symposium, Armidale, Australia, 7 October–8 November; University of New England: Armidale, NSW, Australia, 1982; pp. 229–237. [Google Scholar]
  29. Korsch, R. Early Permian Tectonic Events in the New England Orogen. New England Geology. Voisey Symposium, University of New England: Armidale, Australia, 1982; 35–42. [Google Scholar]
  30. Flood, P.; Fergusson, C. The geological development of the northern New England Province of the New England Fold Belt. In Proceedings of the 1984 Field Conference. Volcanics, Granites and Mineralisation of the Stan Thorpe-Emmaville-Drake Region. Geological Society of Australia, Queensland Division, Brisbane, Australia, 6–10 August 1984; pp. 1–19. [Google Scholar]
  31. Houston, M. The geology and mineralisation of the Drake mine area, northern NSW, Australia. In Proceedings of the New England Orogen, Eastern Australia, NEO 93, University of New England, Armidale, NSW, Australia, 2–4 February 1993; pp. 337–348. [Google Scholar]
  32. Steven, T.A.; Eaton, G.P. Environment of ore deposition in the Creede mining district, San Juan Mountains, Colorado; I, Geologic, hydrologic, and geophysical setting. Econ. Geol. 1975, 70, 1023–1037. [Google Scholar] [CrossRef]
  33. Barton, P.; Bethke, P.M.; Roedder, E. Environment of ore deposition in the Creede mining district, San Juan Mountains, Colorado; Part III, Progress toward interpretation of the chemistry of the ore-forming fluid for the OH Vein. Econ. Geol. 1977, 72, 1–24. [Google Scholar] [CrossRef]
  34. Hayba, D.O. Environment of ore deposition in the Creede mining district, San Juan Mountains, Colorado; Part V, Epithermal mineralization from fluid mixing in the OH Vein. Econ. Geol. 1997, 92, 29–44. [Google Scholar] [CrossRef]
  35. Sander, M.V.; Einaudi, M.T. Epithermal deposition of gold during transition from propylitic to potassic alteration at Round Mountain, Nevada. Econ. Geol. 1990, 85, 285–311. [Google Scholar] [CrossRef]
  36. Henry, C.D.; Elson, H.B.; McIntosh, W.C.; Heizler, M.T.; Castor, S.B. Brief duration of hydrothermal activity at Round Mountain, Nevada, determined from Ar 40/Ar 39 geochronology. Econ. Geol. 1997, 92, 807–826. [Google Scholar] [CrossRef]
  37. Rhys, D.A.; Lewis, P.; Rowland, J. Structural controls on ore localization in epithermal gold-silver deposits: A mineral systems approach. In Applied Structural Geology of Ore-Forming Hydrothermal Systems, SEG Reviews; Society of Economic Geologists: Littleton, CO, USA, 2020; Volume 21, pp. 83–145. [Google Scholar]
  38. Cumming, G. A Revised Stratigraphic Framework for Red Rock, Mozart, and White Rock, Drake Volcanics, Northern New South Wales; Unpublished report; White Rock Minerals Ltd.: Ballarat, Australia, 2013; p. 18. [Google Scholar]
  39. Sarrazin, P.; Chipera, S.; Bish, D.; Blake, D.; Vaniman, D. Vibrating sample holder for XRD analysis with minimal sample preparation. Int. Cent. Diffr. Data Adv. X-Ray Anal. 2005, 48, 156–164. [Google Scholar]
  40. Taylor, J. Computer programs for standardless quantitative analysis of minerals using the full powder diffraction profile. Powder Diffr. 1991, 6, 2–9. [Google Scholar] [CrossRef]
  41. Burkett, D.A.; Graham, I.T.; Ward, C.R. The application of portable X-ray diffraction to quantitative mineralogical analysis of hydrothermal systems. Can. Mineral. 2015, 53, 429–454. [Google Scholar] [CrossRef]
  42. Friedman, I.; O’Neil, J.R. Compilation of stable isotope fractionation factors of geochemical interest. Prof. Pap. 1977, 440, 11. [Google Scholar] [CrossRef]
  43. Leroy, J.L.; Hubé, D.; Marcoux, E. Episodic deposition of Mn minerals in cockade breccia structures in three low-sulfidation epithermal deposits: A mineral stratigraphy and fluid-inclusion approach. Can. Mineral. 2000, 38, 1125–1136. [Google Scholar] [CrossRef]
  44. Ohmoto, H.; Rye, R.; Barnes, H. Geochemistry of Hydrothermal Ore Deposits; John Wiley & Sons: New York, NY, USA, 1979; pp. 517–611. [Google Scholar]
  45. Huston, D. Stable isotopes and their signifi cance for understanding the genesis of volcanic-associated massive sulfide deposits: A review. In Volcanic-Associated Massive Sulfide Deposits: Processes and Examples in Modern and Ancient Settings; Society of Economic Geologists: Littleton, CO, USA, 1999; pp. 157–179. [Google Scholar]
  46. Hoefs, J. Isotope fractionation processes of selected elements. In Stable Isotope Geochemistry; Springer: Berlin/Heidelberg, Germany, 2018; pp. 53–227. [Google Scholar]
  47. Rye, R.O.; Ohmoto, H. Sulfur and carbon isotopes and ore genesis: A review. Econ. Geol. 1974, 69, 826–842. [Google Scholar] [CrossRef]
  48. Matsuhisa, Y.; Morishita, Y.; Sato, T. Oxygen and carbon isotope variations in gold-bearing hydrothermal veins in the Kushikino mining area, southern Kyushu, Japan. Econ. Geol. 1985, 80, 283–293. [Google Scholar] [CrossRef]
  49. Ma, Y.; Jiang, S.-Y.; Frimmel, H.E. Metallogeny of the Late Jurassic Qiucun epithermal gold deposit in southeastern China: Constraints from geochronology, fluid inclusions, and HOC-Pb isotopes. Ore Geol. Rev. 2022, 142, 104688. [Google Scholar] [CrossRef]
  50. Rollinson, H.R. Using Geochemical Data: Evaluation, Presentation, Interpretation; Routledge: London, UK, 2014. [Google Scholar]
  51. Liu, J. Basin fluid genetic model of sediment-hosted micro-disseminated gold deposits in the gold-triangle area between Guizhou, Guangxi and Yunnan. Acta Miner. Sin. 1997, 17, 448–456. [Google Scholar]
  52. Zheng, Y.-F. Carbon-oxygen isotopic covariation in hydrothermal calcite during degassing of CO2. Miner. Depos. 1990, 25, 246–250. [Google Scholar] [CrossRef]
  53. Zheng, Y.-F.; Hoefs, J. Carbon and oxygen isotopic covariations in hydrothermal calcites. Miner. Depos. 1993, 28, 79–89. [Google Scholar] [CrossRef]
  54. Simmons, S.F.; Christenson, B.W. Origins of calcite in a boiling geothermal system. Am. J. Sci. 1994, 294, 361–400. [Google Scholar] [CrossRef]
  55. Zheng, Y. The modeling of stable isotopic system and application for ore deposit geochemistry. Miner. Depos. 2001, 20, 246–250. [Google Scholar]
  56. Simmons, S.; Browne, P. Mineralogical indicators of boiling in two low sulfidation epithermal environments: The Broadlands-Ohaaki and Waiotapu geothermal systems. In Geology and Ore Deposits; Geological Society of Nevada: Reno, NV, USA, 2000; pp. 683–690. [Google Scholar]
  57. Pokrovsky, O.S.; Golubev, S.V.; Schott, J.; Castillo, A. Calcite, dolomite and magnesite dissolution kinetics in aqueous solutions at acid to circumneutral pH, 25 to 150 C and 1 to 55 atm pCO2: New constraints on CO2 sequestration in sedimentary basins. Chem. Geol. 2009, 265, 20–32. [Google Scholar] [CrossRef]
  58. Barnes, H.L. Geochemistry of Hydrothermal Ore Deposits; John Wiley & Sons: Hoboken, NJ, USA, 1997. [Google Scholar]
  59. Shanks, W.P., III. 13.3—Stable Isotope Geochemistry of Mineral Deposits; Elsevie: Amsterdam, The Netherlands, 2014; pp. 59–85. [Google Scholar]
  60. Chaussidon, M.; Lorand, J.-P. Sulphur isotope composition of orogenic spinel lherzolite massifs from Ariege (North-Eastern Pyrenees, France): An ion microprobe study. Geochim. Cosmochim. Acta 1990, 54, 2835–2846. [Google Scholar] [CrossRef]
  61. Field, C.; Fifarek, R. Light-Stable Isotope Systematics in Epithermal Systems: Reviews in Economic Geology; Society of Economic Geologists: Littleton, CO, USA, 1985; Volume 2, pp. 99–128. [Google Scholar]
  62. Arribas, A., Jr. Characteristics of high-sulfidation epithermal deposits, and their relation to magmatic fluid. Mineral. Assoc. Can. Short Course 1995, 23, 419–454. [Google Scholar]
  63. Herbert, H.; Smith, J. Sulfur isotopes and origin of some sulfide deposits, New England, Australia. Miner. Depos. 1978, 13, 51–63. [Google Scholar] [CrossRef]
  64. Richards, J.P. Alkalic-Type Gold Deposits: A Review. In MAGMAS, Fluids and Ore Deposits; Short Course; Mineralogical of Association of Canada: Victoria, BC, Canada, 1995; Volume 23, pp. 367–400. [Google Scholar]
  65. Zhou, J.; Huang, Z.; Bao, G.; Gao, J. Sources and thermo-chemical sulfate reduction for reduced sulfur in the hydrothermal fluids, southeastern SYG Pb-Zn metallogenic province, SW China. J. Earth Sci. 2013, 24, 759–771. [Google Scholar] [CrossRef]
  66. Google Earth Pro, 2021. V7.3.3. 7786. Drake, NSW Australia. Lat -28.886327°, Lon 152.382921°, 531E. Retrieved from June 2021. Available online: https://earth.google.com/web/ (accessed on 22 January 2025).
Figure 1. Simplified map showing the New England Orogen and the location of the Drake Goldfield (indicted by blue star; modified after Murray [7]).
Figure 1. Simplified map showing the New England Orogen and the location of the Drake Goldfield (indicted by blue star; modified after Murray [7]).
Minerals 15 00134 g001
Figure 2. Simplified geological map of the Drake Goldfield showing the location of deposits and prospects within the Drake Volcanics (All Nations, Gladstone Hill, Strass, and Kylo are included in Mt Carrington, modified after White Rock Minerals [6] and Thomson [8]).
Figure 2. Simplified geological map of the Drake Goldfield showing the location of deposits and prospects within the Drake Volcanics (All Nations, Gladstone Hill, Strass, and Kylo are included in Mt Carrington, modified after White Rock Minerals [6] and Thomson [8]).
Minerals 15 00134 g002
Figure 3. Interpreted lithostratigraphic and structural map of the Drake Volcanics and surrounding lithologies developed by Davies [17] from Geological Survey of New South Wales mapping (modified from Bottomer [18]) and interpretation of magnetic data, showing the location of the Drake Quiet Zone (“DQZ”) (1—Mount Carrington Group, including All Nations, Gladstone Hill, Kylo, Strauss, West Copper Deeps, Lady Hampden, Guy Bell, Mozart and Silver King; 2—Red Rock; and 3—White Rock and White Rock North).
Figure 3. Interpreted lithostratigraphic and structural map of the Drake Volcanics and surrounding lithologies developed by Davies [17] from Geological Survey of New South Wales mapping (modified from Bottomer [18]) and interpretation of magnetic data, showing the location of the Drake Quiet Zone (“DQZ”) (1—Mount Carrington Group, including All Nations, Gladstone Hill, Kylo, Strauss, West Copper Deeps, Lady Hampden, Guy Bell, Mozart and Silver King; 2—Red Rock; and 3—White Rock and White Rock North).
Minerals 15 00134 g003
Figure 4. Images of drill cores with carbonate veins from the Drake Goldfield: (a) sub-parallel calcite-magnesite veins in dacitic tuff from All Nations; (b) sub-parallel late-stage magnesite–calcite vein-cutting dacitic tuff from Kylo; (c) fine-grained calcite spots in dacitic tuff from Red Rock; (d) hydrothermal breccia with calcite infill from Silver King; (e) zoned dolomite and magnesite vein in dacitic tuff from Strauss; (f) euhedral calcite crystals in dacitic tuff from West Copper Deeps. Abbreviations: Dol—dolomite; Cal—calcite; Mgs—magnesite.
Figure 4. Images of drill cores with carbonate veins from the Drake Goldfield: (a) sub-parallel calcite-magnesite veins in dacitic tuff from All Nations; (b) sub-parallel late-stage magnesite–calcite vein-cutting dacitic tuff from Kylo; (c) fine-grained calcite spots in dacitic tuff from Red Rock; (d) hydrothermal breccia with calcite infill from Silver King; (e) zoned dolomite and magnesite vein in dacitic tuff from Strauss; (f) euhedral calcite crystals in dacitic tuff from West Copper Deeps. Abbreviations: Dol—dolomite; Cal—calcite; Mgs—magnesite.
Minerals 15 00134 g004
Figure 5. Typical textures and overprinting relationships of carbonates from across the Drake Goldfield: (a) recrystallized magnesite-overprinting dacitic tuff from All Nations; (b) calcite vein from Kylo; (c) calcite vein from Lady Hampden; (d) dolomite surrounding volcanic quartz from Red Rock; (e) ankerite overprint calcite from Silver King; (f) dolomite–quartz contact from White Rock North. Abbreviations: Dol—dolomite; Cal—calcite; Mgs—magnesite.
Figure 5. Typical textures and overprinting relationships of carbonates from across the Drake Goldfield: (a) recrystallized magnesite-overprinting dacitic tuff from All Nations; (b) calcite vein from Kylo; (c) calcite vein from Lady Hampden; (d) dolomite surrounding volcanic quartz from Red Rock; (e) ankerite overprint calcite from Silver King; (f) dolomite–quartz contact from White Rock North. Abbreviations: Dol—dolomite; Cal—calcite; Mgs—magnesite.
Minerals 15 00134 g005
Figure 6. Images of drill core samples with vein sulfides and photomicrographs of sulfides and their associations and textures in reflected light (RL) from the Drake Goldfield: (a) chalcopyrite–pyrite–quartz vein within andesitic tuff from Gladstone Hill; (b) pyrite–sphalerite–quartz vein with open space in dacitic tuff from Kylo; (c) pyrite vein within dacitic tuff from Mozart; (d) coarse-grained low-Fe sphalerite–chalcopyrite–quartz vein from Red Rock; (e) low-Fe sphalerite–chalcopyrite–galena–pyrite–quartz vein within bleached dacitic tuff from Strauss; (f) low-Fe sphalerite–chalcopyrite vein within andesitic tuff from West Copper Deeps; (g) chalcopyrite and sphalerite replacing stage II pyrite from Kylo; (h) recrystallized pyrite within host andesite from Silver King; (i) sphalerite–galena intergrowth replacing pyrite with chalcopyrite exsolved in sphalerite from White Rock North. Abbreviations: Ccp—chalcopyrite; Py—pyrite; Sph—sphalerite; Gn—galena; Qtz—quartz.
Figure 6. Images of drill core samples with vein sulfides and photomicrographs of sulfides and their associations and textures in reflected light (RL) from the Drake Goldfield: (a) chalcopyrite–pyrite–quartz vein within andesitic tuff from Gladstone Hill; (b) pyrite–sphalerite–quartz vein with open space in dacitic tuff from Kylo; (c) pyrite vein within dacitic tuff from Mozart; (d) coarse-grained low-Fe sphalerite–chalcopyrite–quartz vein from Red Rock; (e) low-Fe sphalerite–chalcopyrite–galena–pyrite–quartz vein within bleached dacitic tuff from Strauss; (f) low-Fe sphalerite–chalcopyrite vein within andesitic tuff from West Copper Deeps; (g) chalcopyrite and sphalerite replacing stage II pyrite from Kylo; (h) recrystallized pyrite within host andesite from Silver King; (i) sphalerite–galena intergrowth replacing pyrite with chalcopyrite exsolved in sphalerite from White Rock North. Abbreviations: Ccp—chalcopyrite; Py—pyrite; Sph—sphalerite; Gn—galena; Qtz—quartz.
Minerals 15 00134 g006
Figure 7. (a) Mosaic image of sample ANDD012 123.55; (b) micro-XRF mapping image of sample ANDD012 123.55 as a two-element overlay that includes Ca and Mg; (c) mosaic image of sample KYDD003 188; (d) micro-XRF mapping image of sample KYDD003 188 as a two-element overlay that includes Ca and Mg. Note: The red color depicts calcite, the purple color depicts magnesite; and red overlapping purple depicts magnesium-bearing calcite, with the intensity of each color representing the relative concentration.
Figure 7. (a) Mosaic image of sample ANDD012 123.55; (b) micro-XRF mapping image of sample ANDD012 123.55 as a two-element overlay that includes Ca and Mg; (c) mosaic image of sample KYDD003 188; (d) micro-XRF mapping image of sample KYDD003 188 as a two-element overlay that includes Ca and Mg. Note: The red color depicts calcite, the purple color depicts magnesite; and red overlapping purple depicts magnesium-bearing calcite, with the intensity of each color representing the relative concentration.
Minerals 15 00134 g007
Figure 8. Comparative diagram of δ13CVPDB reservoir and carbonates (blue indicates range for most analyses) from the Drake Goldfield [50].
Figure 8. Comparative diagram of δ13CVPDB reservoir and carbonates (blue indicates range for most analyses) from the Drake Goldfield [50].
Minerals 15 00134 g008
Figure 9. δ18O versus δ13C diagram for carbonates from the Drake Goldfield [51].
Figure 9. δ18O versus δ13C diagram for carbonates from the Drake Goldfield [51].
Minerals 15 00134 g009
Figure 10. (a) Cumulative histogram of sulfur isotope compositions of sulfide minerals from the Drake Goldfield. (b) δ34S values of sulfide minerals from the Drake Goldfield in comparison to common sources of sulfur (the ones marked with yellow stars are silver deposits, and the ones without any marks are Au deposits) [46,61,62,63].
Figure 10. (a) Cumulative histogram of sulfur isotope compositions of sulfide minerals from the Drake Goldfield. (b) δ34S values of sulfide minerals from the Drake Goldfield in comparison to common sources of sulfur (the ones marked with yellow stars are silver deposits, and the ones without any marks are Au deposits) [46,61,62,63].
Minerals 15 00134 g010
Table 1. The range in carbon and oxygen isotopic compositions of carbonates from Drake Goldfield.
Table 1. The range in carbon and oxygen isotopic compositions of carbonates from Drake Goldfield.
MineralNumber of Samplesδ13CVPDB (‰)Average Value for δ13CVSMOW (‰)δ18OVSMOW (‰)Average Value for δ18OVSMOW (‰)
Ankerite4−9.33–−5.92−7.39+5.81–+7.99+6.63
Calcite68−21.32–+1.42−9.13−0.92–+17.94+7.77
Dolomite17−13.02–−5.15−8.47+3.40–+10.59+5.47
Magnesite15−10.44–−5.37−7.99+8.15–+15.84+11.89
Siderite1−13.71-+17.11-
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Quan, H.; Graham, I.; Worland, R.; Adler, L.; Dietz, C.; Madayag, E.; Wang, H.; French, D. Interpreting the Complexity of Sulfur, Carbon, and Oxygen Isotopes from Sulfides and Carbonates in a Precious Metal Epithermal Field: Insights from the Permian Drake Epithermal Au-Ag Field of Northern New South Wales, Australia. Minerals 2025, 15, 134. https://doi.org/10.3390/min15020134

AMA Style

Quan H, Graham I, Worland R, Adler L, Dietz C, Madayag E, Wang H, French D. Interpreting the Complexity of Sulfur, Carbon, and Oxygen Isotopes from Sulfides and Carbonates in a Precious Metal Epithermal Field: Insights from the Permian Drake Epithermal Au-Ag Field of Northern New South Wales, Australia. Minerals. 2025; 15(2):134. https://doi.org/10.3390/min15020134

Chicago/Turabian Style

Quan, Hongyan, Ian Graham, Rohan Worland, Lewis Adler, Christian Dietz, Emmanuel Madayag, Huixin Wang, and David French. 2025. "Interpreting the Complexity of Sulfur, Carbon, and Oxygen Isotopes from Sulfides and Carbonates in a Precious Metal Epithermal Field: Insights from the Permian Drake Epithermal Au-Ag Field of Northern New South Wales, Australia" Minerals 15, no. 2: 134. https://doi.org/10.3390/min15020134

APA Style

Quan, H., Graham, I., Worland, R., Adler, L., Dietz, C., Madayag, E., Wang, H., & French, D. (2025). Interpreting the Complexity of Sulfur, Carbon, and Oxygen Isotopes from Sulfides and Carbonates in a Precious Metal Epithermal Field: Insights from the Permian Drake Epithermal Au-Ag Field of Northern New South Wales, Australia. Minerals, 15(2), 134. https://doi.org/10.3390/min15020134

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop