You are on page 1of 44

Ore Geology Reviews 32 (2007) 37 – 80

www.elsevier.com/locate/oregeorev

Ore controls in the Charters Towers goldfield, NE Australia:


Constraints from geological, geophysical and numerical analyses
O.P. Kreuzer a,b,⁎, T.G. Blenkinsop b , R.J. Morrison c , S.G. Peters d
a
Centre for Exploration Targeting (M006), School of Earth and Geographical Sciences, The University of Western Australia,
35 Stirling Highway, Crawley, WA 6009, Australia
b
Economic Geology Research Unit, James Cook University, Townsville, QLD 4811, Australia
c
Citigold Corporation Ltd, PO Box 414, Charters Towers, QLD 4820, Australia
d
United States Geological Survey, Reston, VA 20192, United States

Received 8 May 2006; accepted 23 November 2006


Available online 1 February 2007

Abstract

The approach taken in this paper, namely synthesising a wealth of previous information with new data and a genetic model, in
combination with integrated numerical analyses, led to new insights into the geological controls on the localisation of auriferous
veins and residual prospectivity of the Charters Towers goldfield, NE Australia. The method also has implications for the
assessment of other “mature” goldfields worldwide.
Despite a number of different ore controls having operated within the Charters Towers goldfield, the controlling factors can be
linked to a single genetic model for orogenic, granitoid-hosted lode-gold mineralisation in a brittle deformation regime (D4) of NE–
SW to NNE–SSW shortening, under conditions of supralithostatic fluid pressure and low stress difference. Spatial autocorrelation
results suggest district-scale alignment of the auriferous veins parallel to and overlapping with the ESE–WNW- to E–W-striking
Charters Towers–Ravenswood lineament, a major crustal boundary in the basement to the Ravenswood batholith. At the camp-
scale, auriferous veins have abundance and proximity relationships with NW–SE-, NNW–SSE-, NE–SW- and ENE–WSW-
oriented lineaments, suggesting that structures that controlled gold deposition in one camp did not necessarily control
mineralisation in other camps. Fractal dimensions obtained with the box-counting method range from 1.02 to 1.10, whereas veins
in the Charters Towers City camp are characterised by a significantly higher fractal dimension of 1.28. This discrepancy may be
taken to imply that most or all outcropping and near-surface deposits within the Charters Towers City camp have been found and
that new discoveries are more likely to occur at greater levels of depth, or outside the boundaries of this camp.
The new understanding has implications for the assessment of the residual prospectivity of the Charters Towers goldfield, where
large areas of prospective rock types and structures (e.g., approximately 40% of the Charters Towers–Ravenswood lineament) are
hidden under cover. This parameter space was inaccessible to the historic prospectors and has received relatively little attention
from recent explorers. The following steps are suggested for the development of a targeting strategy for lode-gold exploration in
areas of the goldfield under cover: (1) identify from geological and geophysical data the ENE–WSW (±15°) and NNW–SSE
(± 15°) striking structures and geological boundaries within a 20-km-wide corridor parallel to and centred upon the Charters
Towers–Ravenswood lineament, the potential control on gold deposit distribution at the regional- to district-scale, (2) interpret
from geological and geophysical data the distribution of pre-Middle Devonian granitoids within these areas that are the preferred
host rocks of the payable gold deposits, (3) deduce from geophysical data the ENE–WSW (±15°) and NNW–SSE (± 15°) striking
structures that cut or bound the intrusions identified in step 2, (4) locate segments along the structures identified in step 3 that

⁎ Corresponding author. Centre for Exploration Targeting (M006), School of Earth and Geographical Sciences, The University of Western Australia,
35 Stirling Highway, Crawley, WA 6009, Australia. Tel.: +61 8 6488 5807; fax: +61 8 6488 1178.
E-mail address: okreuzer@cyllene.uwa.edu.au (O.P. Kreuzer).

0169-1368/$ - see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.oregeorev.2006.12.001
38 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

deviate most from the geometry of a straight line (e.g., potential bends or splays) and/or intersect other structures or geological
contacts, or both, and (5) define and rank potential targets within the prospective areas identified in step 4 and systematically test
the best ones.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Charters Towers; Intrusion-hosted; Lode-gold; Ore controls; Fractal analysis; Spatial autocorrelation; Prospectivity

1. Introduction porphyry-style mineralisation similar to the Permo-


Carboniferous Mt Leyshon deposit (e.g., Scott and van
The Charters Towers goldfield in north Queensland, Eck, 2003).
Australia, includes hundreds of auriferous quartz-sulfide The lack of exploration success earned Charters Towers
veins (Hartley and Dash, 1993) (see Appendix A for a reputation as a mature goldfield of poor prospectivity.
Australian map grid coordinates of lode-gold deposits However, rather than an absence of additional gold
and occurrences), mainly hosted by Early Ordovician to deposits, Kreuzer and Alston (2004) argued that the dearth
Early Devonian granitoids of the Ravenswood batholith, of discovery may be linked to the lack of systematic
northern Tasman orogen (Figs. 1 and 2). Crosscutting exploration in areas away from the known gold deposits
relationships (e.g., Peters, 1987a,b; Hutton and Rienks, and occurrences and the scarcity of exploratory drill holes
1997), structural information (e.g., Kreuzer, 2004) and with greater than 100 m down-hole depth (Fig. 4). The
isotopic age data (Morrison, 1988; Perkins and potential for additional deposits to be discovered away
Kennedy, 1998; Kreuzer, 2005) indicate that veins of from the known gold occurrences might be considerable,
the Charters Towers goldfield formed during a single given that large areas of the Ravenswood batholith are
episode of gold mineralisation between Late Silurian poorly exposed, deeply weathered or under Tertiary to
and Early Devonian times, coincident with D4 of Quaternary cover. These areas have not been adequately
intrusions of the Ravenswood batholith. explored (Kreuzer and Alston, 2004), even though
Between 1872 and 1920, the Charters Towers previous interpretations of high-resolution geophysical
goldfield produced in excess of 6.6 million ounces data (Isles, 1994; King, 1998) indicate that prospective
(Moz) at a high average ore grade of 34 g/t Au, achieved rock types and structures are present. Effective exploration
through selective mining and hand sorting. The veins targeting under cover requires an understanding of the
also yielded more than 1 Moz Ag, 3700 t Pb and 1000 t controls on the location and size of gold deposit formation
Cu metal (Levingston, 1972). Approximately two-thirds at the mine to regional scale (e.g., McCuaig and Hronsky,
of the gold were extracted from two large (Brilliant 2000; Hronsky, 2004), which, at present, are poorly
∼ 2.1 Moz Au; Day Dawn-Mexican ∼ 1.6 Moz Au) and constrained for Charters Towers.
two medium-sized (Queen-Sunburst ∼0.19 Moz Au; The aim of the current study is to investigate the
Victory-Papuan ∼0.17 Moz Au) deposits, hosted by auriferous veins of the Charters Towers goldfield and their
contiguous fracture systems (Reid, 1917; Blatchford, host rocks at the mine to regional scale, integrating (1) the
1953; Levingston, 1972; Fig. 3 and Table 1). outcomes of previous studies (e.g., Peters, 1987a,b, 1990,
Most exploration campaigns in the Charters Towers 1993a,b; Peters and Golding, 1989; Hartley et al., 1989;
goldfield since 1960 mainly focused on searching for Hartley and Dash, 1993; Hartley, 1996; Hutton and Rienks,
extensions to, and repetitions of, known lode-gold 1997; Hutton et al., 1997; Hodkinson, 1998; Dominy et al.,
deposits and testing of gold occurrences in areas of 1999; Raine, 2001; Kreuzer, 2003, 2004, 2005, 2006;
outcrop (Hartley and Dash, 1993; Hutton et al., 1997; Kreuzer and Alston, 2004; Kreuzer et al., 2002; Morrison
Kreuzer and Alston, 2004). However, except for small- to et al., 2004), and (2) the extensive collective research and
medium-sized resources, such as those at Black Jack, exploration experience of the authors in the Charters
Disraeli, Great Britain, Hadleigh Castle, Stockholm and Towers goldfield with (3) new information from detailed
Warrior East (Kay, 1993; Hutton et al., 1997; Hodkinson, field and mine mapping, drill core logging, geophysical
1998), no new lode-gold deposits N 0.25 Moz have been image interpretation, spatial autocorrelation and fractal
found since the discovery of the Day-Dawn Mexican and analysis. These data are herein used to better define the
Brilliant reefs in 1879 and 1889, respectively. Regional controls on gold deposition at the mine- to regional-scale,
exploration in areas of cover has been limited to a single and to develop a new conceptual targeting model to help
programme of shallow pattern drilling (grid size approx- stimulate exploration of areas where prospective rock types
imately 800 m) by Normandy Exploration, targeting and structures are under cover.
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 39

2. Methodology

2.1. Geophysical surveys

New and previous geophysical image interpretations


(Isles, 1994; King, 1998), presented in this study,
utilized the following magnetic and gravity datasets:

(1) A ground magnetic survey by Terra Search Pty


Ltd covering the area surrounding the Hadleigh
Castle mine.
(2) High-resolution reduced-to-pole total magnetic
intensity (TMI), and first and second vertical
derivative (FVD, SVD) images of the Charters
Towers area that were generated from an aero-
magnetic survey commissioned by Charters
Towers Gold Mines Ltd (now Citigold Corporation
Ltd). The survey was flown in 1999 and 2000 by
UTS Geophysics, employing state-of-the-art
equipment. Line spacing was 50 m and flying
height 50 m over the detailed survey areas. Tie line
spacing was 500 m. Flight height had to be
increased to 300 m over built-up areas and where
landholders did not agree. The total survey length
was 4110 line-km.
(3) A mosaic of reduced-to-pole TMI, FVD and
first horizontal derivative (FHD) images cover-
ing the Lolworth–Ravenswood terrane that were
produced by Normandy Exploration Ltd (ac-
quired by Newmont Mining Corporation in
2002) from open file and proprietary aeromag-
netic surveys with line spacings generally
between 200 and 300 m. These images formed
the basis of interpretations by Isles (1994) and
King (1998).
(4) A Bouguer gravity anomaly map (based on density
2600 kg/m3) of part of the Lolworth–Ravenswood
terrane, generated from a survey commissioned by
Normandy Exploration Ltd (acquired by Newmont
Mining Corporation in 2002).

2.2. Spatial autocorrelation

Spatial autocorrelation, also known as Fry analysis


(after Fry, 1979), was originally developed as a method
for measuring strain from patterns of objects in rocks that
Fig. 1. (A) Simplified map of the Tasman orogen, eastern Australia, initially showed a random distribution. The technique of
showing major subdivisions (modified after Scheibner and Veevers,
spatial autocorrelation of point data employs a method of
2000). (B) Tectonic elements of north Queensland (modified from Hutton
et al., 1997; Scheibner and Veevers, 2000). B also illustrates the outcrop recording the distance and bearing from each point of a
and inferred subsurface distribution of Ordovician to Devonian intrusions particular dataset to every other point. For n points there
and location of Late Silurian to Early Devonian goldfields in north are n2–n spatial relationships (translations) that can be
Queensland. illustrated in graphs (Fry or translations plots) and rose
40
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80
Fig. 2. Simplified geological map of the central and western parts of the Ravenswood batholith, showing the locations of lode-gold deposits known or inferred to be of Late Silurian to Early Devonian
age and structural interpretation of the area based on geological mapping and geophysical image interpretation. Numbers refer to frequently quoted structures. Geology after Hutton and Rienks (1997);
structures and lineaments based on unpublished interpretations of geophysical data by Isles (D. Isles, Gravity Diamonds Ltd) and King (S. King, Solid Geology Ltd) and Peters (1987a), Hartley et al.
(1989) and Kreuzer (2003); gold deposit locations based on Hutton and Rienks (1997) and unpublished data by Morrison (G.W. Morrison, Klondike Exploration Services). Grid is Australian map grid
(AMG).
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 41

Fig. 3. (A) Production history of the Charters Towers goldfield over the period 1872 to 1920. (B) Contained-ounces-of-gold diagram for 35 well-
documented deposits of the Charters Towers goldfield, including the eight richest ones (i.e., 80,000 to 2,100,000 oz Au). Data source: Table 1.

diagrams (Fry, 1979; Lagarde et al., 1990; Vearncombe 2.3. Fractal analysis
and Vearncombe, 1999).
Our approach followed a method similar to that of Fractal geometry describes the shapes of irregular
Vearncombe and Vearncombe (1999), who demonstrat- objects (e.g., coastlines, clouds, mountains) that
ed the applicability of spatial autocorrelation to the cannot be adequately described by Euclidean geom-
analysis of distribution patterns of ore deposits. Fry etry. These irregular objects, know as fractals (fractus:
analysis was carried out, using the spatial autocorrela- Latin for broken, interrupted or irregular), are
tion software FryPlot (developed by T.G. Blenkinsop). essentially self-similar at all scales, meaning that any
Translations plots produced with FryPlot have no piece of the fractal design contains a miniature of the
geographic reference. Hence, only directions and entire design (e.g., Mandelbrot, 1967, 1983; Turcotte,
spacings of alignment of gold deposits and occurrences 1992). The fractal dimension (D) describes the degree
can be described, and alignments can be compared to of irregularity of a fractal, which is bound from below
structures of similar orientation. by the topological dimension and from above by the
42 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Table 1 of squares N(r) of size r, required to cover all gold


Orientation, strike length and minimum gold content of well-documented deposits is measured as a function of r:
deposits
Orientation Deposit Production Strike length
N ðrÞ / r−D ð1Þ
(i.e., vein system) (oz Au) (m)
WNW–ESE to Brilliant 2,100,000 1600 where the fractal dimension, D, is:
ENE–WSW Captain 260 550
Clarks 14,000 365
DN ðrÞ
Moonstone −D ¼ ð2Þ
Day Dawn- 1,400,000 2400 Dr
Mexican
Florence 4200 700 The version of the fractal analysis software Boxcount
Hadleigh Castle a 239,000 250
(developed by T.G. Blenkinsop) that was used in this
Madelaine 110 205
Moonstone 13,000 270 study is limited to two-dimensional spatial analysis of
Stockholm 81,000 150 outcropping and near-surface deposits. Hence, the
Union 4400 125 results of this study have no bearing on the assessment
Warrior 4400 20 of the discovery potential of extensions to known gold
Wellington 36,000 300
deposits at depth.
N–S to Black Jack 19,000 250
NNW–SSE Columbia 32,000 300
John Bull 41,000 450 2.4. Gold deposit and occurrence database
Just in Time 27,000 650
Lady Maria 56,000 750 Spatial autocorrelation and fractal analysis of veins
Maude St Ledger 3,300 400
of the Charters Towers goldfield via FryPlot and
Moonstone Cross 14,000 122
North Australian 25,000 450 Boxcount software required data input as XY coordi-
Old Queen Cross 37,000 400 nates, taken from mineral occurrence databases (see
Rainbow- 80,000 335 Appendix A) covering the Charters Towers (Hartley and
Wyndham Dash, 1993) and Ravenswood (Hartley, 1996)
St George 20,000 250
1:100,000 map sheets. These databases include Austra-
Victory 165,000 250
NE–SW Bonnie Doon 1100 61 lian map grid coordinates for most mineral occurrences
Identity 43,000 750 within the Charters Towers and Ravenswood areas
Imperial 15,000 150 gathered from field investigations and maps and reports
Queen b 185,000 1400 dating from the last century to the present. Grid
Ruby 24,000 180
NW–SE Cumberland 5000 70
coordinates given in the Charters Towers and Ravens-
Golden 28,000 650 wood mineral occurrence databases generally represent
Alexandra the surface projection of the “centre of gravity” of the
Great Britain 3000 120 recorded mineral occurrences (e.g., location of the
Swedenborg 10,500 400 deepest shaft or widest vein section along a particular
St Patrick 85,000 305
vein). Data accuracy ranges from ± 20 to ± 50 m given
Welcome 4000 700
a
that hand-held global positioning system instruments
Cummulative production for the period 1874 to 2004 (B. Duck,
were not available for field use in Australia at the time of
SMC Gold Ltd, written comm.); all other data taken from: Reid (1917)
and Levingston (1972). data collection (J.S. Hartley, written comm., 2005).
b
Includes the Sunburst (∼ 16,000 oz Au) and Golden Gate
(∼ 45,000 oz Au) deposits. 3. Geological framework

3.1. Regional and district geological setting


Euclidean dimension (e.g., Ledesert et al., 1993;
Karcz and Scholz, 2003). Veins of the Charters Towers goldfield are mainly
The box-counting method (e.g., Turcotte, 1992) was hosted by intrusions of the Ravenswood batholith, a
used to measure the degree of clustering of lode-gold composite plutonic mass in the eastern portion of the
deposits and occurrences within the Charters Towers Lolworth–Ravenswood terrane of the Thomson fold
goldfield, following the approaches of Carlson (1991), belt, northern Tasman orogen (Figs. 1 and 2).
Blenkinsop (1994, 1995) and Blenkinsop and Sander- The Tasman orogen (Fig. 1A) of eastern Australia
son (1999). In the box-counting technique, the number is subdivided into five fold belts, representing the
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80
Fig. 4. Map of the distribution of mineral occurrences (gold and base metals) in the Charters Towers district, also showing mines (gold and base metals) and locations of drill holes N100 m down-hole
depth as of May 2004 (modified from Kreuzer and Alston, 2004).

43
44 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

successive, generally westward-directed subduction and batholiths (Hutton et al., 1997; Hutton and
accretion of Neoproterozoic to Triassic terranes to the Rienks, 1997). A magmatic arc setting is implied
Precambrian Gawler and North Australian cratons. by (A) at least some Early Ordovician intrusions
Important stages of deformation and magmatism in the are co-magmatic with arc lavas of the Seventy
evolution of the Tasmanides include: the Late Cambrian Mile Range Group (Hutton and Rienks, 1997),
Delamerian orogeny, the Late Ordovician to Early (B) the common continental magmatic-arc geo-
Silurian Benambran orogeny, the Early Devonian chemical signature of mafic intrusions of the
Bowning-Bindi orogeny, the Middle Devonian Tabber- Ravenswood batholith (Hutton and Crouch,
abberan orogeny, the Middle Carboniferous Kanimblan 1993), (C) the trace element and calc-alkaline
orogeny, and the Late Permian to Middle Triassic New major-element geochemistry of granitoids of the
England orogenies. Metamorphic grades are generally Ravenswood batholith (Hutton and Crouch,
low across the Tasman orogen, with most rocks at 1993), and (D) the batholiths forming part of a
subgreenschist to greenschist conditions (e.g., Coney linear, over 700-km-long belt of Middle Silurian
et al., 1990; Walshe et al., 1995; Wellman, 1995; to Middle Devonian intrusions that is subparallel
Solomon and Groves, 2000; Scheibner and Veevers, to the eastern margin of the North Australian
2000; Gray and Foster, 2005; Glen, 2005). craton (Withnall et al., 1997).
The Lolworth–Ravenswood Terrane (e.g., Hutton et al.,
1997; Scheibner and Veevers, 2000; Fig. 1B), although 3.2. Local geology
commonly assigned to the poorly exposed Thomson fold
belt (e.g., Murray and Kirkegaard, 1978; Murray, 1986) of Intrusions of the Ravenswood batholith (Fig. 2) are
the northern Tasman orogen, is characterised by ENE– mainly oxidised (magnetite-series), biotite- and horn-
WSW-striking structures that are perpendicular to the blende-bearing I-type granitoids (Richards, 1980; Hutton
dominant NNW–SSE structure recorded elsewhere in the and Rienks, 1997; Hutton et al., 1997). Based on field
Thomson fold belt (e.g., Henderson, 1986; Wellman, and petrological observations, and geochemical and
1995; McElhinny et al., 2003). Major Early to Middle geophysical data, the intrusive history of the Ravens-
Palaeozoic elements of the Lolworth–Ravenswood terrane wood batholith has been divided into three major
include the following: episodes of magmatic activity: (1) mainly granitic,
Early to Middle Ordovician intrusions (490 ± 6 Ma to
(1) Meta-igneous and metasedimentary rocks of 463 ± 3 Ma) that were subsequently strained or recrys-
greenschist- to amphibolite-grade, such as those tallised (Tenison-Woods and Rienks, 1992; Hutton and
of the Neoproterozoic to Early Cambrian Argen- Rienks, 1997), (2) Middle Silurian to Middle Devonian
tine, Cape River, Charters Towers and Running plutons (426 ±4 Ma to 382 ± 5 Ma) that are typically
River Metamorphics. Geological and geochemi- hornblende- and biotite-bearing granodiorites and tona-
cal data are consistent with deposition of the lites, making up 60% of the surface area of the
precursor rocks to these metamorphic rocks in a Ravenswood batholith, and (3) Late Carboniferous to
rifted continental margin environment, possibly Early Permian plutonic and volcanic rocks (311 ± 3 to
related to the break-up of the supercontinent 283 ± 9 Ma) that are less voluminous (≤6% of the surface
Rodinia (Hutton et al., 1997; Withnall et al., area), cropping out primarily in the eastern and
1997). southeastern Ravenswood batholith (Hutton and Rienks,
(2) An E–W-striking belt of greenschist or lower- 1997; Hutton et al., 1997).
grade marine volcanic and clastic sedimentary The basement to the intrusions is composite (Fig. 2).
rocks known as the Seventy Mile Range Group. To the south, it consists mainly of greenschist or lower-
Stratigraphic and geochemical data indicate that grade volcanic and sedimentary rocks of the Cambrian
the tectonic setting of the Late Cambrian to Early to Ordovician Seventy Mile Range Group and the Kirk
Ordovician Seventy Mile Range Group was that River Beds (Henderson, 1986; Berry et al., 1992; Hutton
of an extensional back-arc (Henderson, 1986; et al., 1997). To the north, the basement is dominated by
Berry et al., 1992; Stolz, 1995). greenschist to amphibolite-grade metasedimentary
(3) Greenschist or lower-grade intrusive rocks of rocks of the Neoproterozoic or Early Cambrian Charters
Early to Middle Ordovician and Middle Silurian Towers Metamorphics. Meta-igneous rocks, such as the
to Middle Devonian age forming three large Bucklands Hill diorite, occur as dykes and sills that
plutonic rock masses. From west to east, these are intruded the Charters Towers Metamorphics during the
the Reedy Springs, Lolworth, and Ravenswood Middle Cambrian (Hutton and Rienks, 1997).
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 45

A study by Hutton (2004) on the geochemistry, The second deformation event (D2) produced
isotopic characteristics and zircon populations of mylonitic, NW–SE- and NE–SW-striking shear zones,
granitoids in the Lolworth–Ravenswood terrane indi- cutting pre-Middle Silurian rocks and D1 structures
cates that source rocks to intrusions of the Ravenswood (including S1 fabric deflections by D2 structures). S–C
batholith are of Late Mesoproterozoic age (1.0 to fabrics (S2) preserved in the NW–SE-striking Weir Hill
1.2 Ga), and, thus, older than the Neoproterozoic to shear zone (Fig. 2), characterised by a gently SE-
Early Cambrian basement to the Ravenswood batholith. plunging stretching lineation, and similar shear zones
Hutton (2004) also suggests that this Grenvillian-age south of Charters Towers imply dextral senses of shear,
source is unique to the Ravenswood batholith. The whereas S2 fabrics preserved by NE–SW-striking shear
Lolworth and Reedy Springs batholiths in the western zones are suggestive of sinistral movement. Given these
part of the Lolworth–Ravenswood terrane are thought to relationships, the D2 deformation event may have been
be underlain by crust that is similar to that exposed in the linked to ∼ N–S-directed regional shortening sometime
Georgetown Inlier (Fig. 1B). The Grenvillian-age crust after D1 and prior to the Middle Silurian.
underneath the Ravenswood batholith appears to be Middle to Late Silurian extension (D3) was linked to
unique and may be a remnant basement block that has either broad doming across the Ravenswood batholith or
been removed during the Neoproterozoic break-up of extension along a N–S or NE–SW axis. Such a regime is
Rodinia or a Late Mesoproterozoic mobile belt on the inferred from pluton emplacement into graben structures
flank of the North Australian craton (Hutton, 2004). and clusters of NW–SE to N–S- and NE–SW-striking
Magnetic and gravity data imply that intrusions of the swarms of mafic dykes (e.g., Middle Silurian dykes at
Ravenswood batholith continue under Cenozoic cover, Charters Towers and Hadleigh Castle), implying
forming an approximately 100-km-wide, WNW–ESE- injection of mantle-derived magma into the crust.
striking belt of plutons. These were intruded by the The fourth deformation (D4) was characterised by
Lolworth batholith, and are bounded to the north by the brittle deformation and reactivation of earlier structures
Clarke River fault and to the south by the Cape River and coincided with the formation of the Late Silurian to
Metamorphics (Hutton et al., 1997). Gravity modelling Early Devonian veins of the Charters Towers goldfield,
by Stockill and Hutton (1991) suggests that the Ravens- mainly hosted by fractures and low-displacement
wood batholith is a 5 to 6 km thick tabular body that is (commonly b1.5 m) faults. Subhorizontal extension
exposed at or close to its roof zone (Hutton and Rienks, veins, hydraulic breccia, fault separations, fault striations
1997). and graphical analysis of fault slip data suggest that gold
deposition occurred in a regime of NE–SW to NNE–
3.3. Deformation history of the Ravenswood batholith SSW shortening, under conditions of supralithostatic
fluid pressure and low stress difference (cf. Etheridge,
The Early to Middle Palaeozoic deformation history 1983; Cox et al., 1991). The proposed timing of D4 of the
of the Ravenswood batholith (Fig. 2) has been Ravenswood batholith corresponds well with the timing
previously described by Reid (1917), Peters (1987a,b, of deformation (e.g., Seventy Mile Range Group: Berry
1990), Hutton and Rienks (1997), Hutton et al. (1997) et al., 1992; Hodgkinson-Broken River fold belt:
and Kreuzer (2003, 2004). A summary is given below. Hammond, 1986; Withnall and Lang, 1993) and gold
The first deformation (D1), recorded in pre-Middle mineralisation (e.g., Bain et al., 1998) in other parts of
Silurian rocks only, produced ∼ E–W- and less common North Queensland (Fig. 5). Later deformations had little
NW–SE-striking shear zones. The largest of these effect on the pre-Middle Devonian rocks and vein
zones, the E–W-striking, anastomosing Alex Hill systems. Reverse and normal separations of veins in
shear zone (Fig. 2), is at least 50 km long and up to workings at Charters Towers and Hadleigh Castle
2 km wide. Mylonitic rocks within the Alex Hill shear commonly range between 0.1 and 1 m. Historic records
zone are characterised by a dominantly sinistral S–C (e.g., Reid, 1917) show several faults cutting the vein
fabric and gently E-plunging mineral stretching linea- systems at Charters Towers. The largest of them, the SE-
tion, implying sinistral strike-slip movement at the time dipping Caledonia fault, gives a normal separation of
of the D1 deformation event. The NW–SE-striking D1 27 m to the workings in the Caledonia, Victory and
structures have steeply plunging stretching lineations, Queen Mines.
which were taken by Hutton et al. (1997) to imply Taken as a whole, geological, structural, geochemical
reverse movement. Field and petrological relations and geophysical data suggest that, from Cambrian to
suggest that D1 was linked to NE–SW- to N–S- Carboniferous times, the Lolworth–Ravenswood ter-
orientated shortening. rane was situated within a possible back-arc magmatic
46 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Fig. 5. Schematic summary chart of tectonic events recorded by the Lolworth–Ravenswood terrane and other tectonic elements in northeast
Queensland. Data compiled from Berry et al. (1992), Bain and Draper (1997), Zucchetto et al. (1999), Fergusson et al. (2005) and Kreuzer (2003).

environment. Cycles of basin formation and deforma- 4. Veins of the Charters Towers goldfield
tion, recorded by rocks of the Seventy Mile Mount
Range (Berry et al., 1992), Ravenswood batholith 4.1. Vein and ore zone geometry
(Hutton and Rienks, 1997) and Burdekin and Drum-
mond basins (Henderson et al., 1998), suggest repeated Veins of the Charters Towers goldfield are tabular
adjustments of regional stress patterns (Kreuzer, 2004). structures, commonly ranging in thickness from 0.1 to
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 47

Fig. 6. (A) Simplified geological map of the main mining area at Charters Towers, showing the principal auriferous veins (after Kreuzer, 2006). Inset:
Lower hemisphere, equal area projection illustrating the orientation of the principal auriferous veins (after Kreuzer, 2004). (B) Plan view of the
historic workings within the main mining area at Charters Towers. Black = development; yellow = ore zones. (C) E–W section of the historic
workings. (D) N–S section of the historic workings. Grid is Australian map grid (AMG).

1 m and having strike lengths between 0.1 and 1 km. Towers area (Fig. 6) are hosted by gently to moderately
Most veins have been mined to 0.4 km below surface or (25° to 55°) N-, NE, and E-dipping structures, although
less. However, the Day Dawn-Mexican and Brilliant rare NW- and S-dipping veins were recorded by Marks
reefs are up to 4.5 m thick, have strike lengths of up to (1913) and Reid (1917). Veins of the Puzzler area (Fig. 2)
2.4 km and have been stopped to vertical depths of up to dip gently (5° to 35°) to the N and W, whereas veins at
0.9 km (Reid, 1917; Blatchford, 1953; Levingston, 1972; Hadleigh Castle (Fig. 7) are hosted by gently to
Peters, 1987a,b). Deep drilling in 1984 by a joint venture moderately (25° to 55°) SSE-, SE- or SW-dipping
of BHP Ltd and Homestake Ltd demonstrated that structures. Ore zones within the veins occur with a
auriferous veins at Charters Towers persist to vertical periodicity typically on the order of 200 to 300 m, and
depths of at least 1270 m and suggests that there is no generally extend 200 to 700 m in the down plunge
change in the makeup of veins and gold-related wall-rock direction and 70 to 300 m normal to the plunge direction
alteration over this depth interval. Veins of the Charters (Fig. 6B to D).
48 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Fig. 6 (continued ).

4.2. Host rocks 4.3. Auriferous vein types, mineralogy and paragenesis

Most veins (∼90%) are hosted in oxidised I-type Three vein types are present in the Charters Towers
granite, granodiorite or tonalite of the medium- to high-K goldfield (Kreuzer, 2006): fault-fill (∼ 89% of all veins),
calc-alkaline Hogsflesh and Millchester supersuites, and extension (∼ 8%) and stockwork-like (∼ 3%) veins.
the high-K calc-alkaline Broughton River suite (Kreuzer, Only fault-fill veins are of economic significance.
2005). These suites were previously defined by Peters Kreuzer (2006) identified four paragenetic stages in
(1987a,b) and Hutton and Crouch (1993). Veins that are the fault-fill veins (Fig. 8). Stages I and II were
hosted by dioritic to gabbroic intrusions (∼8%) and by significant in terms of vein growth, wall-rock sulfidation
meta-igneous and metasedimentary basement rocks and wall-rock alteration. Gold was introduced during
(∼2%) are less common, and only four veins in host stage III, later than the bulk of sphalerite of stage III but
rock other than granitoid produced more than 5000 oz Au simultaneous with galena and chalcopyrite. Narrow,
(Reid, 1917; Peters, 1987a,b). discontinuous calcite veins of stage IV mark the waning
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80
Fig. 7. (A) Simplified geological map of the Hadleigh Castle area, also illustrating the principal auriferous veins at the Hadleigh Castle and Robinson Crusoe mines, and the distribution of gold-related wall-rock
alteration assemblages (after Kreuzer, 2006). Grid is Australian map grid (AMG). (B) Contoured poles to auriferous vein orientations at Hadleigh Castle U/G and O/C (after Kreuzer, 2004) and Robinson Crusoe
O/C (all plots are lower hemisphere, equal area projections). (C) Cross-section at 10250 m E (local mine grid), Hadleigh Castle mine, illustrating the vein system and zones of gold-related wall-rock alteration
(modified from Hodkinson, 1998). Distribution of epidote veining and epidote–hematite–chlorite alteration inferred from reconnaissance mapping at key locations in the foot-wall of the 01 lode. Key to
abbreviations: O/C =open cut mine; U/G =underground mine (after Kreuzer, 2006).

49
50 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

of gold-related hydrothermal activity or a later unrelated pyrite of stage II. Less commonly, gold particles are
stage. sited at sulfide/sulfide and sulfide/quartz boundaries
Although the proportions of vein-forming minerals may (∼14%) or as inclusions in sphalerite (∼ 12%) and
vary from one location to another, the mineralogy of the arsenopyrite (∼ 2%). Notable amounts of free gold in
veins is simple and uniform. Auriferous veins are mainly hypogene ore were only recorded at two locations (Mary
composed of quartz (commonly 70 to 90 vol.%) and Lou reef, ∼5 km SE of Charters Towers: Marks, 1913;
sulfide minerals (commonly 10 to 20 vol.%). Pyrite, galena Day Dawn reef: Reid, 1917). Calcite and ankerite and
and sphalerite comprise more than 90% of the sulfide external fragments, such as wall rock material, each
volume. Chalcopyrite is minor, although locally it is the make up 3 to 5 vol.% of the vein filling (Kreuzer, 2006).
most abundant sulfide mineral. Arsenopyrite is common in
metasedimentary rock-hosted veins, such as the Great 4.4. Ore versus barren zones
Britain reef and the eastern extremity of the Queen reef,
whereas it is rarely recorded in the granitoid-hosted veins. Ore zones are locally very rich, with several zones
Maximum dimensions of gold particles commonly within the New Queen Cross, Talisman and Brilliant reefs
range from 40 to 80 μm. Most gold particles (∼ 70%) are yielded average ore grades of N 62 g/t Au. Ore zones are
spatially associated with galena or chalcopyrite, or both, everywhere restricted to the quartz veins, whereas grab
and attached to crystal, etch-pit and fracture surfaces of samples from zones of gold-related wall-rock alteration
typically yield assay values of less than 2 g/t Au. In contrast
to low-grade and barren sections, the ore zones are
characterised by large quantities (10 to 90 vol.%) of sulfide
minerals (heavily fractured pyrite and galena, sphalerite or
chalcopyrite, or both), well-developed modification and
deformation textures, and high ratios of comb quartz/buck
quartz, grey quartz/buck quartz and calcite or ankerite/buck
quartz (Kreuzer, 2006). Although relatively consistent
within the ore zones, gold distribution is discontinuous and
spotty (nuggetty) at the vein-scale (Kreuzer et al., 2002),
also displaying high variance (Morrison et al., 2004). It is
common for low-grade or barren sections of the veins to
pass over short distances into ore zones, and vice versa.

4.5. Gold-related hydrothermal wall-rock alteration

Veins of the Charters Towers goldfield are everywhere


surrounded by zones of gold-related wall-rock alteration
that grade, from unaltered rocks toward the veins, into
outer epidote–hematite–chlorite alteration zones and
inner sericite–ankerite alteration zones (Fig. 9). Zones
of gold-related wall-rock alteration are commonly up to
two or three times the width of the enclosed veins but can
be as wide as 100 m where they surround structurally
complex zones, containing multiple veins (Peters, 1987a,
b; Peters and Golding, 1989; Hartley et al., 1989;
Hodkinson, 1998; Kreuzer, 2006).
Distal epidote–hematite–chlorite zones can be further
subdivided into outer epidote and inner chlorite subzones;
both subzones are red and well developed in granitoid host
rocks. The epidote subzone is characterised by igneous
biotite and hornblende that were partially replaced by chlo-
rite, calcite and minor pyrite, whereas feldspar was partially
Fig. 8. Paragenetic sequence of ore, gangue and alteration minerals in replaced by sericite and calcite. Feldspar crystals were
veins of the Charters Towers goldfield (after Kreuzer, 2006). also reddened due to the introduction of microcrystalline
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 51

Fig. 9. Divisions and subdivisions of ideally zoned gold-related wall-rock alteration envelopes surrounding granitoid-hosted veins of the Charters
Towers goldfield. Graph also illustrates the stability of primary igneous minerals with decreasing distance to auriferous veins (after Kreuzer, 2006).

hematite. The epidote subzone is well exposed within the sericite, implying that they may have been more resistant to
Hadleigh Castle area (Fig. 7) where epidote veins (1) cut alteration.
Middle Silurian diorite dykes and Middle to Late Silurian Inner sericite–ankerite alteration zones are green
tourmaline veins, but nowhere cut Late Silurian to Early regardless of host rock type. Contacts with both the distal
Devonian veins of the Charters Towers goldfield or Late zones of wall-rock alteration and the enclosed veins are
Carboniferous to Early Permian dykes and intrusions, (2) generally sharp. The main distinctive criterion of the
cluster within 1 km of the gold deposits but are scarce proximal alteration zone is the replacement of all primary
elsewhere, and (3) are everywhere in direct contact with igneous constituents except quartz. Relict igneous textures
distal, but nowhere in contact with proximal, zones of gold- and pseudomorphs after primary igneous minerals,
related wall-rock alteration. Epidote minerals that sealed especially feldspar and biotite, by sericite are common.
these fractures may have formed from calcium, aluminium, Quartz is everywhere strained and locally recrystallised.
iron and silica that were liberated into the passing solutions Three divisions have been recognised: outer sericite–
through plagioclase breakdown reactions within the inner ankerite, intermediate sericite and inner sericite–pyrite
sericite–ankerite alteration zones of gold-related wall-rock subzones. The outermost sericite–ankerite subzone is
alteration (cf. Zhou et al., 2003). The chlorite subzone is green and rusty brown, and characterised by a high pro-
characterised by the incomplete replacement of hornblende portion of calcite and ankerite relative to other secondary
and biotite by chlorite, minor calcite and rare pyrite. minerals. Toward the vein, the sericite–ankerite subzone
Sericite growth was mainly restricted to plagioclase, gradationally gives way to the sericite subzone, which is
whereas potassium–feldspar crystals contain only minor mainly composed of sericite group minerals and quartz.
52 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Calcite and ankerite are less common. The innermost 4.7. Timing of gold deposition
sericite–pyrite subzone is composed of sericite group
minerals, quartz, ankerite, calcite, and abundant pyrite Hydrothermal muscovite from zones of gold-related
(max. dimension ≤3 cm) and minor galena, sphalerite, wall-rock alteration surrounding the Day Dawn and
chalcopyrite and trace gold. Grab samples typically yield Brilliant reefs yielded K–Ar (n = 4; Morrison, 1988)
assay values of less than 2 g/t Au (Kreuzer, 2006). and Ar–Ar (n = 1; Perkins and Kennedy, 1998) ages
Because of the replacement of primary igneous ranging from 416 ± 4 Ma to 397 ± 4 Ma. Argon–Ar ages
magnetite by pyrite and other secondary minerals, (n = 12) measured on hydrothermal muscovite from the
proximal zones of gold-related wallrock alteration induce 04 lode (Hadleigh Castle mine) fall within a similar
linear negative magnetic anomalies that can be detected range of 412.2 ± 2.4 Ma to 400.1 ± 4.9 Ma (Kreuzer,
by ground and high-resolution aeromagnetic surveys. 2005; Fig. 10).
However, many payable gold deposits are surrounded by
zones of wall-rock alteration that are less than 10 m wide 4.8. Relationships between gold deposition, deforma-
(e.g., Peters, 1987a,b; Hartley et al., 1989) and thus tion and plutonism
potentially undetected by aeromagnetic surveys.
Geological and geochronological data (Peters, 1987a,
4.6. Fluid characteristics b; Morrison, 1988; Hutton and Rienks, 1997; Perkins and
Kennedy, 1998; Kreuzer, 2005) indicate that gold
Trace metal values for chip channel samples from the B mineralisation postdated peak metamorphism (D1/D2)
lode at Hadleigh Castle mine indicate significant average by at least 20 my (Kreuzer, 2004) and that it overlapped in
Au (∼45,200×), Cu (∼140×), Pb (∼1,860×) and Zn space and time with the emplacement of Late Silurian to
(∼270×) enrichment of the veins with respect to average Early Devonian I-type plutons (Kreuzer, 2005; (Figs. 1, 2
Au, Cu, Pb and Zn contents of granitoids of the and 10)). Despite this association, the auriferous veins
Ravenswood batholith and elsewhere. The trace metal lack an obvious causative intrusion (e.g., Peters, 1987a,b;
data also illustrate that Au correlates well with Ag, Fe, As Peters and Golding, 1989; Dowling and Morrison, 1989)
and, to a lesser extent, with Te. Antimony correlates only and crosscutting relationships point toward the veins
weakly with gold content, whereas the correlation between having postdated the emplacement, crystallisation and
Au and W is negative. No correlation exists between Au brittle fracturing of their host intrusions (e.g., Peters,
and Mo and Au and Bi, metals that are commonly 1987a,b; Kreuzer, 2003, 2005). The similar K–Ar and
associated with Au in intrusion-related deposits (Kreuzer, Ar–Ar ages and uniformity of auriferous veins and zones
2005). Significant enrichment in Te (100 to N 1000 times of gold-related wall-rock alteration within an area of at
with respect to average Te contents of granitoids of the least 1200 km2 and to depths of over 1 km, and the lack of
Ravenswood batholith) of auriferous veins (Kreuzer, zoning of vein compositions and gold-related wall-rock
2005) may be taken as an indication of a magmatic alteration with respect to individual intrusions strongly
contribution to the ore-forming fluids (e.g., Cooke and suggest that ore components were sourced from a large
McPhail, 2001). Fluid inclusion data suggest that gold was system of energy and mass flux rather than from
precipitated from CO2-poor, low to moderately saline individual host plutons (Kreuzer, 2005). Moreover, Pb
fluids (∼5 to 11 wt.% NaCl equiv.) at temperatures isotope data rule out Pb derivation from any exposed
between ∼240 and 300 °C (Peters, 1987b; Peters and granitoids of the Ravenswood batholith (Black et al.,
Golding, 1989). A temperature of ore deposition of 310 °C 1997). Fig. 11 illustrates that any mineralisation associ-
is indicated by sphalerite–galena isotope fractionation data ated with the mainly oxidised, unfractionated Ravens-
(Peters, 1987a,b; Peters and Golding, 1989). In addition to wood batholith melts should be dominated by Cu–Au or
the more dilute fluids, Kreuzer (2003, 2005) recorded W (cf. Blevin et al., 1996) rather than Au only.
saline solutions (∼19 to 28 wt.% NaCl eq.) in quartz- and Gold deposition within the Charters Towers goldfield
sphalerite-hosted fluid inclusions that were trapped during and plutonism within the Ravenswood batholith over-
the paragenetic stage of gold deposition (i.e., stage III). lapped in time with the emplacement of the nearby
These data define an apparent mixing and cooling trend Lolworth (414 to 382 Ma K–Ar and U–Pb ages: Hutton
between a hotter saline and a cooler, more dilute fluid. et al., 1997; Hutton, 2004) and Reedy Springs batholiths
Fluid inclusion-based pressure estimates fall mainly within (410 to 403 Ma U–Pb and K–Ar ages: Hutton et al., 1997),
the range of 1.4 to 3.8 kbar, suggesting that the ore-bearing shown in Fig. 1B. There also is temporal overlap between
fluids may have been trapped at depths between 5 and gold mineralisation (based on 426 to 398 Ma K–Ar and
14 km (Kreuzer, 2005). Rb–Sr ages; Bain et al., 1998) and granitoid emplacement
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 53

Fig. 10. Temporal overlap of granitoid emplacement in the Ravenswood batholith and crystallisation of hydrothermal muscovite in alteration
envelopes of auriferous veins of the Charters Towers district (after Kreuzer, 2005).

(based on 431 to 420 Ma Rb–Sr and U–Pb ages; Withnall Kreuzer, 2005), as an indication that the auriferous
et al., 1997) in the Georgetown Inlier (Fig. 1B). veins belong to a group of granitoid-hosted lode-gold
deposits, commonly classified as orogenic. Examples
4.9. Classification of gold mineralisation in the of this group include the lode-gold deposits of the
Charters Towers goldfield Etheridge (Bain et al., 1998), Grass Valley (Johnston,
1940), Parcoy-Pataz (Haeberlin et al., 2004) and
Sillitoe (1997) and Sillitoe and Thompson (1998) Jiaodong (Qiu et al., 2002) districts.
proposed an intrusion-related origin for the auriferous
veins of the Charters Towers goldfield, whereas Groves 5. Controls on gold deposition
et al. (1998, 2003), Goldfarb et al. (1998, 2001) and
Bierlein and Crowe (2000) assigned the veins to the 5.1. Hand specimen-scale
class of orogenic gold deposits.
Kreuzer (2005) took the similarity of host rocks, Structural and petrographic analysis of barren and
ore element association, alteration assemblage, struc- auriferous vein specimens by Kreuzer (2006) indicate
tural control, and tectonic setting (see Table 4 in that pyrite is abundant only where the veins recorded
54 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Fig. 11. Conceptual diagram showing the relationship between compositions of granitoids from eastern Australia, oxidation and fractionation state
(field shown encloses approximately 3500 analyses) and main metal assemblages of related mineralisation (Blevin et al., 1996). Whole-rock
geochemistry data of intrusions of the Ravenswood batholith from Hutton and Crouch (1993) and the OZCHEM database of Geoscience Australia
(bhttp://www.ga.gov.au/gda/accessa.jspN). Diagram after Kreuzer (2005).

multiple stages of deformation-induced fault rupturing workers (e.g. , Day Dawn reef: Levingston, 1972;
and fluid penetration. Sulfide minerals and gold Peters, 1987a,b, 1990; Great Britain reef: Kay, 1992;
particles of stage III are virtually everywhere attached Ladybird reef: Rypkema, 1978; Maude St Ledger reef
to crystal, etchpit and fracture surfaces of earlier and B lode: Kreuzer, 2003, 2004). Kreuzer (2003, 2004,
sulfide minerals of stage II, in particular pyrite. This 2006) illustrated that textures and geometries of
gold-sulfide link cannot be explained by fluid mixing, auriferous veins at Charters Towers and Hadleigh Castle
which likely triggered gold deposition at the trap-scale are indicative of localised dilational deformation, and
(Kreuzer, 2006), but is in agreement with adsorption– that ore zones commonly formed where left-hand bends
reduction precipitation (chemisorption–physisorption: of ENE–WSW- (± 15°) or right-hand bends of NNW–
cf. Knipe et al., 1992; Widler and Seward, 2002), SSE- (± 15°) striking faults coincided with a marked
having operated as additional depositional process at decrease in the dip of the host structures. The association
the mm- to cm-scale. of ore zones in NNW–SSE-striking structures with
right-hand bends and in ENE–WSW-striking veins with
5.2. Mine- to camp-scale left-hand bends is consistent with releasing bends
developed within the proposed regime of NE–SW to
Ore zones within the auriferous veins are commonly NNE–SSW shortening at the time of gold mineralisa-
located at or within fault bends, jogs, intersections, tion (Kreuzer, 2003, 2004). Historic mine plans
relays and termination splays (Reid, 1917; Levingston, published by Levingston (1972) serve to illustrate that
1972; Rypkema, 1978; Peters, 1987a,b, 1990; Hodkin- a 5° to 10° change of fault orientation may have been
son, 1998; Dominy et al., 1999; Raine, 2001; Kreuzer, adequate for significant dilation to occur down dip or
2003, 2004; Fig. 12). along strike of a host structure.
A common spatial association of ore zones in the The largest and most productive auriferous vein
auriferous veins with more gently-dipping sections of systems (Brilliant, Day Dawn, Hadleigh Castle and
the host faults has been recorded by a number of Queen) are characterised by numerous diverging and

Fig. 12. Structural controls on ore deposition. (A) Splay fault-hosted veins at the contact of the Towers Hill granite with the Washington diorite (John
Bull reef, level No. 4 and crosscut 948S, Black Jack mine). Note how vein orientations are controlled by the pre-existing lithological contact
(modified after Denmead and Levingston, 1950). (B) Plan of 975 m RL, Hadleigh Castle mine, showing complex and well mineralised splay fault
terminations at the eastern and western termination of the 02 lode (modified from Hodkinson, 1998). (C) Vein depth contour plan of the Swedenborg
reef, illustrating how the plunge of the southern ore zone is controlled by a pre-existing vertical mylonitic shear zone (modified from Peters, 1987a,b).
(D) Plan of the underground workings on the Brilliant Deeps reef, indicating that the payable ore was deposited at a left-hand bend in the host
structure of the Brilliant reef. At the bend, the Brilliant structure splits into a hanging- and a foot-wall splay, hosting the Brilliant Deeps Hanging-wall
and Foot-wall reefs (modified from Reid, 1917). Middle Silurian diorite dykes in the wall rocks of the Brilliant Deeps reef suggest that this structure is
not hosted by the Millchester Creek tonalite but by an Ordovician intrusion. (E) Cross-section of the Mills Day Dawn United mine, showing splay
faults and fault relays down-dip of the Day Dawn reef, (modified from Reid, 1917). (F) Cross-section of the Great Britain mine, illustrating how the
richest ore zones (No 2 and 3 levels) are associated with a thicker, more gently-dipping segment of the quartz vein (modified from Kay, 1992).
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 55

rejoining converging splays, both along strike and especially where their dips are gentler than those of
down dip (e.g., Reid, 1917; Levingston, 1972; the master faults. Prominent bifurcations were
Hodkinson, 1998). These bifurcations generally dip reported from the Mills Day Dawn United mine,
in the direction of the master fault and strike sub- where the geometry of the Day Dawn Reef resembles
parallel to it. Diverging and rejoining converging that of linked en-echelon fractures (Reid, 1917;
splays were important sites of gold deposition, Levingston, 1972). Termination splays are known to
56 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

host several significant ore zones, such as those at the


western end of the Brilliant Deeps and the eastern
ends of the Brilliant and Queen reefs (Levingston,
1972; Peters, 1987a,b, 1990). Termination splays were
also important at Hadleigh Castle, where a proportion
of the payable gold was contained within complex
bifurcations, such as those at the western and eastern
extremities of the 02 lode (Hodkinson, 1998).
Structural analyses by Peters (1987a,b, 1990) suggest
that the location, shape and plunge of a number of ore zones
were controlled by pre-existing, steep-dipping to vertical
cross structures. This model is supported by historic mine
or vein contour plans, showing ore zones that are partially
bounded by pre-existing structures, such as shear zones,
faults or geological contacts (Fig. 12A, C). Cross structures
are also evident at Hadleigh Castle mine, where the N–S-
striking Western fault zone appears to have governed the
orientation of adjacent auriferous quartz veins on 965 m RL
(I.P. Hodkinson, written comm., 2001).
Historic mine plans and records may be taken to
imply that, locally, gold deposition occurred at vein
intersections. Examples given by Peters (1987a,b, 1990)
include ore zones, such as the Pfeiffer (Day Dawn reef)
and Victoria (Brilliant reef), that plunge along the
intersections of the Day Dawn and Talisman and the
Brilliant and Victory reefs, respectively.
Structural ground preparation was of great impor-
tance in the Charters Towers goldfield (Peters, 1993b).
Many pre-existing discontinuities were misoriented for Fig. 13. Aeromagnetic and gravity crustal boundaries in part of north
reactivation in the strain field present at the time of gold Queensland (modified from the Geoscience Australia National Geosci-
ence Datasets GIS: bhttp://www.ga.gov.au/map/national/N).
mineralisation. Reactivation angles of steeply-dipping
or vertical structures, such as the Middle Silurian diorite
dykes at Charters Towers (Fig. 6), would have been Silurian to Early Devonian gold mineralisation at
more than twice the optimal value. That a number of Charters Towers and Late Carboniferous to Early
these diorite dykes were loci for hydraulic fracturing and Permian gold precipitation at Ravenswood, ∼ 50 km to
auriferous vein formation during Late Silurian to Early the east of Charters Towers. This model is consistent
Devonian times (Kreuzer, 2004) is implied by (1) veins with the outcomes of the geophysical and numerical
in the Towers Hill area that strike parallel to diorite analyses present below.
dykes (Fig. 6), and (2) the Day Dawn (Reid, 1917;
Levingston, 1972) and Warrior (Citigold Corporation 6. Interpretations of geophysical data
Ltd, unpublished data) reefs that follow Middle Silurian
diorite dykes for up to 2 km along strike. For the purpose of this study lineaments that were
detected by aero- and ground magnetic surveys over the
5.3. District- to regional-scale Charters Towers region are taken to represent exposed or
unexposed structures, representing fractures or fracture
Hutton and Rienks (1997) proposed a model in zones, faults or fault zones, or crustal discontinuities
which an E–W-striking crustal boundary in the (cf. Chernicoff et al., 2002).
basement of the Ravenswood batholith (Fig. 13),
inferred from geochemical (Hutton et al., 1994, 1997) 6.1. Charters Towers area
and geophysical data (Geoscience Australia Geosci-
ence Datasets GIS 2002: http://www.ga.gov.au/map/ The new interpretation of aeromagnetic data of the
national/), was fundamental in the localisation of Late Charters Towers area in Fig. 14 illustrates three major
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 57

Fig. 14. Aeromagnetic interpretation (this study) of the Charters Towers area based on total magnetic intensity (TMI) and first vertical derivative
(FVD) images that are part of detailed survey of the area commissioned by Citigold Corporation Ltd. Superimposed geological boundaries taken from
Hutton and Rienks (1997); distribution of felsic and mafic dykes based on Reid (1917), Peters (1987a) and Kreuzer (2003); location of lode-gold
deposits based on Hartley and Dash (1993). Note that the main mining area in the city of Charters Towers could not be surveyed from the same height
as the surrounding area but had to be captured from much greater altitude, thus suffering from significant loss of detail (N. Storey, 2003, Citigold
Corporation Ltd, pers. comm.).
58 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

groups of lineaments, striking NW–SE to NNW–SSE AMG 7770500 and 7775000 mN. The WNW–ESE-
(group I), NE–SW to ENE–WSW (group II) and E–W striking lineaments are mainly restricted to the south-
to WNW–ESE (group III). western corner of Fig. 14, coincident with the WNW–
Lineaments of group I are interpreted as 0.5 to 5 km- ESE-striking, mylonitic Swedenborg shear zone and
spaced fracture zones that can be traced along strike for pre-Middle Silurian aplite dykes of similar orientation
2 to 20 km, or more in the case of the Merrie Monarch (Kreuzer, 2003).
fault (Peters, 1987a,b; Hartley et al., 1989). Lineaments
of group I are probably amongst the oldest structures in 6.2. Hadleigh Castle area
this area, given that they are more prominent in the pre-
Ordovician Charters Towers Metamorphics and Ordo- The new interpretation of aero- and ground magnetic
vician intrusions of the Ravenswood batholith than data covering the Hadleigh Castle area (Figs. 15 and 16)
lineaments of group II and III. Moreover, the orientation illustrates two major groups of lineaments, striking NE–
of lineaments of group I is similar to that of dykes, SW to E–W (group I) and NNW–SSE to NW–SE
bedding and fold axes that are exposed in the (group II).
metamorphic basement (i.e., Charters Towers Meta- Lineaments of group I are interpreted as 2- to 4-km-
morphics) to the Ravenswood batholith (Figs. 2 and 6). spaced and up to 20-km-long fracture zones. As
Little is known about the sense and amount of illustrated in Fig. 15, most structures of group I define
displacement along structures of group I, except where (faulted?) boundaries between the different granitoids in
rock units of the Charters Towers Metamorphics are the Hadleigh Castle area. The most prominent structure
separated by the Merrie Monarch fault in a sinistral or of group I is the Boori lineament, previously described
sinistral-oblique sense. by Tenison-Woods and Rienks (1992) as a regional-
Lineaments of group II are interpreted as 5 km- scale fault zone in the basement to the Ravenswood
spaced fracture zones that have strike lengths ranging batholith. Unlike any other lineament in the Hadleigh
from 0.1 to 15 km, or longer in case of the Stockholm– Castle area, the structure that coincides with the Boori
Gladstone lineament (Peters, 1987a,b; Hartley et al., lineament, the Basal fault zone (Kreuzer, 2003, 2004), is
1989). Most lineaments of group II appear to postdate exposed. Given that the Basal fault zone hosts the 01
those of group I, given that the NE–SW- to ENE– lode at Hadleigh Castle mine (Fig. 7), its minimum age
WSW-striking structures disrupt the continuity of, or can be constrained as Late Silurian to Early Devonian.
offset, the NW–SE- to NNW–SSE-striking structures. Hence, the Boori lineament and other lineaments of
The Stockholm–Gladstone lineament likely controlled group I to the north and south may have formed at the
the orientation of Middle Silurian diorite dykes at same time and earlier than the NW–SE-striking Rishton
Charters Towers, suggesting it may have formed at lineament (group II), which appears to have offset the
least prior to the Middle Silurian. This proposition is Boori lineament in a sinistral sense.
supported by the disrupting of the Stockholm– Ground magnetic data (Fig. 16) serve to illustrate the
Gladstone lineament and its poor definition within the intimate spatial association between veins of the Charters
Middle to Late Silurian Millchester Creek Tonalite Towers goldfield (e.g., those at Hadleigh Castle, Hadleigh
(Peters, 1987a,b; Hutton and Rienks, 1997). The East, Captain) and the Basal fault zone, which is
Stockholm–Gladstone lineament is spatially associated coincident with a 100 to 400 m wide and more than 2
with important gold deposits (e.g., Stockholm, km long negative magnetic anomaly. Geological mapping
N81,000 oz Au; Ruby, N 24,000 oz Au; Identity, (Fig. 7) indicates that this magnetic anomaly overlaps
N50,000 oz Au), implying that it was reactivated and with the distal and proximal zones of gold-related wall-
served as a fluid pathway at the time of gold rock alteration. However, the distal zone of wall-rock
deposition. There is no obvious spatial association alteration and associated magnetite replacement by pyrite
between the major gold deposits at Charters Towers and other alteration minerals (e.g., Peters, 1987a, b;
(i.e., Brilliant, Day Dawn, Mexican and Queen) and a Kreuzer, 2003, 2006) at Hadleigh Castle mine is nowhere
particular lineament other than their proximity to the more than 100 m wide (Hodkinson, 1998). Hence, the
Sandy Creek and Stockholm–Gladstone lineament width of the negative magnetic anomaly cannot be readily
(Fig. 14). explained by gold-related wall-rock alteration alone.
The few lineaments of group III are mainly in the Work by Gunn et al. (1989) at the Crossroads prospect
southern portion of the area shown in Fig. 14. (Fig. 7) suggests that the great width of the magnetic
Discontinuous, 0.3- to 3-km-long and 1- to 2-km- anomaly is related to the presence of higher than
spaced E–W-striking structures are evident between background values of kaolin, carbonate, epidote or silica,
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 59

Fig. 15. Aeromagnetic interpretation (this study) of the Hadleigh Castle area based on total magnetic intensity (TMI) and first vertical derivative (FVD)
images that are part of detailed surveys covering part of the Lolworth–Ravenswood terrane. Distribution of superimposed cover sequences and some
geological boundaries taken from Hutton and Rienks (1997); location of lode-gold deposits based on Hartley (1996).

or both. This view is supported because the negative lineaments of group II were cogenetic with the Rishton
magnetic anomaly extending into the mapped zone of lineament, their minimum age would be older than the Late
distal gold-related wall-rock alteration that is charac- Carboniferous to Early Permian age of intrusions, against
terised by a high density of epidote veins. which the Rishton lineament terminates. Additional
Lineaments of group II form mainly discontinuous, observations that may help to constrain the timing of
300-m to 4-km-spaced zones that can be traced along strike lineaments of group II include that: (1) these structures
for 500 m to 15 km. The timing of formation is poorly have the same orientation as Middle Silurian diorite dykes
constrained for most of these structures. However, if most (Kreuzer, 2003) to the north of Hadleigh Castle mine
60 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Fig. 16. Interpretation (this study) of a total magnetic intensity (TMI) image covering the Hadleigh Castle mine and its surroundings. The image is
based on a ground magnetic survey by Terra Search Pty Ltd that was commissioned by SMC Gold Ltd. Note the great width (100 to 500 m in plan
view) of the linear zone of magnetite destruction, spatially associated with the Basal fault zone.

(Fig. 7), (2) these structures are more abundant within extending into and having been reactivated in the
Ordovician plutons than younger intrusions, and (3) some basement to the basin (King, 1998).
of these structures terminate against the (faulted) margins Lineaments of group II form distinct E–W to ENE–
of Silurian to Devonian intrusions. Hence, the bulk of WSW-striking fracture zones, spaced at 10 to 20 km
lineaments of group II may be taken to have formed prior intervals and coincident with previously identified
to Late Silurian to Early Devonian times. structures, such as the Sheep Station shear zone and
the Stockholm–Gladstone, Boori and Connolly linea-
6.3. Lolworth–Ravenswood terrane ments (Peters, 1987a, b; Hartley et al., 1989). Locally,
the continuity of these up to 30-km-long structures
Interpretations of regional-scale aeromagnetic data appears to be disrupted by N–S- and NW–SE- to
by Isles (1994) and King (1998) illustrate distinct NNW–SSE-striking lineaments. Geological and geo-
patterns of lineaments (Figs. 17 and 18), mainly stri- physical criteria imply that a large proportion of the
king: (1) NW–SE (group I), (2) ENE–WSW (group II), ENE–WSW-striking structures formed prior to or at the
(3) E–W to WNW–ESE (group III), (4) NE–SW (group time of gold mineralisation. For example, the Stock-
IV), and (5) N–S (group V). holm–Gladstone lineament was intruded by Middle
Lineaments of group I occur as 5- to 10-km-spaced Silurian diorite dykes and terminates against the Middle
zones that can be traced along strike for up to 50 km. to Late Silurian Millchester Creek tonalite (Fig. 6), and
Some of these lineaments overlap with previously part of the Boori lineament (i.e., the Basal fault zone of
described NW–SE-striking structures, such as the Kreuzer, 2004) is host to the Late Silurian to Early
mylonitic, pre-Middle Silurian Weir Hill, Sellheim and Devonian 01 lode (Fig. 7).
Macrossan shear zones, and the Merrie Monarch fault Lineaments of group III are evident as WNW–ESE-
(Figs. 17 and 18), along which pre-Middle Cambrian striking, up to 30-km-long and 10- to 15-km-spaced
metamorphic basement units were separated in a structures, including the previously identified mylonitic
sinistral sense (Fig. 14; Peters, 1987a,b; Hartley et al., Oak Meadows, Swedenborg and Mt Deane shear zones
1989; Kreuzer, 2003, 2004). Given that lineaments of that are known to have formed prior to emplacement of
group I are common everywhere in the Lolworth– Middle Silurian diorite dykes and Middle Silurian to
Ravenswood terrane except for the Early Devonian Middle Devonian plutons (Peters, 1987a,b; Hartley et al.,
plutons of the Lolworth batholith and that some of these 1989; Kreuzer, 2003, 2004). The continuity of structures
structures terminate against the Early Devonian Deane of group III appears to be disrupted by N–S- and NE–
granodiorite of the Ravenswood batholith, most linea- SW-striking lineaments, such as those to the west and to
ments of group I may have formed prior to the Early the south of Charters Towers. As illustrated in Figs. 17
Devonian. The extension of a zone of NW–SE-striking and 18, a group of WNW–ESE-striking structures, here
structures that forms the boundary between the Lol- termed the Charters Towers–Ravenswood lineament,
worth and Ravenswood batholiths, here termed the pass through the main mining area at Charters Towers.
Central lineament (Fig. 17), into the area of the These structures are taken to form part of a regional-scale
Drummond basin may be explained by this lineament E–W- to ESE–WNW-striking crustal boundary or suture
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80
Fig. 17. Major structures and subsurface geology to the west of Charters Towers, previously interpreted by Hartley et al. (1989), Isles (1994) and King (1998) from detailed geophysical surveys
covering part of the Lolworth–Ravenswood terrane.

61
62
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80
Fig. 18. Grouped and colour-coded lineaments, previously interpreted by Isles (1994) and King (1998) from detailed geophysical surveys covering part of the Lolworth–Ravenswood terrane.
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 63

Fig. 19. Bouguer anomaly gravity map covering part of the Lolworth–Ravenswood terrane (source: Normandy Exploration Ltd; based on density
2600 kg/m3). Red shades indicate areas of positive gravity values produced by high average densities in the Earth's crust and upper mantle; blue
shades indicate areas of negative gravity values produced by low average densities (e.g., Kearey et al., 2002). Superimposed batholith and selected
pluton outcrop boundaries based on Hutton and Rienks (1997) and Hutton et al. (1997); lode-gold deposit locations based on G.W. Morrison (unpubl.
data).

zone (Fig. 13), previously inferred from geochemical and Based on the aeromagnetic data, lineaments of group
geophysical datasets and proposed to cut the basement of V are the least obvious structures within the Lolworth–
the Lolworth–Ravenswood terrane along a line passing Ravenswood terrane. Three broad but discontinuous
through the mining camps of Charters Towers, Hadleigh zones of N–S-striking, 10- to 20-km-spaced lineaments,
Castle and Ravenswood (Hutton and Rienks, 1997; such as the Southern Cross and Puzzler lineaments
Geoscience Australia 2005: http://www.ga.gov.au/map/ (Peters, 1987a,b; Hartley et al., 1989), are present.
national/). Hartley et al. (1989) assigned a pre-Middle Silurian age
Lineaments of group IV are common within the to the Southern Cross lineaments, based on the
Lolworth–Ravenswood terrane. They are relatively mylonitic foliation preserved in rocks that are cut by
closely spaced (5 to 10 km), also disrupting the this structure. The Puzzler lineament is evident in the
continuity of, and cutting, earlier lineaments of group field as the border between the Early to Middle
II, such as those preserved in the Cambrian to Ordovician Plumtree Creek tonalite and the Middle
Ordovician Seventy Mile Range Group to the south of Silurian to Early Devonian Puzzler granodiorite. It is
Charters Towers. It is likely that several lineaments of also evident as a boundary along which earlier NW–SE-
group IV, especially those coinciding with the NE–SW- striking structures that are preserved in the Plumtree
striking “Mt Leyshon corridor” of Morrison (1988), may Creek tonalite terminate against the Puzzler granodio-
have formed between Late Carboniferous and Early rite, which is mainly cut by N–S- to NNE–SSW-striking
Permian times. However, given that a large proportion structures. Given that locally these structures host
of the veins of the Charters Towers goldfield is hosted auriferous veins of the Charters Towers goldfield,
by NE–SW-striking faults and that earlier NE–SW- lineaments of group V may be mainly pre-Middle
striking structures could have been reactivated between Silurian structures.
Late Carboniferous and Early Permian times, it is likely The colour-shaded Bouguer gravity anomaly map
that many (or most?) lineaments of group IV were (Fig. 19) of part of the Lolworth Ravenswood terrane
formed during the pre-Middle Devonian evolution of the clearly delineates several geological features, including:
Lolworth–Ravenswood terrane. (1) the NW–SE-striking boundary between metamorphic
64 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

basement rocks of the Lolworth–Ravenswood terrane WSW- and NNW–SSE-striking structures had a
(low to high gravity values) and rocks of the North fundamental control on the spatial distribution of the
Australian craton (low gravity values) to the west, (2) the gold deposits within this area. Structures parallel to
spatial distribution of rocks of the Seventy Mile Range the alignments include the Sandy Creek and Stock-
Group within an ESE–WNW- to ENE–WSW-striking holm–Gladstone lineaments, Merrie Monarch fault,
zone of high gravity values, (3) the spatial distribution of Middle Silurian diorite dykes, geological boundaries,
rocks of the Lolworth batholith, represented by a broad, and structures preserved in the Charters Towers
elongate, E–W- to ESE–WNW-oriented zone of low Metamorphics (folds, foliations, shear zones and
gravity values (−38 mgal: Hutton et al., 1997), and (4) faults) and Early to Middle Ordovician intrusions
small, approximately circular domains (numbered 1 to 4) (narrow mylonitic shear zones).
of low gravity values within the Ravenswood batholith. The Charters Towers South data illustrate dominant
Domains 1 to 4 in Fig. 19 may be expressions of NNE–SSW to NE–SW and E–W to ENE–WSW, and
deep-seated and steep-sided rather than sheet-like minor N–S to NNW–SSE and ESE–WNW alignments
plutons (e.g., Hutton et al., 1997) or igneous bodies (Fig. 20B). Parallel structures include the Stockholm–
that have lower average densities than adjacent intru- Gladstone and Diamantina Road lineaments, Merrie
sions. Alternatively, some circular zones of low gravity Monarch fault, the Swedenborg and Oak Meadows
values (i.e., domains 1, 2 and 3) may represent tongues shear zones, Middle Silurian diorite dykes, geological
of the Lolworth batholith, extending into the area of the boundaries, and structures preserved in the Charters
Ravenswood batholith (Peters, 1987a,b). Towers Metamorphics (folds, foliations, shear zones and
Domain 1 is of particular interest for the formulation faults) and Early to Middle Ordovician intrusions
of a genetic model for veins of the Charters Towers (narrow mylonitic shear zones).
goldfield, given that this gravity anomaly is eschewed The Hadleigh Castle data illustrate a single dominant
by auriferous veins and spatially associated with the ENE–WSW alignment (Fig. 20C). This ENE–WSW
outcrop area of the Deane granodiorite (Rb–Sr whole control is not surprising given that most veins within this
rock ages of 411 ± 2 Ma to 409 ± 2 Ma: Hutton and area are spatially associated with the ENE–WSW-
Rienks, 1997; Fig. 10) that was emplaced at or close to striking Boori lineament and (sub-) parallel structures
the time of mineralisation (range of Ar–Ar and K–Ar (Fig. 15). More subtle controls on gold deposit location
ages of gold-related hydrothermal muscovite of 416 ± are masked in the rose diagram by the dominance of the
4 Ma to 397 ± 4 Ma: Morrison, 1988; Perkins and ENE–WSW direction of alignment, whereas closely
Kennedy, 1998; Kreuzer, 2005). Peters (1987a,b) spaced WNW–ESE alignments are clearly illustrated in
suggested that the distribution of the veins of the the translations plot. These are parallel to geological
Charters Towers goldfield along the margin of domain 1 contacts, Early Ordovician to Middle Silurian diorite
could be interpreted as either accidental or linked to and aplite dykes, exposed fracture zones and the Rishton
auriferous vein formation at regular distances from a lineament and (sub) parallel structures.
hidden igneous body at depth. However, the latter model Autocorrelation of the 239 gold deposits of the Charters
must remain speculative until the age of auriferous veins Towers District dataset yielded 56,882 spatial relationships,
hosted by the Deane granodiorite (see maps published in illustrating a dominant ESE–WNW to E–W trend of
Hutton and Rienks, 1997) is constrained by isotopic age alignment, and minor ENE–WSW and NNE–SSW
data. trends (Fig. 20D). The ENE–WSW and NNE–SSW
trends are well illustrated in the translations plot,
7. Spatial autocorrelation (Fry analysis) whereas the ESE–WNW to E–W pattern may be ob-
scured by the high density of point data in the centre of
7.1. Results the diagram.

The distribution pattern of the Charters Towers 7.2. Interpretation of results


City data illustrates three directions of alignment of
gold deposits and occurrences: E–W to NW–SE, Spatial autocorrelation at the district scale (Charters
ENE–WSW and NNW–SSE (Fig. 20A; Table 2). Towers District data) may be taken to imply that gold
These are: (1) consistent with the principal orienta- deposit location was controlled at the largest scale by
tions of auriferous veins at Charters Towers, striking ESE–WNW- to E–W-striking structures. The alignment
N–S to NNW–SSE, E–W and NE–SW, and (2) may at the district scale is interpreted to be linked to control
be taken to suggest that E–W- to NW–SE-, ENE– on gold deposition by structures within and connected to
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 65

Fig. 20. Fry plots and rose diagrams, illustrating the results of the spatial autocorrelation (Fry analysis) of gold deposits in the Charters Towers district.
(A) Charters Towers City dataset. (B) Charters Towers South dataset. (C) Hadleigh Castle dataset. (D) Charters Towers District dataset. (E) Overlay of
datasets (A) to (D), illustrating the relative frequency (darker colours indicate greater frequency) of alignment as revealed by spatial autocorrelation,
and the variability of alignment at the camp scale. See Table 2 for a comparison of alignment revealed by spatial autocorrelation and the orientation of
known shear zones, faults and lineaments.
66
Table 2
Results of spatial autoeorrelation (Fry analysis) applied to 235 gold deposits and occurrences
Datasets Gold Spatial Principal directions of Spacing of Parallel structures Spacing of parallel
deposits relationships alignments alignments lineaments/structures

O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80


(in km) (in km)
Charters 120 14,280 ENE–WSW ∼2.3 Narrow mylonitic shear zones, diorite dykes, geological contacts, ∼0.001 to 3
Towers Sandy Creek and Stockholm–Gladstone lineaments and (sub-)parallel structures
City E–W to WNW–ESE ∼0.7 Narrow mylonitic shear zones, diorite and aplite dykes, geological contacts, lineaments ∼0.001 to 1
NW–SE to NNW–SSE ∼2.5 structures in the metamorphic basement (folds, foliations, shear zones, faults), bb0.001 to 10
narrow mylonitic shear zones, geo logical contacts, diorite dykes, Merrie Monarch
lineament and (sub-)parallel structures
Charters 78 6,006 E–W to ENE–WSW ∼2.7 to 5.3 Structures in the metamorphic basement (folds, foliations, shear zones, faults), bb0.001 to 10
Towers narrow mylonitic shear zones, diorite and aplite dykes, geological contacts,
South Stockholm–Gladstone lineament and (sub)parallel structures
NNE–SSW to NE–SW ∼2.9 Narrow mylonitic shear zones, aplite dykes, geological contacts, ∼0.001 to 10
Diamantina Road lineament and (sub-)parallel structures
N–S to NNW–SSE ∼0.9 to 1.1 Diorite dykes, Menie Monarch lineament and (sub-)parallel structures ∼0.001 to 5
WNW–ESE nd structures in the metamorphic basement (folds, foliations, shear zones, faults), bb0.001 to 10
Swedenborg and Oak Meadows shear zone, narrow mylonitic shear zones,
igneous foliation (Broughton River granodiotite), aplite dykes, geological contacts
Hadleigh 37 1,332 ENE–WSW ∼1.7 to 2.2 Boon lineament (=Basal fault zone) and (sub-)parallel structures, geological contacts ∼2 to 7.5
Castle WNW–ESE to NE–SW ∼1.1 fractures, geological contacts, diorite and aplite dikes, ∼0.001 to 2
Rishton lineament and (sub-)parallel structures
Charters 235 56,882 E–W to ESE–WNW nd Inferred crustal boundary (Charters Towers–Ravenswood lineament) ∼0.9 to 100
Towers NNE–SSW ∼0.9 Narrow mylonitic shear zones, aplite and diorite dykes, geological contacts, lineaments ∼0.1 to 10
District ENE–WSW ∼2.9 to 5.1 Narrow mylonitic shear zones, aplite and diorite dykes, geological contacts, l ∼0.7 to 12.1
ineaments: e.g. Stockholm–Gladstone, Boon (Basal fault zone) and Connolly lineaments
nd = not discernible on Fry plot due to extremely high data density.
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 67

a previously recognised ESE–WNW- to E–W-striking relationships for box sizes of b 1.0 to 1.5 km, respec-
crustal boundary in the basement of the Lolworth– tively (Table 3). This roll-off effect is a common feature
Ravenswood terrane (Fig. 13; Wellman, 1995; Hutton of fractal analyses and was interpreted by Blenkinsop
and Rienks, 1997; Geoscience Australia, 2005: http:// and Sanderson (1999) in the context of the spatial
www.ga.gov.au/map/national/). distribution of gold deposits to reflect undersampling
Spatial autocorrelation at the camp scale raised an linked to discovery of a random sample of the total
interesting point. Rose diagrams (Fig. 20) illustrate number of deposits.
some common alignments but, taken as a whole, there The fractal dimensions of veins of the Charters Towers
are significant differences between the Charters Towers goldfield range from 1.02 to 1.10, although gold deposits
City, Charters Towers South and Hadleigh Castle assigned to the Charters Towers City subset have a
subsets (Fig. 20E). The cause(s) of these differences significantly higher D value of 1.28 (Table 3). This
are unknown. They may simply be linked to the inherent discrepancy may be interpreted in terms of: (1) most or all
natural variability of orientation of structures that gold deposits of the Charters Towers City camp having
controlled gold deposition in each of the camps. been discovered (i.e., fractal has been well sampled), or (2)
Alternatively, they could be related to camp-scale gold deposits of the Charters Towers City camp being
heterogeneities in the strain field at the time of gold hosted by structures with D values higher than those of
deposition. The variability of patterns of gold deposit structures elsewhere in the Charters Towers goldfield. Each
alignment suggests that structures that localised gold scenario has to be considered, given that there are
deposition in one camp did not necessarily control insufficient data to choose between them.
mineralisation in other camps. Proposition (1) is supported by fractal modelling
results reported by Blenkinsop and Sanderson (1999).
8. Fractal analysis One of the end-member models of gold deposit
distribution tested by Blenkinsop and Sanderson
8.1. Results (1999) is applicable to the Charters Towers goldfield
scenario: a district containing well-explored major and
The Charters Towers City data (Fig. 21A) show an minor mining camps, where most gold deposits have
approximately linear relationship for box sizes of been detected. By including an undiscovered mining
r ≥ 1.5 km. The fractal dimension can be obtained camp to this model, Blenkinsop and Sanderson (1999)
from the slope of the straight-line part of the data, giving found that the fractal dimension of deposit distribution
a D value of 1.28 (Table 3). For box sizes b 1.5 km, the became smaller and the straight-line part of the data
data depart from this linear relationship. The Charters more gentle than those of the original dataset. Hence, the
Towers South data illustrate an approximately linear lower D values of the Charters Towers South, Hadleigh
relationship for box sizes ranging from 1.5 to 6.0 km, Castle and Charters Towers District subsets, and the
yielding a fractal dimension of 1.10 (Fig. 21B). The more gentle slopes of the straight-line parts of these data
Hadleigh Castle data yield a D value of 1.02 and display with respect to the Charters Towers City subset
a linear relationship for box sizes ranging from 1.0 to (Fig. 21E), may imply that new deposits are more likely
10.0 km (Fig. 21C). The Charters Towers District data to be discovered outside the boundaries of the Charters
(Fig. 21D) give a D value of 1.09, which is based on the Towers City camp.
approximately linear relationship for box sizes ranging Proposition (2) is consistent with the records of
from 1.5 to 10.0 km. By excluding the Charters Towers Reid (1917) and Levingston (1972), describing the
City records from the Charters Towers District data, a most productive veins (Day Dawn-Mexican, Brilliant
slightly lower fractal dimension of 1.02 is obtained for and Queen) as those of greatest structural complexity.
box sizes ranging from 1.5 to 10.0 km (Table 3). A If proposition (2) is valid, a different character of the
normalised diagram for comparison of the Charters fault network (i.e., more bends and splays) could
Towers City, Charters Towers South, Hadleigh Castle explain the difference in the fractal dimension. In this
and Charters Towers District datasets is presented in scenario, work addressing the fractal properties of
Fig. 21E. fault systems in the Lolworth–Ravenswood terrane
would assist with directing exploration into areas of
8.2. Interpretation of results greatest fault complexity.
An interesting aspect of this fractal analysis is that the
As illustrated by the log–log plots of N(r) against r D value of the Charters Towers District dataset is very
(Fig. 21), the data depart from approximately linear similar to that calculated from Archaean lode-gold
68 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Fig. 21. Number of boxes, N(r), containing one or more gold deposits as a function of box size, r. (A) Charters Towers City dataset.
(B) Charters Towers South dataset. (C) Hadleigh Castle dataset. (D) Charters Towers District dataset. (E) Normalised diagram for datasets
(A) to (D).
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 69

Table 3 time of mineralisation and the exact timing, spatial


Fractal properties of gold deposits, Charters Towers goldfield distribution, magnitude and significance of tectonother-
Dataset D n r (km) R S mal events that shaped this terrane. Injection of large
Charters Towers City 1.28 119 1.0–10.0 − 0.9952 0.0289 volumes of melt and associated heat transfer into the
Charters Towers South 1.10 78 1.5–6.0 − 0.9941 0.0492 middle or upper crust (cf. Withnall et al., 1997) of a
Hadleigh Castle 1.02 37 1.0–10.0 − 0.9822 0.0518 potential Cordilleran-type magmatic arc environment
Charters Towers District a 1.09 313 1.5–10.0 − 0.9962 0.0144
(e.g., Kreuzer, 2005) at the same time as deformation
Charters Towers District b 1.02 194 1.5–10.0 − 0.9960 0.0139
(D4 of Kreuzer, 2004), uplift and erosion of the
Abbreviations: D = fractal dimension; n = number of deposits; r = box
Lolworth–Ravenswood terrane (e.g., Kreuzer, 2004;
size; R = correlation coefficient; S = standard error of regression.
a
All Charters Towers District data (n = 313 deposits). Fergusson et al., 2005), is the most likely source of
b
Charters Towers District data without the Charters Towers City energy for mobilisation and transport of ore components
subset (n = 194 deposits). from source to trap regions. The trigger for this
tectonothermal event is unknown. However, Goldfarb
deposits of the Zimbabwe craton (D ≈ 1: Blenkinsop and et al. (2001) and Hagemann and Cassidy (2000)
Sanderson, 1999; Blenkinsop, 2004), raising the suggested that plate interaction, plume tectonics or
possibility of similarities in the underlying controls on combinations of both are critical far-field processes in
gold deposit localisation in both districts. the formation of many lode-gold systems.
If the heat transfer concept of Withnall et al. (1997) is
9. The Charters Towers lode-gold system appropriate, the mappable ingredients and guides for
area selections at the regional-scale would be accumula-
The mineral systems approach (Wyborn et al., 1994) tions of Middle Silurian to Early Devonian intrusions
promotes construction of geological process models in with compositions and inferred volumes similar to those
which ore deposits are considered as focal points of present in the Charters Towers and Etheridge goldfields.
much larger systems of energy and mass flux, having
operated at scales from regional to local (e.g., Lord 9.2. Source of ore components
et al., 2001; Hronsky, 2004). A key step in this mo-
delling approach is the identification of the critical Little is known about the sources of metals, ligands
success factors (i.e., the critical geological elements and fluids, their levels of concentration and their role in
and processes) without which the formation of a par- the spatial distribution of the auriferous veins despite
ticular ore deposit would have been precluded. previous and recent advances in the understanding of
According to Wyborn et al. (1994), Hagemann and their genesis (e.g., Peters, 1987a,b, 1993b; Peters and
Cassidy (2000), Lord et al. (2001) and Huston et al. Golding, 1989; Dowling and Morrison, 1989; Black
(2004), the critical geological process factors in the et al., 1997; Kreuzer, 2003, 2004, 2005, 2006) and the
genesis of lode-gold systems can be summarized as nature of the basement to the Ravenswood batholith
follows: (1) sources of energy that drive a particular (Hutton, 2004). However, agreement exists among these
system, (2) sources from which ore components (fluids, workers that the geological and geochemical informa-
metals and ligands) can be mobilised by melts or fluids, tion may be taken to imply that the local intrusions
(3) active transport pathways along which the mobilised mainly were passive hosts and did not contribute
ore components can migrate, (4) effective, narrow fluid significantly to the metal budget of the auriferous veins.
or melt channels (i.e., trap zones) along which fluid flow A model that accounts for the published geological
become focused, and (5) physico-chemical processes and fluid characteristics of the auriferous veins was
that promote ore deposition within these trap zones. proposed by Kreuzer (2005). In this model, and based
Applied to the Charters Towers goldfield, the mineral on the concept of Stüwe (1998), fluids of metamorphic
systems approach serves as a framework that can help origin were released through deformation and devola-
explorers to better understand why auriferous veins are tilisation of the lower crust of the Lolworth–Ravens-
located where they are and what processes were essential wood terrane approximately 20 my after peak
to forming these deposits (cf. Wyborn et al., 1994). deformation (D1 and D2 of Kreuzer, 2004) of the
upper crust. Immediate transport of these metamorphic
9.1. Source of energy fluids into the upper crust was prohibited by the low
permeability and porosity of the thick mass of freshly
Significant uncertainty exists about the tectonic accumulated intrusions of the Ravenswood batholith
setting of the Lolworth–Ravenswood terrane at the until a new episode of fracturing linked to regional
70 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

shortening (D4 of Kreuzer, 2004) promoted metamor- An understanding of the strain field present at the time
phic fluid flow into dilatant structures. Interaction of the of mineralisation is necessary to identify zones of
metamorphic fluids with the intrusions triggered wall- dilational and contractional deformation that were critical
rock alteration and quartz and sulfide deposition of in focusing fluid flow and localising ore deposition within
stages I and II (Fig. 8). This first episode of the structural array (e.g., Sibson, 1996; Allibone et al.,
hydrothermal fluid flow was followed by the release 2002) of the Lolworth–Ravenswood terrane. Only those
of saline fluids from deep-seated zones of crustal structures that were active during the time of gold
melting, possibly linked to underplating of the crust deposition are likely to host payable auriferous veins,
by mafic magmas (cf. Hutton and Rienks, 1997; Hutton given that permeability is only enhanced and pressure
et al., 1997) or asthenospheric mantle upwelling as gradients that drive fluid flow can only be maintained
proposed for granitoid-hosted lode-gold deposits of the while fault and shear zones are being actively deformed
Jiadong province, Eastern China (Qiu et al., 2002; Chen (e.g., Sibson, 1990; Cox, 1999). In the Charters Towers
et al., 2005). Gold and sulfide deposition of stage III was goldfield, the identification of structures that were active
caused by channelling of fluids that originated from at the time of mineralisation is straightforward only where
deep-seated zones of crustal melting into active exposed fractures are filled with auriferous quartz veins or
structures previously affected by stages I and II and relative sense and timing of displacement along concealed
mixing them with cooler, more dilute fluids that structures can be inferred from geological or geophysical
circulated in the upper crust (cf. Schreiber et al., data with a high degree of confidence. In all other
1990a,b; Mishra and Panigrahi, 1999; Haeberlin, 2002). situations, judgments about the age and nature of a
In this model, the sources of ore components and particular structure are made under conditions of
processes of mobilisation of ore components are uncertainty. However, NNW–SSE- (±15°) and ENE–
threefold: deformation of the lower crust and associated WSW- (±15°) striking structures are more likely to have
metamorphic devolatilisation, deep-seated crustal melt- active at the time of mineralization than structures of
ing and associated magmatic devolatilisation, and different orientation, given that: (1) virtually all auriferous
downward flow of meteoric fluids into faults that were veins are hosted by NNW–SSE- (±15°) and ENE–WSW-
actively being deformed. According to this model, (±15°) striking fractures, suggesting that these orienta-
prospective areas should be those where Middle Silurian tions were most favourable for the generation of new, and
to Early Devonian granitoid emplacement was contem- reactivation of pre-existing, structures (Fig. 22A), and (2)
poraneous with uplift and brittle deformation, and at the district scale, auriferous veins have abundance and
magmatism can be demonstrated to have postdated proximity relationships with NW–SE, NNW–SSE, NE–
peak deformation by tens of millions of years. SW and ENE–WSW-striking lineaments.
Given the same fracture density, groups of short
9.3. Pathways fractures are less well connected than those of long
fractures (Odling, 1997; Hodkiewicz et al., 2005). In
Hutton and Rienks (1997) suggested that, at the addition, longer and more continuous structures gener-
district-scale, transport of ore components from source ally have the potential to tap fluids from larger volumes
regions to trap zones could have been controlled by the of rock (e.g., Cox et al., 2001). These links are manifest
Charters Towers–Ravenswood lineament, an ESE– at Charters Towers (Fig. 22B; Table 1), where most gold
WNW- to E–W-striking suture zone in the basement of metal was extracted from the vein systems that are
the Lolworth–Ravenswood terrane. Such a concept is longest and most continuous (i.e., Day Dawn-Mexican,
consistent with the spatial autocorrelation results, imply- Brilliant and Queen reefs).
ing an ESE–WNW to E–W alignment of the auriferous
veins at the scale of the district. In the context of 9.4. Trap zones
percolation theory (cf. Cox, 1999, 2005; Cox et al., 2001),
the Charters Towers–Ravenswood lineament may be Fluid flow within shear or fault zones is essentially
taken as the backbone element, having provided a direct controlled by gradients in permeability and hydraulic
link between source regions and trap zones. Reactivation head. These gradients are highest at splay faults, fault
of the Charters Towers–Ravenswood lineament during bends, jogs and intersections (e.g., Cox, 1999; Cox et al.,
D4 may have locally enhanced its permeability, thereby 2001), localities that are commonly mineralised within
facilitating fluid flow up and along the backbone and into structures that host auriferous veins of the Charters
reactivated and newly created fractures (dangling ele- Towers goldfield (e.g., Reid, 1917; Peters, 1987a,b;
ments) connected to the backbone. Kreuzer, 2004). According to Allibone et al. (2002), a
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 71

Fig. 22. (A) Contained-ounces-of-gold diagram for the same dataset as in Fig. 3B but grouped according to the strike orientation of the host structures:
large deposits are hosted by E–W-striking structures only; NW–SE-striking structures host mainly small deposits. Data source: Table 1. (B) Crude
relationship between contained ounces of gold and strike length of structures hosting the thirty-five well-documented gold deposits (c = correlation
coefficient).

clear understanding of the deformation history, sense of significant dilation (and gold deposition) to occur at right-
shear during mineralisation, and timing of ore genesis is hand bends of NNW–SSE- (±15°) and left-hand bends of
the key to defining the structural controls on mineralisa- ENE–WSW- (±15°) striking structures (Kreuzer, 2004).
tion and developing a predictive structural targeting No payable gold accumulations have been reported from
strategy. Structural data from auriferous vein systems zones of contractional deformation (e.g., Reid, 1917;
analysed by Kreuzer (2004) are consistent with NE–SW Peters, 1987a,b; Kreuzer, 2004).
(±20°) orientation of the D4 shortening axis during the A proximity relationship is evident between the
time of gold mineralisation, resolving into reverse-dextral spatial distribution of the auriferous veins and domains
shear on ∼N–S-striking and reverse-sinistral slip on ∼E– of low gravity values in the Lolworth and Ravenswood
W-striking host structures. Although problems are raised batholiths, coinciding with distinct intrusions or com-
in defining a bend (e.g., scale and position of observation), plex intrusive bodies (Figs. 2 and 19). The low gravity
historic mine plans serve to illustrate that a 5° to 10° values may indicate that these bodies have high silica
change of strike orientation may have been adequate for contents, and thus are highly brittle. More information
72 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

(e.g., isotopic age data of veins and intrusions) would be Seward, 2002) may have been important depositional
required to constrain this relationship for the Lolworth processes and controls on gold enrichment at the mm- to
batholith. The proximity relationship between aurifer- cm-scale, given that sphalerite, galena, chalcopyrite and
ous veins and domains of low gravity values in the gold minerals of stage III are virtually everywhere
Ravenswood batholith could be explained with pre- spatially associated with aggregates of fractured pyrite
ferred localisation of auriferous veins (particularly those of stage II (Kreuzer, 2006).
N0.1 Moz) at or near geological boundaries where
focusing of fluid flow and physicochemical reactions 10. Implications for prospectivity and targeting
were most likely enhanced as a result of strain
heterogeneities from different mechanical properties of The prospectivity of the goldfield may be considered
adjacent rock units (cf. Ridley, 1993) and chemical and in three categories (Towsey et al., 2002): (1) known
rheological gradients. auriferous vein systems, (2) undiscovered (new) aurifer-
A critical trap factor may be the apparent association ous vein systems, and (3) other styles of mineralisation.
relationship between the spatial distribution of the
auriferous veins and felsic to intermediate, oxidised I- 10.1. Known auriferous vein systems
type intrusions of the medium- to high-K calc-alkaline
Hogsflesh and Millchester supersuites and the high-K Virtually all payable auriferous veins of the Charters
calc-alkaline Broughton River suite (Kreuzer, 2005). It Towers goldfield that were exposed at surface were
could be argued that more auriferous veins are hosted by located within 2 weeks after the initial discovery of gold
intrusions that belong to these particular suites since at the base of Towers Hill in 1871 (Kay, 1993). It was
they make up a greater proportion of the outcrop area of the efficiency of the historic prospectors and thorough-
the Ravenswood batholith than other rock types. ness of subsequent prospecting and exploration activi-
However, cumulative production records suggest that ties in areas of outcrop that contributed to the reputation
the affinity of auriferous veins for felsic to intermediate of Charters Towers as mature goldfield of limited
intrusions of the Hogsflesh and Millchester supersuites prospectivity. The efficiency of previous exploration
and the Broughton River suite is genuine: (1) most of the and high D value of the Charters Towers City data
N6.6 Moz Au produced in Charters Towers were suggest that the probability of finding additional,
extracted from veins hosted by Ordovician and Middle outcropping or near-surface deposits within the Charters
Silurian to Early Devonian intrusions of the Hogsflesh Towers City camp is low. However, potential may exist
and Millchester supersuites and the Broughton River at depth for discovery of extensions to, or new structures
suite, and (2) all but two (Sunburst-Golden Gate and parallel to, known auriferous veins given the limitation
Imperial reefs) out of 23 auriferous veins with of the fractal analysis to two dimensions, scarcity of
documented production greater 15,000 oz (Table 3) drillholes with greater than 100 m downhole depth, great
are hosted by Ordovician and Middle Silurian to Early vertical extent of similar grantoid-hosted lode-gold
Devonian granitoids of these suites. systems elsewhere (e.g., 1.7 km at Parcoy-Pataz, Peru:
Schreiber et al., 1990a,b; Haeberlin, 2002; Haeberlin
9.5. Physicochemical processes et al., 2004), and evidence from historical reports for
inefficient and unsystematic exploration practices at
The range of temperatures and salinities recorded by depth (Towsey et al., 2002).
Kreuzer (2005) from quartz- and sphalerite-hosted fluid
inclusions may be taken to indicate mixing between 10.2. Undiscovered (new) auriferous vein systems
upward-flowing, deep-sourced, hotter, more saline
fluids and downward-flowing, shallow-sourced, cooler, The best chance of detecting new, sizeable lode-gold
more dilute waters as the cause of gold deposition. systems within the Charters Towers district is in areas
Hence, a key factor in quantifying the prospectivity of where prospective rock types of, and structures cutting,
the Charters Towers goldfield in the third dimension is the Ravenswood batholith are poorly exposed or hidden
to establish the vertical range over which the deep under the Tertiary to Quaternary cover (commonly 2 to
sourced fluids mixed with the cooler more dilute fluids 40 m thick: Aspandiar et al., 2003). Most of these areas
occurred, and the mechanisms by which the fluids were not only inaccessible to the historic prospectors but
mixed with such uniformity at the district-scale. also have received relatively little attention from recent
Physi- and chemisorption of gold complexes on to explorers. Given that structure was the first order control
sulfide surfaces (cf. Knipe et al., 1992; Widler and on the spatial distribution of auriferous veins in the
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 73

Charters Towers goldfield, the following steps in the mineralisation, the controlling factors can be related to a
development of a targeting strategy for areas under single genetic model for orogenic, granitoid-hosted
cover are suggested: lode-gold mineralisation in a D4 brittle deformation
regime of NE–SW to NNE–SSW shortening, under
(1) To identify from geological and geophysical data conditions of supralithostatic fluid pressure and low
the ENE–WSW (± 15°) and NNW–SSE (± 15°) stress difference. It is salutary for exploration that many
striking structures and geological boundaries possible factors, on different scales, need to be
within a 20-km-wide corridor parallel to and considered to predict targets, even though mineralisation
centred upon the Charters Towers–Ravenswood was likely to have occurred in a single phase in the Late
lineament that appears to have controlled the Silurian to Early Devonian, and there is no evidence for
spatial distribution of gold deposition at the remobilisation.
regional- to district-scale. In this example, which is considerably simpler than
(2) To interpret from geological and geophysical data many examples of older orogenic lode-gold deposits,
the distribution of felsic to intermediate, pre- some of the important controls on mineralisation and
Middle Devonian intrusions within these areas. implications for exploration can be detailed as follows:
(3) To deduce from geophysical data the ENE–WSW
(± 15°) and NNW–SSE (± 15°) striking struc- (1) Ore zones within the auriferous veins are
tures that cut or bound the intrusions identified in commonly located at or within fault bends, jogs,
step 2. intersections, relays and termination splays along
(4) To locate segments along the longest of the ENE–WSW- (± 15°) and NNW–SSE- (± 15°)
structures identified in step 3 that deviate most striking structures.
from the geometry of a straight line (e.g., potential (2) Geophysical data illustrate three major groups of
bends or splays) and/or intersect other structures lineaments in the Charters Towers (group I: NW–
or geological contacts, or both. SE to NNW–SSE; group II: NE–SW to ENE–
(5) To define and rank potential targets within the WSW; group III: E–W to WNW–ESE) and two in
prospective areas identified in step 4 and to the Hadleigh Castle (group I: NE–SW to E–W;
systematically test the best ones. group II: NNW–SSE to NW–SE) areas, whereas
the Lolworth–Ravenswood terrane is characterised
10.3. Other styles of mineralisation by five major groups of lineaments (group I: NW–
SE; group II: ENE–WSW; group III: E–W to
The Charters Towers goldfield is known for Late WNW–ESE; group IV: NE–SW; group V: N–S).
Silurian to Early Devonian lode-gold and Late Carbon- Strike orientations of the auriferous veins are
iferous to Early Permian porphyry-related, breccia-hosted parallel to the strike orientations of all lineament
gold mineralisation (i.e., Mt Leyshon: Figs. 2 and 4). groups listed above.
However, rocks and sediments within the Charters Towers (3) A proximity relationship is evident between the
goldfield host additional gold occurrences, such as spatial distribution of auriferous veins and
sheeted auriferous veins and vein stockworks assigned domains of low gravity values in the Ravenswood
as Permian or Carboniferous, and alluvial placers in batholith, coinciding with distinct intrusions. The
present-day drainage systems and lithified placers in relationship is interpreted to reflect preferred
older, Tertiary to Quaternary leads and residuals (e.g., localisation of auriferous veins at or near
Hartley and Dash, 1993; Hutton and Rienks, 1997; Ewers, geological boundaries where focusing of fluid
1997; Towsey et al., 2002). While none of these styles flow and physicochemical reactions could have
have proven payable, future exploration in the Charters been enhanced as a result of strain heterogeneities
Towers goldfield should remain vigilant to encompass from different mechanical properties of adjacent
and test for deposit styles other than the auriferous veins rock units and chemical and rheological gradients
described in this paper. (cf. Hodkiewicz et al., 2005).
(4) Spatial autocorrelation suggests an ESE–WNW
11. Conclusions to E–W alignment of the auriferous veins at the
district-scale, supporting the concept of Hutton and
A variety of sites of gold mineralisation can be Rienks (1997) of gold deposit localisation along
demonstrated in the Charters Towers goldfield, NE the Charters Towers–Ravenswood lineament (new
Australia. Despite a number of different controls on name), an ESE–WNW- to E–W-striking suture
74 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

zone in the basement of the Lolworth–Ravenswood support through School of Earth Sciences and Interna-
terrane previously inferred from geophysical and tional Postgraduate scholarships. Project setup and
geochemical data. At the camp-scale, auriferous supervision by, and discussions with Andrew Allibone,
veins have abundance and proximity relationships Nick Oliver and Simon Dominy are greatly appreciated.
with NW–SE-, NNW–SSE-, NE–SW- and ENE– Gregg Morrison (Klondike Exploration Services), Steve
WSW-oriented lineaments. The variability in strike King (Solid Geology), Malcolm Stallman (Normandy
orientation between different camps implies that Exploration) and Mike Sexton (Newmont Australia) are
structures that controlled gold deposition in one thanked for having made available important geological
camp did not necessarily control mineralisation in and geophysical datasets. Support of this project by and
other camps. discussions with previous and current staff members of
(5) Auriferous veins in the Charters Towers goldfield Citigold Corporation Ltd (Nigel Storey), Glengarry
are generally highly clustered. Fractal dimensions Resources Ltd (Tony Alston, David Richards), SMC
range from 1.02 to 1.10, although auriferous veins Gold Ltd (Brett Duck, Ian Hodkinson) also are greatly
in the main mining area are characterised by a appreciated. RJM gratefully acknowledges the support
significantly higher fractal dimension of 1.28. and permission of Citigold Corporation Ltd. Reviews by
This discrepancy may be taken to imply that most David Groves (University of Western Australia), Art
or all outcropping and near-surface deposits of the Schultz and Klaus Schultz (US Geological Survey), and
Charters Towers City camp have been discovered Ore Geology reviewers Alexander Yakubchuk (Gold
(i.e., the fractal has been well sampled) and that Fields International Services) and Craig Hart (Univer-
new discoveries are more likely to occur outside sity of Western Australia), and Nigel Cook (Editor-in-
the boundaries of this camp. However, this does Chief, Ore Geology Reviews) are greatly appreciated
not downgrade the potential within the Charters and helped to improve the manuscript.
Towers City camp for discovery at depth of
extensions to, and new structures parallel to,
known auriferous veins. Appendix A
(6) The best chance of detecting new, sizeable near-
surface lode-gold systems in the Charters Towers Australian map grid (AMG) coordinates of gold
district is in areas where prospective rock types deposits and occurrences (sources: Hartley and Dash,
and structures of the Ravenswood batholith are 1993; Hartley, 1996)
hidden under the Tertiary to Quaternary cover.
Exploration targeting in these areas will have to Name/reference # Grid reference
rely on geophysical data to identify prospective mE mN
areas along the Charters Towers–Ravenswood CT City
lineament. Aberdeen 423700 7777000
(7) Large portions of the Ravenswood batholith are Alabama 421480 7778720
poorly exposed or hidden under the cover. These Band of Hope 419530 7781260
Bonnie Dundee 424100 7779800
areas were inaccessible to the historic prospectors
Bramble Lode 423950 7779120
and have received relatively little attention from Brilliant 423280 7779030
recent explorers. Hence, we strongly reject the Captain 426100 7777420
notion of Charters Towers being a mature Chance 424980 7775900
goldfield of poor prospectivity. Clark's Gold Mine 421200 7778270
Columbia 423480 7781300
Commodore 420850 7778170
The approach taken in this paper, namely synthesiz- Contest 421620 7778470
ing a wealth of previous information with new data and Coronation 423300 7779100
a genetic model, in combination with spatial autocor- Craven's Caledonia 424180 7779090
relation and fractal analysis, could have implications for CT Consolidated 422600 7778300
Dan O'Connell 423500 7780050
the assessment of other “mature” goldfields worldwide.
Day Dawn 421200 7779500
Daybreak 423700 7775600
Acknowledgements Duke of Cornwall 426700 7777700
East Mexican 422800 7779200
This paper forms part of a doctoral study at James East Sunburst 425600 7780150
Eclipse 423600 7777960
Cook University by OPK, who received financial
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 75

Appendix A (continued ) Appendix A (continued )


Name/reference # Grid reference Name/reference # Grid reference
mE mN mE mN
Gladstone 424870 7777805 Perfect Cure 426900 7776750
Golden Alexandra 425000 7775700 Perseverance 423200 7779200
Golden Gate 424880 7780020 Phoebe 423500 7779460
Golden Sunset 420850 7780350 Picanniny 424500 7775280
Golden Surprise 424600 7778200 Pilot 426000 7774900
Grand Secret 421600 7779500 Poverty 422500 7775940
Guiding Star 420600 7777550 Queen 424000 7779700
Havelock 421560 7777810 Queenslander 426740 7777000
Hidden Treasure 421050 7778640 Rainbow 422250 7778380
Homerule 423370 7779660 Rainbow-Wyndham 422150 7778680
Identity 425740 7778060 Recompense 420500 7777530
Independence 427100 7777150 Resolution 423320 7777520
Irishman 425200 7777900 Rose of Denmark 426900 7777700
John Bright 427100 7777460 Rose of England 424440 7777540
Just-in-Time Lode 424050 7779300 Royal George 423330 7779750
Kelly's Queen 423550 7779550 Ruby 423950 7777460
Kelly's Welcome 420320 7778200 RubyUnited 424100 7777600
Kroko 424100 7779000 Sir William Wallace 421200 7777650
Lady Antrim 426300 7778320 Sisters 420780 7779000
Lady Carrington 424050 7778900 Sons of Freedom 425600 7775000
LadyDon 427450 7777300 St Patrick 423430 7780120
Lady Florence 426300 7778360 Standard 421570 7778450
LadyMaria 421150 7778500 Sunburst 424920 7779990
Liverpool 424750 7780340 Bucklands Hill 420900 7782200
Lord Nelson 421130 7777500 Great Britain 419380 7781230
Madelaine 421300 7777570 Iceland 419850 7781300
Mafekin 427600 7778250 Shamrock 419330 7781320
Marquis of Lome 422200 7775740 Sunlight 420940 7778150
Marshall's Queen 423880 7779750 Talisman 422300 7779300
Martin Lyons 421340 7777690 Telegraph 429740 7778000
Mary 422450 7778180 Third Chance 423930 7779300
MaryLou 420850 7775900 Tuckett's Lode 423540 7779140
Maude St Ledger 423730 7779280 Union Lode 420390 7778180
Mayday 424250 7779100 Venus 426300 7778600
McDonald 424700 7777580 Vesuvius 424480 7779630
Millchester 427400 7777550 Victoria 424070 7779170
Millicans Caledonia 424370 7778980 Victory 423560 7779400
Mississippi 421750 7777920 Vixen 423100 7776380
Monte Christo 420620 7778300 Waniola 422740 77706U
Moonstone 423590 7779230 Warrior 424860 7774780
Morris' Lode 423530 7779164 Washington 425400 7775300
Mountain Maid 421420 7778940 Welcome 425700 7775190
Mystery Deeps 423500 7780850 Wellington 421100 7777460
Nellie Grant 426700 7777540 Worcester 423500 7779300
Nell's Extended 424800 7777230 Zerviles 427600 7776700
New Queen Cross 423900 7779150
New St Andrews 423600 7780000 CT South
Normanby 423000 7778500 #175670 417510 7766960
North Australian 421680 7778800 #176666 417600 7766600
North German 424540 7776000 #190709 419050 7770950
North Queen 424300 7779970 #191660 419150 7766000
North Star 423340 7780500 #192712 419200 7771200
Old Queen Cross 423800 7779400 #194670 419450 7767000
Pacific 425250 7775350 #197654 419660 7765360
Papuan 423830 7779270 #197658 419750 7765850
Pauline 427400 7777300 #203707 420260 7770700
Peabody 421500 7778560
(continued on next page)
76 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Appendix A (continued ) Appendix A (continued )


Name/reference # Grid reference Name/reference # Grid reference
mE mN mE mN
#210744 421000 7774400 Scandinavian 418600 7769900
#215739 421500 7773900 Silent Friend 424270 7773060
#217725 421700 7772530 Six Mile 415620 7773450
#221731 422070 7773090 Southern Cross 410800 7771500
#253730 425300 7773000 Stock Squad West 412930 7769750
#253734 425300 7773400 Stockholm Cross 418360 7774340
Anglo France 418120 7761580 Swedenborg 418800 7770100
Beary Creek 412800 7768900 Try Again 421260 7765500
Beary Creek West 412530 7768750 Try Again Group 412300 7771650
Black Jack Lode 418060 7772100 Union Jack 420100 7769920
Black Jack PC 418000 7771870 West Imperial 425260 7773940
Black Jack South 418820 7771870 Wheel of Fortune 424600 7773600
Black Knob 414300 7757240
Bonnie Doon 425550 7773350 HC
Butler Block 418300 7772280 #482726 448223 7772581
Caroline 422220 7773260 #486730 448554 7772981
Christmas Box 420800 7773050 #489733 448946 7773347
Clara 420900 7773250 #495729 449507 7772921
Clarke River Veins 414250 7757640 #495731 449516 7773051
Cumberland 420960 7771300 #497730 449671 7773043
Cumberland West 420900 7771300 #500724 449950 7772430
Curlew 409900 7769900 Alabama 455250 7773630
Curlew South 409500 7769000 Albion 458900 7776500
Debbie's Find 411750 7758100 Alma 453300 7773800
Democrat 414500 7756800 Arena 448821 7772766
Democrat East 414770 7757140 Arena West 448696 7772861
Democrat West 414000 7756800 Aurelia 448900 7774500
Golden Ant 420850 7771200 Battery 452100 7773750
Golden Bar 421040 7771040 Blue Duck 462600 7777770
Golden Gate South 411280 7770600 Bosun 447250 7773700
GoldenGateWest 411130 7771060 Captain 458720 7776500
Golden Spider 420980 7771060 Churchill 452600 7773500
Golden Spur 419980 7768940 Commodore 459382 7776903
Goldfinch 419500 7770080 Crossroads 455400 7775400
Gregory's Try Again 421000 7765900 Disraeli 451800 7774000
Hidden Secret 423300 7772670 Eden 452640 7773880
Hidden Treasure 413700 7756960 Elsie Koenig 451450 7774000
Imperial 425500 7774050 Englishman 447900 7770600
Ivory Elephant 420700 7766200 Forget-Me-Not 451000 7773500
John Bull 418200 7772000 Greek Extended 449319 7775261
Lady Musgrave 418980 7767500 Hadleigh Castle 459442 7777018
Ladybird 424700 7774050 Hadleigh Castle East 459935 7777121
Lubra 418250 7771750 Lightening 452300 7774300
LuckyProp 413420 7772660 Mi Mi 449669 7773160
Mabel Jane 423000 7767900 Mi Mi East 450124 7772857
Merrie Monarch 423040 7772080 Mickeries 460540 7779200
Monarch View 423820 7771940 Rishton 451050 7774100
Mount Censis 422200 7772960 Rishton Scrub 451000 7771000
Mount Charles 413500 7768830 Robinson Crusoe 457068 7776144
MtCharlesEast 413650 7768730 Rochford Syndicate 459000 7776000
Mt Pleasant North 413700 7762800 Try No More 451000 7773000
Mt Pleasant South 412980 7760050
Newtown Butler 418480 7772450 AMG coordinates were used as input data for spatial
Powerline 411930 7768230
autocorrelation in FryPlo and fractal analysis in Boxcount.
Republic 418500 7768250
Rollston 424600 7773060 Abbreviations: CT City = Charters Towers City camp;
Same as Usual 422700 7772000 CT South = Charters Towers South camp; HC = Hadleigh
Castle Camp.
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 77

References ern Argentina: geological, geophysical, and remote sensing


evidence. Ore Geology Reviews 21, 127–155.
Allibone, A., Teasdale, J., Cameron, G., Etheridge, M., Uttley, P., Coney, P.J., Edwards, A., Hine, R., Morrison, F., Windrim, D., 1990.
Soboh, A., Appiah-Kubi, J., Adanu, A., Arthur, R., Mamphey, J., The regional tectonics of the Tasman orogenic system, eastern
Odoom, B., Zuta, J., Tsikata, A., Pataye, F., Famiyeh, S., 2002. Australia. Journal of Structural Geology 12, 519–543.
Timing and structural controls on gold mineralization at the Bogoso Cooke, D.R., McPhail, D.C., 2001. Epithermal Au–Ag–Te mineral-
Mine, Ghana, West Africa. Economic Geology 97, 949–969. ization, Acupan, Baguio district, Philippines: numerical simula-
Aspandiar, M.F., Taylor, G., Eggleton, R.A., 2003. Charters Towers tions of mineral deposition. Economic Geology 96, 109–131.
region, Queensland. In: Anand, R.R., de Broekert, P. (Eds.), Cox, S.F., 1999. Deformational controls on the dynamics of fluid flow
Regolith Landscape Evolution. Cooperative Research Centre for in mesothermal gold systems. In: McCaffrey, J.W., Lonergan, L.,
Landscape Environments and Mineral Exploration. bhttp://leme. Wilkinson, J.J. (Eds.), Fractures, Fluid Flow and Mineralization.
anu.edu.au/RegLandEvol/ChartersTowers.pdfN, 4 pp. Geological Society of London, Special Publication, vol. 155,
Bain, J.H.C., Draper, J.J., 1997. North Queensland Geology. Australian pp. 123–140.
Geological Survey Organization Bulletin 240, and Queensland Cox, S.F., 2005. Coupling between deformation, fluid pressures, and
Department of Mines and Energy. Queensland Geology 9, 600 pp. fluid flow in ore-producing hydrothermal systems at depth in the
Bain, J.H.C., Withnall, I.W., Black, L.P., Etminan, H., Golding, S.D., Sun, crust. In: Hedenquist, J.W., Thompson, F.H., Goldfarb, R.J.,
S.-S., 1998. Towards an understanding of the age and origin of gold Richards, J.P. (Eds.), Economic Geology 100th Anniversary
mineralization in the Etheridge Goldfield, Georgetown region, north Volume, pp. 39–75.
Queensland. Australian Journal of Earth Sciences 45, 247–263. Cox, S.F., Wall, V.J., Etheridge, M.A., Potter, T.F., 1991. Deforma-
Berry, R.F., Huston, D.L., Stolz, A.J., Hill, A.P., Beams, S.D., Kuronen, tional and metamorphic processes in the formation of mesothermal
U., Taube, A., 1992. Stratigraphy, structure, and volcanic-hosted vein-hosted gold deposits — examples from the Lachlan fold belt
mineralization of the Mount Windsor subprovince, north Queens- in central Victoria, Australia. Ore Geology Reviews 6, 391–423.
land, Australia. Economic Geology 87, 739–763. Cox, S.F., Knackstedt, M.A., Braun, J., 2001. Principles of structural
Bierlein, F.P., Crowe, D.E., 2000. Phanerozoic orogenic lode gold control on permeability and fluid flow in hydrothermal systems. In:
deposits. In: Hagemann, S.G., Brown, P.E. (Eds.), Gold in 2000. Richards, J.P., Tosdal, R.M. (Eds.), Structural Controls on Ore
Reviews in Economic Geology, vol. 13, pp. 103–139. Genesis. Society of Economic Geologists, Littleton, Reviews in
Black, L.P., Carr, G.R., Sun, S-S., 1997. Applied isotope geochronol- Economic Geology, vol. 14, pp. 1–24.
ogy and geochemistry. In: Bain, J.H.C., Draper, J.J. (Eds.), North Denmead, A.K., Levingston, K.R., 1950. Report on the Black Jack gold
Queensland Geology. Australian Geological Survey Bulletin 240, mine, Charters Towers. Queensland Government Mining Journal 52,
and Queensland Department of Mines and Energy Queensland 128–133.
Geology, vol. 9, pp. 429–447. Dominy, S.C., Hodkinson, I.P., Kidd, R.G., 1999. Meeting the challenges of
Blatchford, A., 1953. Charters Towers goldfield. Proceedings of the narrow-vein gold mining; Hadleigh Castle mine, Charters Towers,
Empire Mining and Metallurgical Congress 1, 796–806. north Queensland, Australia. Transactions, Institution of Mining and
Blenkinsop, T.G., 1994. The fractal distribution of gold mineralisation: Metallurgy Section A: Mining Industry 108, 192–205.
two examples from the Zimbabwe Archaean craton. In: Kruhl, J.H. Dowling, K., Morrison, G.W., 1989. Application of quartz textures to
(Ed.), Fractals and Dynamic Systems in Geoscience. Springer, the classification of gold deposits using North Queensland
Berlin–Heidelberg, pp. 247–258. examples. In: Keays, R.R., Ramsay, W.R.H., Groves, D.I. (Eds.),
Blenkinsop, T.G., 1995. Fractal measures for size and spatial distributions The Geology of Gold Deposits: The Perspective in 1988.
of gold mines: economic applications. In: Blenkinsop, T.G., Economic Geology Monograph, vol. 6, pp. 342–355.
Tromp, P.L. (Eds.), Sub-Saharan Economic Geology. Geological Etheridge, M.A., 1983. Differential stress magnitudes during regional
Society of Zimbabwe, Special Publication, vol. 3, pp. 177–186. deformation and metamorphism: upper bound imposed by tensile
Blenkinsop, T.G., 2004. Applications of fractal geometry to mineral fracturing. Geology 11, 231–234.
exploration. In: Muhling, J., Goldfarb, R., Vielreicher, N., Bierlein, Ewers, G.R., 1997. Mineral and energy deposit styles and potential
F., Stumpfl, E., Groves, D.I., Kenworthy, S. (Eds.), SEG 2004 — resources. In: Bain, J.H.C., Draper, J.J. (Eds.), North Queensland
Predictive Mineral Discovery Under Cover. Centre for Global Geology. Australian Geological Survey Organization Bulletin 240,
Metallogeny, Perth, The University of Western Australia Publica- and Queensland Department of Mines and Energy Queensland
tion, vol. 33, pp. 158–161. Geology, vol. 9, pp. 529–546.
Blenkinsop, T.G., Sanderson, D.J., 1999. Are gold deposits in the crust Fergusson, C.L., Henderson, R.A., Lewthwaite, K.J., Phillips, D.,
fractals? A study of gold mines in the Zimbabwean craton. In: Withnall, I.W., 2005. Structure of the Early Palaeozoic Cape River
McCaffrey, K.J.W., Lonergan, L., Wilkinson, J.J. (Eds.), Fractures, Metamorphics, Tasmanides of north Queensland: evaluation of the
Fluid Flow and Mineralization. Geological Society of London roles of convergent and extensional tectonics. Australian Journal of
Special Publication, vol. 155, pp. 141–151. Earth Sciences 52, 261–277.
Blevin, P.L., Chappell, B.W., Allen, C.M., 1996. Intrusive metallogenetic Fry, N., 1979. Random point distribution and strain measurements in
provinces in eastern Australia based on granite source and composition. rocks. Tectonophysics 60, 89–105.
Transactions of the Royal Society of Edinburgh 87, 281–290. Glen, R.A., 2005. The Tasmanides of eastern Australia. Gray, D.R,
Carlson, C.A., 1991. Spatial distribution of ore deposits. Geology 19, Foster, D.A., 2005. Tasman orogenic belt. In: Vaughan, A.P.M.,
111–114. Leat, P.T., Pankhurst, R.J. (Eds.), Terrane Processes at the Margins
Chen, Y.-J., Piranjo, F., Qi, J.-P., 2005. Origin of gold metallogeny and of Gondwana. Geological Society of London, Special Publica-
sources of ore-forming fluids, Jiaodong province, Eastern China. tions, vol. 246, pp. 23–96.
International Geology Review 47, 530–549. Goldfarb, R.J., Phillips, G.N., Nokleberg, W.J., 1998. Tectonic setting
Chernicoff, C.J., Richards, J.P., Zappettini, E.O., 2002. Crustal of synorogenic gold deposits of the Pacific Rim. Ore Geology
lineament control on magmatism and mineralization in northwest- Reviews 13, 185–218.
78 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Goldfarb, R.J., Groves, D.I., Gardoll, S., 2001. Orogenic gold and Groves, D.I., Kenworthy, S. (Eds.), SEG 2004 — Predictive Mineral
geologic time: a global synthesis. Ore Geology Reviews 18, 1–75. Discovery Under Cover. Centre for Global Metallogeny, The
Gray, D.R, Foster, D.A., 2005. Tasman orogenic belt. In: Selley, R.C., University of Western Australia Publication, vol. 33, pp. 129–133.
Cocks, L.R.M., Plimer, I.R. (Eds.), Encyclopedia of Geology — Huston, D.L., Wygralak, A., Mernagh, T., Vandenberg, L., Crispe, A.,
Volume 1. Elsevier Academic Press, Oxford, pp. 237–251. Lambeck, L., Cross, A., Fraser, G., Williams, N., Worden, K.,
Groves, D.I., Goldfarb, R.J., Gebre-Mariam, M., Hagemann, S.G., Meixner, T., 2004. The Tanami region, northern Australia, a
Robert, F., 1998. Orogenic gold deposits: a proposed classification summary of its geology and mineralization. Australian Institute of
in the context of their crustal distribution and relationship to other Geoscientists News 77, 1–2 and 4–9.
gold deposit types. Ore Geology Reviews 13, 7–27. Hutton, L.J., 2004. Petrogenesis of I- and S-type Granites in the Cape
Groves, D.I., Goldfarb, R.J., Robert, F., Hart, C.J.R., 2003. Gold River — Lolworth area, northeastern Queensland. Unpublished
deposits in metamorphic belts: overview of current understanding, PhD Thesis, University of Queensland, 216 pp.
outstanding problems, future research, and exploration signifi- Hutton, L.J., Crouch, S.B.S., 1993. Geochemistry and petrology of the
cance. Economic Geology 98, 1–29. western Ravenswood batholith. Queensland Department of
Gunn, M.J., Honey, F.R., Lyon, R.J.P., Furnell, R.G., 1989. Geoscan Minerals and Energy, Brisbane, Queensland Geological Record,
AMSS leads to gold mineralization at Crossroads, Queensland. In: vol. 22 (73 pp.).
Broadbent, G., Furnell, R., Morrison, G.W. (Eds.), North Queens- Hutton, L.J., Rienks, I.P., 1997. Geology of the Ravenswood batholith.
land Gold '98 Conference Proceedings. Australasian Institute of Queensland Geology, vol. 8. Queensland Department of Mines and
Mining and Metallurgy, Parkville, p. 139. Energy, Brisbane (60 pp.).
Haeberlin, Y., 2002. Geological and structural setting, age, and Hutton, L.J., Rienks, I.P., Tenison-Woods, K.L., Hartley, J.S., Crouch,
geochemistry of the orogenic gold deposits of the Pataz province, S.B.S., 1994. Geology of the Ravenswood batholith, north
eastern Andean Cordillera, Peru. Unpublished PhD thesis, Queensland. Queensland Geological Record 1994/4 124 pp.
University of Geneva, 182 pp. Hutton, L.J., Draper, J.J., Rienks, I.P., Withnall, I.W., Knutson, J.,
Haeberlin, Y., Moritz, R., Fontbote, L., Cosca, M., 2004. Carbonif- 1997. Charters Towers region. In: Bain, J.H.C., Draper, J.J. (Eds.),
erous orogenic gold deposits at Pataz, eastern Andean Cordillera, North Queensland Geology. Australian Geological Survey
Peru: geological and structural framework, paragenesis, alteration, Organization, Canberra, Bulletin 240, and Queensland Depart-
and 40Ar/39Ar geochronology. Economic Geology 99, 73–112. ment of Mines and Energy, Brisbane, Queensland Geology, vol. 9,
Hagemann, S.G., Cassidy, K.F., 2000. Archaean lode gold deposits. In: pp. 165–224.
Hagemann, S.G., Brown, P.E. (Eds.), Gold in 2000. Reviews in Isles, D., 1994. Regional interpretation of aeromagnetic data from the
Economic Geology, vol. 13, pp. 9–68. Charters Towers and Homestead 1:100,000 sheet areas, “Barring-
Hammond, R.L., 1986. Large scale structural relationships in the ton Project”, Queensland. Unpublished Technical Memorandum to
Palaeozoic of northeastern Queensland: melange and mylonite Poseidon Exploration Ltd.
development, and the regional distribution of strain. Unpublished Johnston, W.D., 1940. The gold quartz veins of the Grass Valley,
PhD thesis. James Cook University, Townsville, Australia, 328 pp. California. United States Geological Survey Professional Paper
Hartley, J.S., 1996. Mineral occurrences — Ravenswood 1:100,000 194, 101 pp.
sheet area. Department of Minerals and Energy Queensland, Karcz, Z., Scholz, C.H., 2003. The fractal geometry of some stylolites
Brisbane, Queensland Geological Record 1996/2, 55 pp. from the Calcare Massiccio Formation, Italy. Journal of Structural
Hartley, J.S., Dash, P.H., 1993. Mineral occurrences — Charters Geology 25, 1301–1316.
Towers 1:100,000 sheet area. Department of Minerals and Energy Kay, J.R., 1992. EPM 8123 Lincoln Hill: exploration report for the
Queensland, Queensland Geological Record 1993/6, 53 pp. twelve months ended 10th July 1992. Unpublished report prepared
Hartley, J.S., Peters, S.G., Beams, S.D., 1989. Current developments on behalf of Mt Leyshon Gold Mines Ltd.
in Charters Towers: geology and gold mineralization. In: Broad- Kay, J.R., 1993. A history of mining in Charters Towers. In:
bent, G., Furnell, R., Morrison, G.W. (Eds.), Proceedings North Henderson, R.A. (Ed.), Guide to the Economic Geology of the
Queensland Gold '98. Australasian Institute of Mining and Charters Towers Region, Northeastern Queensland. Geological
Metallurgy, Parkville, pp. 7–14. Society of Australia Field Excursion Guidebook, James Cook
Henderson, R.A., 1986. Geology of the Mt Windsor Subprovince — a University, Townsville, pp. 25–27.
lower Palaeozoic volcano-sedimentary terrane in the northern Kearey, P., Brooks, M., Hill, I., 2002. Introduction to Geophysical
Tasman orogenic zone. Australian Journal of Earth Sciences 33, Exploration, Third Edition. Blackwell Science, Oxford. 262 pp.
343–364. King, S., 1998. Structural controls on mineralisation in the Charters
Henderson, R.A., Davis, B.K., Fanning, C.M., 1998. Stratigraphy, age Towers region, Queensland. Unpublished report prepared for
relationships and tectonic setting of rift-phase infill in the Normandy Exploration Ltd.
Drummond Basin, central Queensland. Australian Journal of Knipe, S.W., Foster, R.P., Stanley, C.J., 1992. Role of sulfide surfaces
Earth Sciences 45, 579–595. in sorption of precious metals from hydrothermal fluids. Transac-
Hodkiewicz, P.F., Weinberg, R.F., Gardoll, S.J., Groves, D.I., 2005. tions, Institution of Mining and Metallurgy Section B Applied
Complexity gradients in the Yilgarn Craton: fundamental controls on Earth Science 101, 83–88.
crustal-scale fluid flow and the formation of world-class orogenic- Kreuzer, O.P., 2003. Structure, timing and genesis of auriferous quartz
gold deposits. Australian Journal of Earth Sciences 52, 831–841. veins in the Charters Towers goldfield, north Queensland:
Hodkinson, I.P., 1998. Geology of the Hadleigh Castle mine, Charters implications for exploration and prospectivity. Unpublished PhD
Towers. In: Beams, S.D. (Ed.), Economic Geology of Northeast thesis, James Cook University, Townsville, 310 pp.
Queensland — The 1998 Perspective. Geological Society of Kreuzer, O.P., 2004. How to resolve the controls on mesothermal vein
Australia, Melbourne, pp. 230–241. systems in a goldfield characterised by sparse kinematic
Hronsky, J.M.A., 2004. The science of exploration targeting. In: information and fault reactivation — a structural and graphical
Muhling, J., Goldfarb, R., Vielreicher, N., Bierlein, F., Stumpfl, E., approach. Journal of Structural Geology 26, 1043–1065.
O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80 79

Kreuzer, O.P., 2005. Intrusion-hosted mineralization in the Charters Murray, C.G., 1986. Metallogeny and tectonic development of the Tasman
Towers goldfield, north Queensland: new isotopic and fluid Fold Belt System in Queensland. Ore Geology Reviews 1, 315–400.
inclusion constraints on the timing and origin of the auriferous Murray, C.G., Kirkegaard, A.G., 1978. The Thomson Orogen of the
veins. Economic Geology 100, 1583–1603. Tasman Orogenic Zone. Tectonophysics 48, 299–325.
Kreuzer, O.P., 2006. Textures, paragenesis and wall-rock alteration of Odling, N.E., 1997. Fluid flow in fractured rocks at shallow levels in
lode-gold deposits in the Charters Towers district, north Queens- the Earth's crust: an overview. In: Holness, M.B. (Ed.),
land: implications for the conditions of ore formation. Mineralium Deformation-Enhanced Fluid Transport in the Earth's Crust and
Deposita 40, 639–663. Mantle. The Mineralogical Society Series, vol. 8. Chapman and
Kreuzer, O.P., Alston, A.J., 2004. What are the chances of discovery in Hall, pp. 289–320.
the ‘over-explored’ Charters Towers region of north Queensland. Perkins, C., Kennedy, A.K., 1998. Permo-Carboniferous gold epoch of
In: Camuti, K., Young, D. (Eds.), Northern Queensland Explora- northeast Queensland. Australian Journal of Earth Sciences 45,
tion and Mining 2004. Australian Institute of Geoscientists 185–200.
Bulletin, vol. 40, pp. 13–17. Peters, S.G., 1987a. Geology, fluid characteristics, lode controls and
Kreuzer, O.P., Dominy, S.D., Platten, I.M., Raine, M.D., 2002. Ore ore-shoot growth in mesothermal gold-quartz veins, north-eastern
controls and grade distribution in the Charters Towers goldfield. Queensland. Unpublished PhD thesis, James Cook University,
In: Vearncombe, S. (Ed.), Applied Structural Geology for Mineral Townsville, 277 pp.
Exploration and Mining. Australian Institute of Geoscientists Peters, S.G., 1987b. Geology and lode controls of the Charters Towers
Bulletin, vol. 36, pp. 96–99. Goldfield, north-eastern Queensland. Contributions of the Eco-
Lagarde, J.L., Omar, S.A., Roddaz, B., 1990. Structural characteristics nomic Geology Research Unit, vol. 19. James Cook University,
of granitic pluton emplacement during weak regional deformation: Townsville. 117 pp.
examples from Late Cretaceous plutons, Morocco. Journal of Peters, S., 1990. Lode controls of the Charters Towers goldfield,
Structural Geology 12, 805–821. northeastern Queensland. Proceedings of the Australasian Institute
Ledesert, B., Dubois, J., Velde, B., Meunier, A., Genter, A., Badri, A., of Mining and Metallurgy 295, 51–60.
1993. Geometrical and fractal analysis of a three-dimensional vein Peters, S.G., 1993a. Nomenclature, concepts and classification of
network in a fractured granite. Journal of Volcanology and oreshoots in vein deposits. Ore Geology Reviews 8, 3–22.
Geothermal Research 56, 267–280. Peters, S.G., 1993b. Formation of oreshoots in mesothermal gold–
Levingston, K.R., 1972. Ore deposits and mines of the Charters quartz vein deposits: examples from Queensland, Australia. Ore
Towers 1:250,000 sheet area, north Queensland. Geological Geology Reviews 8, 277–301.
Survey of Queensland, Brisbane, Report 57, 103 pp. Peters, S.G., Golding, S.D., 1989. Geologic, fluid inclusion and stable
Lord, D., Etheridge, M.A., Willson, M., Hall, G., Uttley, P., 2001. isotope studies of granitoid-hosted gold-bearing quartz veins,
Measuring exploration success: an alternative to the discovery- Charters Towers, northeastern Australia. In: Keays, R.R., Ramsay,
cost-per-ounce method of quantifying exploration effectiveness. W.R.H., Groves, D.I. (Eds.), The Geology of Gold Deposits —
Society of Economic Geologists Newsletter 45, 1 and 10–16. The Perspective in 1988. Economic Geology Monograph, vol. 6,
Mandelbrot, B.B., 1967. How long is the coast of Britain? Statistical pp. 260–273.
self-similarity and fractional dimension. Science 156, 636–638. Qiu, Y., Groves, D.I., McNaughton, N.J., Wang, L., Zhou, T., 2002.
Mandelbrot, B.B., 1983. The Fractal Geometry of Nature. Freeman, Nature, age and tectonic setting of granitoid-hosted, orogenic gold
New York. 495 pp. deposits in the Jiaodong Peninsula, eastern North China craton,
Marks, E.O., 1913. Outside mines of the Charters Towers goldfield: China. Mineralium Deposita 37, 283–305.
Brisbane. Queensland Geological Survey, Department of Mines Raine, M.D., 2001. Geology, sampling and geostatistical analysis of the B-
Publication, vol. 238. 21 pp. lode, Hadleigh Castle gold mine, Charters Towers, Queensland,
McCuaig, T.C., Hronsky, J.M.A., 2000. The current status and future Australia, Unpublished MSc thesis, University of Wales, Cardiff.
of the interface between the exploration industry and economic Reid, J.H., 1917. The Charters Towers goldfield. Geological Survey of
geology research. In: Hagemann, S.G., Brown, P.E. (Eds.), Gold in Queensland, Brisbane, Publication, vol. 256. 232 pp.
2000. Reviews in Economic Geology, vol. 13, pp. 553–559. Richards, D.N.G., 1980. Paleozoic granitoids of northeastern Aus-
McElhinny, M.W., Powell, C.Mc.A., Pisarevsky, S.A., 2003. Paleo- tralia. In: Henderson, R.A., Stephenson, P.J. (Eds.), The Geology
zoic terranes of eastern Australia and the drift history of and Geophysics of Northeastern Australia. . Geological Society of
Gondwana. Tectonophysics 362, 41–65. Australia, Queensland Division, Brisbane, Proceedings. Third
Mishra, B., Panigrahi, M.K., 1999. Fluid evolution in the Kolar gold Australian Geological Convention, Townsville, pp. 229–246.
field. Evidence from fluid inclusion studies. Mineralium Deposita Ridley, J.R., 1993. The relations between mean rock stress and fluid
34, 173–181. flow in the crust with reference to vein- and lode-style gold
Morrison, G.W., 1988. Palaeozoic gold deposits of northeast Queens- deposits. Ore Geology Reviews 8, 23–37.
land. In: Morrison, G.W. (Ed.), Epithermal and Porphyry Style Rypkema, H.A., 1978. The Ladybird gold mine, Charters Towers,
Gold Deposits in North Queensland. James Cook University, North Queensland. Unpublished B.Sc. (Hons.) thesis, University
Townsville, Contributions of the Economic Geology Research of Queensland, Brisbane, 160 pp.
Unit, vol. 29, pp. 11–22. Scheibner, E., Veevers, J.J., 2000. Tasman fold belt system. In:
Morrison, R.J., Storey, N.J.M., Towsey, C.A.J., 2004. Management of Veevers, J.J. (Ed.), Billion-year Earth History of Australia and
geological risks associated with quartz reef gold deposits, Charters Neighbours in Gondwanaland. GEMOC Press, Macquarie Uni-
Towers, Queensland. In: Dominy, S.C. (Ed.), Mine and Resource versity, Sydney, pp. 154–233.
Geology: Economic Geology Research Unit and Australasian Schreiber, D.W., Amstutz, G.C., Fontbote, L., 1990a. The formation of
Institute of Mining and Metallurgy Symposium. Contribution, auriferous quartz sulfide veins in the Pataz region, northern Peru. A
Economic Geology Research Unit, vol. 62. James Cook synthesis of geological, mineralogical, and geochemical data.
University, pp. 87–106. Mineralium Deposita 25, S136–S140 (Supplement).
80 O.P. Kreuzer et al. / Ore Geology Reviews 32 (2007) 37–80

Schreiber, D.W., Fontbote, L., Lochmann, D., 1990b. Geologic setting, Towsey, C.A.J., Morrison, R.J., Foord, G.E., Storey, N.J.M., 2002. The
paragenesis and physiochemistry of gold quartz veins hosted by plu- Charters Towers gold project: gold production plan. Technical
tonic rocks in the Pataz region. Economic Geology 85, 1328–1347. Report, Charters Towers Gold Mines Ltd, Brisbane, http://www.
Scott, K.M., van Eck, M., 2003. Brahman gold prospect, Charters ctgold.com.au/downloadablefiles/02-09-27GPP-FINAL.pdf, 72 pp.
Towers region, Queensland. Regolith Expression of Australian Ore Turcotte, D.L., 1992. Fractals and Chaos in Geology and Geophysics.
Systems — A Compilation of Geochemical Case Histories and Cambridge University Press, Cambridge. 412 pp.
Conceptual Models. CRC LEME, pp. 1–3. http://crcleme.org.au/ Vearncombe, J., Vearncombe, S., 1999. The spatial distribution of
RegExpOre/Brahman.pdf. mineralization: applications of Fry analysis. Economic Geology
Sibson, R.H., 1990. Conditions of fault-valve behavior. In: Knipe, R.J., 94, 475–486.
Rutter, E.H. (Eds.), Deformation Mechanisms, Rheology and Walshe, J.L., Heithersay, P.S., Morrison, G.W., 1995. Toward an
Tectonics. Geological Society Special Publication, vol. 54, pp. 15–28. understanding of the metallogeny of the Tasman fold belt system.
Sibson, R.H., 1996. Structural permeability of fluid driven fault- Economic Geology 90, 1382–1401.
fracture meshes. Journal of Structural Geology 18, 1031–1042. Wellman, P., 1995. Tasman orogenic system: a model for its
Sillitoe, R.H., 1997. Gold deposits and intrusive rocks. In: Papunen, H. subdivision and growth history based on gravity and magnetic
(Ed.), Mineral Deposits: Research and Exploration, Where do anomalies. Economic Geology 90, 1430–1442.
They Meet? Balkema, Rotterdam, pp. 23–26. Widler, A.M., Seward, T.M., 2002. The adsorption of gold(I)
Sillitoe, R.H., Thompson, F.H., 1998. Intrusion-related gold deposits: hydrosulfide complexes by iron sulfide surfaces. Geochimica et
types, tetono-magmatic settings and difficulties of distinction from Cosmochimica Acta 66, 383–402.
orogenic gold deposits. Resource Geology 48, 237–250. Withnall, I.W., Lang, S.C., 1993. Geology of the Broken River Province,
Solomon, M., Groves, D.I., 2000. The Geology and Origin of North Queensland. Brisbane, Queensland Geology Department of
Australia's Mineral Deposits (reprinted with additional material). Mines and Energy, Queensland Geology, vol. 4. 292 pp.
Centre for Ore Deposits Research, University of Tasmania, and Withnall, I.W., Draper, J.J., Mackenzie, D.E., Knutson, J., Blewett, R.S.,
Centre for Global Metallogeny, University of Western Australia Hutton, L.J., Bultitude, R.J., Wellman, P., McConachie, B.A.,
Publication 32, 1002 pp. Bain, J.H.C., Donchak, P.J.T., Lang, S.C., Domagala, J.,
Stockill, B.D., Hutton, L.J., 1991. Gravity data over the Lolworth– Symonds, P.A., Rienks, I.P., 1997. Review of geological provinces
Ravenswood province: geological and structural implications. In: and basins of north Queensland. In: Bain, J.H.C., Draper, J.J. (Eds.),
Jamieson, W., Pippett, T. (Eds.), Exploration in a Changing North Queensland Geology. Australian Geological Survey Bulletin
Environment. Australian Society of Exploration Geophycisists, 240, and Queensland Department of Mines and Energy Queensland
Eighth Conference and Exhibition, and the Geological Society of Geology, vol. 9, pp. 449–528.
Australia Exploration Symposium, Sydney, Geological Society of Wyborn, L.A.I., Heinrich, C.A., Jaques, A.L., 1994. Australian
Australia, Abstracts 30, pp. 68–69. Proterozoic mineral systems: essential ingredients and mappable
Stolz, A.J., 1995. Geochemistry of the Mount Windsor Volcanics: criteria. Proceedings of the Australian Institute of Mining and
implications for the tectonic setting of Cambro-Ordovician Metallurgy Annual Conference, Darwin, pp. 109–115.
volcanic-hosted massive sulfide mineralization in northeastern Zhou, T., Phillips, G.N., Denn, S., Burke, S., 2003. Woodcutters
Australia. Economic Geology 90, 1080–1097. goldfield: gold in an Archaean granite, Kalgoorlie, Western
Stüwe, K., 1998. Tectonic constraints on the timing and relationships Australia. Australian Journal of Earth Sciences 50, 553–569.
of metamorphism, fluid production and gold-bearing quartz vein Zucchetto, R.G., Henderson, R.A., Davis, B.K., Wysoczanski, R.,
emplacement. Ore Geology Reviews 13, 219–228. 1999. Age constraints on deformation of the eastern Hodgkinson
Tenison-Woods, K.L., Rienks, I.P., 1992. New insights into the province, north Queensland: new perspectives on the evolution of
structure and subdivision of the Ravenswood Batholith — a the northern Tasman orogenic zone. Australian Journal of Earth
geophysical perspective. Exploration Geophysics 23, 353–360. Sciences 46, 105–114.

You might also like