Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/260286782

Microbial influence in the growth of alluvial gold from Watts Gully, South
Australia

Article · January 1993

CITATIONS READS

9 222

1 author:

John L Keeling
Geological Survey of South Australia
205 PUBLICATIONS 1,983 CITATIONS

SEE PROFILE

All content following this page was uploaded by John L Keeling on 23 March 2016.

The user has requested enhancement of the downloaded file.


Is s u e d by ISSN 0584- 3219

THE GEOLOGICAL SURVEY


OF SOUTH AUSTRALIA

Contents

• BAMPTON K.F. — Origin and age of the


Burra copper orebody — a discussion.
• PREISS W.F. and DREXEL J.F. — Origin
and age of the Burra copper orebody —
a reply.
• KEELING J.L. — Microbial influence in the
growth of alluvial gold from Watts Gully,
South Australia.
• MILLER D.T. and SCHURING J. —
The Sandison Subgroup, north of
Buckaringa Gorge, Flinders Ranges.

JUNE 1993 NUMBER 126


Origin and age of the Burra copper orebody —
a discussion
K.F. Bampton*

Drexel and McCallum (1986) proposed a novel model for the origin and age of the Burra
orebody, based on new evidence of ‘an extensive zone o f volcanics identified as a result of
their 1980-81 mapping of the Burra pit in its final weeks of operation. The following
alternative hypothesis derives not from new work but from a review of existing Burra data
and comparison with some other copper deposits in the Adelaide Geosyncline and elsewhere.

Dickinson (1942) divided the east-dipping host sequence to the Burra orebody into four
informal units. These have been included in the Torrensian Skillogalee Dolomite by sub­
sequent authors (Wright, 1975) with slightly differing descriptions and are here designated
from west to east as follows:
1A — complexly folded dolomite and marble, cut off by marble breccia
1B — kaolinised siltstone
1C — laminated-flaggy dolomite
1D — massive dolomite.

The 1A breccia, which occupies the western side of the pit (Figs 1 and 2) has been subject
to varied interpretation over the years. However, a specific study by Ey (1974) clearly
established its diapiric characteristics and Drexel and McCallum (1986) have presented
further supporting data. It is thus not a Skillogalee Dolomite member but intruded material of
undetermined older age.

The contact between the marble breccia and the mineralised 1B siltstone immediately to
the east traditionally has been known as the Kingston Fault. Similarly, the less well-defined
eastern boundary of mineralisation, corresponding to the 1D dolomite contact, has been
known as the Tinline Fault. It is here proposed that Tinline rather than Kingston Fault is the
diapir boundary, making the 1B siltstone, 1C dolomite and included orebodies all part of the
diapir. Only the 1D unit would be part of the host Skillogalee Dolomite. These relationships
are illustrated on Figures 1 and 2, which should be contrasted with Drexel and McCallum’s
equivalent Figures 2 and 3.

The apparent problem of 1C dolomite being so intensely altered as to be ‘reduced to a


friable, porous sandstone/siltstone-like rock’ (IB) by 'dissolution of carbonate, kaolinisation,
potash metasomatism, chloritisation and silicification' is immediately removed — they are
separate lithotypes. There is likewise no requirement for redolomitisation adjacent to the
Kingston Fault, as coarsely recrystallised dolomite juxtaposed to leached rocks, is a common
occurrence in diapirs. The ore textures of jigsaw breccia and relatively intact host are also
compatible with the orebodies being hosted by a number of partially fragmented diapir rafts.
The need for an assumed cross fault (Wright, 1975) in the centre of the pit is also eliminated
— it is a major xenoclast boundary.

Further, the long-standing enigma of the central syenite porphyry, its changing attitude
and the more recently reported extensive zone of adjacent volcanics, are readily
accommodated by the model. Volcanics of generally intermediate composition, but ranging
* Ore Reserve Evaluation Services, Adelaide

2
from acid to basic, are characteristic of Adelaide Geosyncline diapirs, whereas they are
unknown elsewhere in the Skillogalee Dolomite or other rocks of the Burra region. Depth of
cover and age of emplacement problems mentioned by Drexel and McCallum (1986) for their
volcanic vent mode, do not arise.

There are a number of prior local parallels for significant copper mineralisation in diapiric
rafts, including Blinman (Coats, 1964) and Mountain of Light (Dewar and Hatcher, 1975).
On purely lithological grounds, a marble-rich body ‘overlain’ by volcanics within the diapir at
Burra is suggestive of a partly disrupted upper Wywyana Formation/lower Wooltana
Volcanics position. Note that in the Boucaut Volcanics (Wooltana equivalents closest to
Burra), acid lavas and pyroclastics predominate (Preiss, 1987). However, this is not thought
to be critical as many different units are mineralised in other diapirs and mineralisation would
appear to be at least syn and possibly post-diapiric as suggested by discordant mineralisation
in abutting host rocks. Time of emplacement is thus proposed to be post Torrensian (the
diapir intrudes Burra Group rocks) and, if not syn-diapiric, then coeval with the intrusion of
Cambro-Ordovician granitoids which occur 40 km to the east.

3
In addition to explaining some previous difficulties at Burra, the foregoing model also
explains why ore cuts out so suddenly along strike and has the appeal of regional geological
compatibility as opposed to invoking a unique post-Ordovician volcanic centre for which
there is minimal wider corroborative evidence.

KEYWORDS: ECONOMIC GEOLOGY!Structural geologylOre genesisICopper ore/Diapirs/Copper mines!


Tinline Fault!Kingston Fault!Adelaide GeosynclinelSkillogalee DolomitelBurre minelSI 54-05 BURRAI6630-I

4
REFERENCES

Coats, R.P., 1964. The geology and mineralisation of the Blinman Dome Diapir. South Australia. Geological
Survey. Report o f Investigations, 26.
Dewar, G.J. and Hatcher, M.I., 1975. The geology and mineralisation of Copley Diapir, northern Flinders Ranges,
South Australia. In: The Australasian Institute of Mining and Metallurgy Annual Conference, South Australia,
1975. Australasian Institute o f Mining and Metallurgy. Publication Series, 4/75:483-493.
Dickinson, S.B., 1942. The structural control of ore deposition in some South Australian copperfields. South
Australia. Geological Survey. Bulletin, 20.
Drexel, J.F. and McCallum, W.S., 1986. Origin and age of the Burra copper orebody. South Australia. Geological
Survey. Quarterly Geological Notes, 98.
Ey, J.L., 1972. Investigations of the 1A breccia, Burra Mine SA. Poseidon Ltd Report, 3/72 (unpublished).
Preiss, W.V. (Compiler), 1987. The Adelaide Geosyncline — late Proterozoic stratigraphy, sedimentation,
palaeontology and tectonics. South Australia. Geological Survey. Bulletin, 53.
Wright, R.G., 1975. Burra copper deposit, South Australia. In: Knight, C.L. (Ed.), Economic geology of Australia
and Papua New Guinea, 1, Metals. Australasian Institute o f Mining and Metallurgy. Monograph Series,
5:1030-1044.

Origin and age of the Burra copper orebody —


a reply
W.V. Preiss and J.F. Drexel

INTRODUCTION
Bampton (this QGN) has proposed an interesting solution to a number of enigmatic
aspects of the Burra orebody (Fig. 1). Like the volcanic-neck model suggested by Drexel and
McCallum (1986), his alternative explanation focuses on the unique geological characteristics
of the mine area, features that are not seen in the country rocks along strike. Unfortunately,
the critical evidence to verify the two models is no longer accessible. Dykes of intrusive
feldspar porphyry have been mined out entirely. The volcanic rocks crucial to Drexel and
McCallum’s interpretation, and the breccias interpreted by Bampton as an extension of the
diapiric material west of the Kingston Fault, are restricted to the deepest part of the open cut,
which is now flooded.

Since publication of Drexel and McCallum’s paper, one of us (WVP) has been carrying
out detailed regional mapping of the eastern BURRA map area. The purpose of this reply is
to outline some of the results of this mapping and its relevance to interpretation of the Burra
orebody.

STRATIGRAPHY
Regional mapping has confirmed that the stratigraphy recognised by Dickinson (1942) can
be traced consistently between Burra and Robertstown. Today, these carbonate units
(described by him as ‘limestone’) are assigned to the Skillogalee Dolomite, consisting of a
lower, dominantly pale grey or cream dolomite overlain by a blue-grey, commonly
cryptalgal-laminated dolomite upper member, with irregular black chert replacement. The
lower member contains a distinctive marker unit characterised by very thinly laminated dark
grey limestone but also containing some thin dolomite interbeds. It may have been one of
these interbeds that was sampled by Dickinson (1942) for chemical analysis, which led him to

5
6
describe the marker as ‘dolomite’. Closely associated with the limestone marker, and
generally stratigraphically below it, is a dark grey, laminated, calcareous and carbonaceous
siltstone and very fine-grained sandstone. It should be noted that Dickinson regarded the
breccias at Burra as of tectonic origin, and not as part of the country rock stratigraphy, a view
with which we concur.

STRUCTURE
Contrary to the implication by Bampton, the Kingston Fault, and not the Tinline Fault, is
the dominant structural feature of the Burra area. The Kingston Fault, as exposed in the open
cut, is a north northwest trending, steeply east-dipping fault and shear zone; local
subhorizontal slickensides indicate that at least the last phase of movement is likely to have
involved strike-slip displacement. The Kingston Fault has been mapped as having a strike
extent well beyond the limits of the pit, at least as far as the golf course 1.5 km to the north
and about 30 km along the spine of the range to the south, towards Robertstown. The Tinline
Fault, on the other hand, has very limited extent and displacement, and has not been
recognised outside the open cut (Fig. 2).

The Kingston Fault is also important in that, regionally, it forms the boundary between
highly recrystallised pale-coloured dolomite (lower Skillogalee) to the west and fine-grained,
well preserved dolomites (both lower and upper members) to the east. The mid-greenschist
grade of regional metamorphism does not change appreciably throughout the eastern Nackara
Arc, and the reason for this contrast in recrystallisation remains unclear. It is possible,
however, that the Kingston Fault has juxtaposed slightly different crustal levels within the
Skillogalee Dolomite.

Folding in the Burra vicinity east of the Kingston Fault is characterised by gently to
moderately steeply north-plunging en echelon meso to macro-scale folds with steep axial
surfaces and sinistral vergence (there are also minor plunge reversals). The Kingston Fault is
part of an extensive lineament (Fig. 1) interpreted as an expression of the eastern boundary of
the Q2 corridor of O ’Driscoll (1982), which Preiss (1987) suggested acted as a wrench
system to which the upright folds of the Nackara Arc could be related. The en echelon folds
at Burra are believed to be an expression of wrench deformation, suggesting that the Kingston
Fault suffered convergent, sinistral strike-slip displacement during the main phase of folding
in the Delamerian Orogeny. The more irregular fold style in the Skillogalee Dolomite west of
the fault may be due to complication by the intrusion of diapiric breccia and by the
accommodation of buckled dolomite beds scraped off by wrench movement on the Kooringa
Fault, a splinter branching to the west from the Kingston Fault and truncating a regional
syncline in Umberatana Group rocks (Fig. 2).

A consequence of the fold style within the open cut is the possible thickening of some
units in fold hinges. The calcareous siltstone bed below the limestone marker unit is well
exposed on the northern haul road of the Burra Mine (Fig. 3), where it dips steeply east and is
only a few metres thick. It can be traced, through a series of meso-scale folds, into the area of
the orebody, where it is intensely altered, and through to the southern edge of the pit, where it
is again relatively fresh and very gently dipping.

7
Figure 2. Geology o f the Burra township and environs.

8
INTERPRETATION
Mapping at the bottom of the pit by Drexel and McCallum (1986) showed that relict
bedding is preserved in the intensely leached ‘friable, porous sandstone/siltstone-like rock’
and that this is commonly gently dipping. Derivation of such a material from the leaching of
a relatively pure dolomite presents a problem. Bampton suggests that the leached rock is
explained by not being part of the Skillogalee Dolomite, but instead a diapiric xenoclast
carrying copper mineralisation in a manner analogous to the cupriferous dolomite at Blinman.
On the contrary, we suggest that the protolith of the leached rock may have been the
calcareous siltstone unit below the limestone marker in the lower Skillogalee Dolomite,
tectonically thickened in a fold hinge.

Another enigmatic feature of the Burra Mine is a body of hard, coarsely crystalline
carbonate rock with 20-30 mm dolomite rhombs, carrying copper mineralisation, adjacent to
the Kingston Fault at the bottom of the pit. Drexel and McCallum (1986) interpreted this as
due to redolomitisation of the ore and leached host rock, while Bampton regards it as another
diapiric xenoclast. We do not necessarily discount the latter alternative, or the possibility that
it represents a lower, recrystallised level of the Skillogalee Dolomite situated in an anticlinal
core. However, petrographic evidence (Farrand, 1983) strongly supports a replacement origin
for this carbonate, which is massive and structureless. In contrast, bedding is generally
preserved in marble of the Skillogalee Dolomite west of the Kingston Fault, and indeed in
most carbonate xenoclasts in diapirs.

CONCLUSION AND ECONOMIC IMPLICATIONS


There are two objections to Bampton’s interpretation. Firstly, the Kingston Fault, and not
the Tinline Fault, is the dominant structural break at Burra. As such, it would have to
juxtapose precisely two small diapiric bodies (both within the confines of the open cut), one
unmineralised and exposed on the west side of the pit, and the other mineralised and altered,
to the east. Our proposal retains the Kingston Fault as the boundary between diapiric breccia
to the west and hydrothermally altered and mineralised Skillogalee Dolomite to the east.
Secondly, Drexel and McCallum’s mapping during mining demonstrated that the porphyry
intrudes Skillogalee Dolomite (Fig. 3; Plate 1), and not a xenoclast in a diapir.

The only difference from Drexel and McCallum’s (1986) interpretation is that the host
rock for mineralisation is the tectonically thickened calcareous siltstone unit below the
extensive limestone marker, rather than leached dolomite, which could not account for the
large amount of silt and sand residue in the mineralised material. We agree with Bampton
that there is no need for a cross-fault as suggested by Wright (1975); the northern limit of
mineralisation coincides with the limit of hydrothermal alteration.

Both the model of Drexel and McCallum (1986) and that of Bampton involve special
circumstances not seen in other parts of the Nackara Arc, in the one case a unique volcanic
vent, in the other a unique mineralised xenoclast in a small diapir. Despite exhaustive
searching by one of us (WVP) along the extensive breccias associated with the southern
extension of the Kingston Fault, it has not been possible to find any evidence of a diapiric
body other than that at Burra. The breccias along the extension of the Kingston Fault are all
explicable by brittle fracturing in the Skillogalee Dolomite. On the other hand, structural
evidence suggests that the Kingston Fault acted as a sinistral wrench during folding, and was
subsequently reactivated as a strike-slip fault after formation of the orebody. It has not yet
been possible to determine unambiguously the sense of movement of this latest displacement,

9
Figure 3. Geology o f the Burra Mine (based on unpublished mapping byJ.F . D rexeland
W.S. M cCallum) showing location o f the porphyry dyke and interpreted volcanic rocks.

but the orebody appears to have been cut off. There is therefore a possibility that the
displaced portion of the orebody might be preserved somewhere along strike on the western
side of the Kingston Fault. If the latest movement is also sinistral, there is a chance that this
‘other half of the orebody might be located south of Burra, yet no evidence for it has been
found during detailed regional mapping. On the other hand, if the latest displacement were
dextral, the ‘other half might be expected to occur to the north of Burra, perhaps in one of the
large areas of Quaternary alluvial cover.

10
Plate 1. Porphyry dyke intruding gently dipping dolomite o f the lower member, Skillogalee Dolomite,
in the lower benches o f the Burra Mine. (Photo no. 41841)

KEYWORDS: ECONOMIC GEOLOGY!Regional geology!Structural geologyIDiapirslOre genesis! Copper ore!


Geological mappingICopper minesITinline Fault!Kingston Fault!Adelaide GeosynclinelSkillogalee Dolomite!
Burra MinelSI 54-05!Burra 6630-1.

REFERENCES

Dickinson, S.B., 1942. The structural control of ore deposition in some South Australian copperfields. South
Australia. Geological Survey. Bulletin, 20.
Drexel, J.F. and McCallum, W.S., 1986. Origin and age of the Burra copper orebody. South Australia. Geological
Survey. Quarterly Geological Notes, 98.
'Farrand, M.G., 1983. Comments on the petrography of forty-two thin sections of samples from the Burra copper
mine, SA. South Australia. Department o f Mines and Energy. Report Book, 83/28.
O’Driscoll, E.S.T., 1982. Patterns of discovery the challenge for innovative thinking. Petroleum Exploration o f
Australia. Journal, 1:11-31.
Preiss, W.V., (Compiler) 1987. The Adelaide Geosyncline late Proterozoic stratigraphy, sedimentation,
palaeontology and tectonics. South Australia. Geological Survey. Bulletin, 53.
Tucker, D.H. and Collerson, K.D., 1972. Lamprophyric intrusions of probable caibonatitic affinity from South
Australia. Geological Society o f Australia. Journal, 19:387-392.
Wright, R.G., 1975. Burra copper deposit, South Australia. In: Knight, C.L. (Ed.), Economic geology of Australia
and Papua New Guinea, 1, Metals. Australasian Institute o f Mining and Metallurgy. Monograph Series,
5:1039-1044.

11
Microbial influence in the growth of alluvial
gold from Watts Gully, South Australia
J.L. Keeling

INTRODUCTION
Scanning electron microscope (SEM) examination of three sand-sized gold grains from
clayey alluvium at Watts Gully, 40 km northeast of Adelaide, showed evidence of
bacterioform gold in two of the grains. The structure of gold pseudomorphs of bacteria-like
colonies, and the morphology of individual cells resemble Pedomicrobium-like, budding
bacteria. The presence of gold-coated bacteria-like cells indicate that chemical processes
were involved in the near surface formation of alluvial gold at Watts Gully, and raise
speculation about the role of metal precipitating bacteria in the localised accumulation of
placer gold in the Adelaide Hills.

Natural accumulations of gold in eluvial and alluvial placer deposits have provided some
of the earliest, richest and most accessible sources of gold, supplying more than two thirds of
the world’s total gold production (Boyle, 1987). Yet the nature of gold in placer deposits and,
in particular, the origin of gold nuggets have been the subject of debate and scientific
argument for over a hundred years (Boyle, 1987). In particular, the coarse size of some gold
nuggets and characteristic high gold content of many placer grains often differ substantially
from that of gold in primary source rocks from which the deposits were derived. Wilkinson
(1867), Newbery (1868) and Egleston (1881) considered that nuggets, and alluvial gold
generally, were formed by chemical precipitation from groundwater circulating within the
deposits. The theory of formation of nuggets by chemical accretion was a widely held view
in the late 1800s and early 1900s, a view embraced by Maclaren (1908) in his treatise on the
geological occurrence and geographical distribution of gold. These views were opposed by
Liversidge (1893), Lindgren (1911, 1933), MacKay (1921) and others who regarded placer
gold as detrital and derived by weathering or mechanical disintegration of primary deposits.
Any addition by chemical precipitation they regarded as insignificant. The theory of detrital
origin of placer gold has dominated the literature since the early 1900s. Increase in fineness
(gold content) in placer gold has been regarded as due principally to dissolution of more
soluble silver, while the coarse size of nuggets reflected either variation in the nature of
primary gold distribution, or was the result of mechanical effects of flattening, agglomeration
and rounding during transportation.

Supergene mobilisation of gold, either as a chloride complex under acidic, oxidising


conditions or as a thiosulphate complex under alkaline, reducing conditions, is recognised as
an important process in gold ore enrichment in weathered profiles (Emmons, 1917; Cloke and
Kelly, 1964; Lakin etal., 1974; Mann, 1984; Webster and Mann, 1984; Webster, 1986;
Krupp and Weiser, 1992). Inorganic gold complexes in solution are stable only under
specific chemical conditions and are readily reduced with changes in Eh or pH, resulting in
precipitation of metallic gold. In this regard, chloride complexes are much less stable than
thiosulphate complexes (Lakin, et al., 1974). Freise (1931) documented examples from
Brazil where replenishment of alluvial gold in gravel deposits apparently formed by chemical
precipitation from groundwater. In experiments using organic-rich water, Freise
demonstrated that dissolution and transportation of gold was controlled by humic acid in the
water. Following investigations of the effect of organic acids on gold, Ong and Swanson
(1969) concluded that organic acids have the capacity to reduce gold chloride solutions to
form organic-protected colloids of metallic gold that were stable, and capable of being

12
transported in the general range of composition of natural waters. Baker (1973,1978)
showed that humic acid was both an effective solvent of gold metal and a mobilising agent for
precipitated metal humates. Humic acids form the bulk of the organic matter present in many
Tasmanian podzolic soils, and under cool temperature, high rainfall conditions, could be the
dominant cause of high mobility of metals in the soil (Baker, 1973). The significance of
gold-organic complexes in mobilising gold metal were confirmed by Bowell et al. (1993) in a
study of the role of fulvic acid in the dissolution and transportation of gold in tropical rain
forest soils at the Ashanti mine in Ghana.

Recognition of dissolution and mobilisation of gold under atmospheric conditions revived


early ideas on chemical formation of gold nuggets (Wilson, 1984; Boyle, 1987) and also
prompted consideration of the role of microorganisms in the geochemical cycle of gold
(Lakin et al., 1974; Beveridge et al., 1983). Microbial involvement in the deposition and
concentration of gold during Precambrian times has been argued for carbonaceous zones in
the rich Witwatersrand placer deposits in South Africa (Hallbauer and van Warmelo, 1974;
Hallbauer, 1975).

The most compelling evidence of biological contribution in the precipitation of gold in


placer deposits was reported by Watterson (1992), and Brooks and Watterson (1992). Placer
gold particles from Alaskan deposits show a high proportion of grains with lacelike networks
of micrometre filiform gold interpreted as low-temperature pseudomorphs of a
Pedomicrobium-\ike budding bacteria (Watterson, 1992). Based on SEM studies of 18 000
gold particles from Lillian Creek, together with several hundred particles from eight other
Alaskan placer deposits, Watterson (1992) concluded that the majority of placer gold particles
appeared to include gold that had accumulated in and on the cells of common bacteria.
Assuming that Watterson’s interpretations are correct, then the presence of bacterioform
pseudomorphs on gold grains provides readily observed evidence of chemical accretion and
probable growth of gold under near-surface conditions. The bacteria species responsible for
gold deposition remain to be identified and as such, the range of environmental conditions
conducive to survival and growth of the bacteria are uncertain. That range may prove to be
broad given that similar bacterioform gold was recorded for samples of filigree gold from
possible mesothermal quartz veins in the Xiniiang-Uygur Autonomous Region of China
(Watterson, 1992).

ADELAIDE HILLS PLACER GOLD


Placer gold in the Adelaide Hills has been mined since the 1840s, mainly from small but
widespread deposits formed by weathering and erosion of Palaeoproterozoic inliers, and
Neoproterozoic to Cambrian metamorphic rocks of the Adelaide Geosyncline. A common
feature of many deposits was their initial development as small but payable workings in soil
or shallow alluvial sediments, often in restricted catchments. Prospecting by costeans, shafts
and drives on nearby gold-bearing quartz veins generally showed these to be of marginal
grade. Larger deposits, including Barossa (100 000 oz) and Echunga (130 000 oz), were
mostly in gravels of probable early Tertiary age where gold accumulated usually near or at
the contact with underlying clay of deeply weathered bedrock (Robertson, 1991). No
systematic work has been done to reconstruct Tertiary drainage patterns in these areas and
therefore the provenance of the gold is uncertain, although widely believed to have been
derived from local bedrock sources. The presence of microbial gold provides a means of
discriminating between secondary and primary gold in placer deposits and could explain the
presence of coarse gold in placers for which no comparable gold has been found in nearby
bedrock.

13
Gumeracha Goldfield
Gold at the Gumeracha workings, 40 km northeast of Adelaide, was found mostly as nug­
gets in alluvium and soil in ephemeral creeks, and as filamentous gold in deeply weathered
schist of Adelaidean age (Fig. 1). The goldfield was discovered by Watts in early 1884 and
was worked quietly for about a year (Woodward, 1886). Reports of recovery of good quanti­
ties of gold led to a rush in 1885 when much of the mining and prospecting was done. Best
returns were from shallow alluvial deposits in wooded creeks draining coarse, crumpled
schist and gneiss that form a broad, 500-1 000 m wide zone of sheared and metasomatically
altered rocks, probable equivalents of basal Burra Group sediments of early Adelaidean age
(Mills, 1973). Watts Gully was established as the first payable gold diggings and proved to
be by far the richest of the workings on the Gumeracha Goldfield. According to Woodward
(1886), the gully:
was very rich to a little way above the fork near the top, and particularly rich just at it, where some large
nuggets were obtained, which as far as can be gathered, were not taken from the deepest gutter but rather
upon the western bank. Lately, some nice cemented nuggets were obtained behind the blacksmith’s shop;
these had been so little rolled that they could not have travelled far from the reef.

Some 1 000 oz (31 kg) of gold were reportedly produced from the Gumeracha Goldfield,
with the bulk of production as eluvial or alluvial gold from Watts Gully. The richness of the
gully, restricted gold distribution, and character of the gold suggested a nearby reef source.
Quartz veins near the alluvial workings and at the head of the gully were prospected by
drives, shafts and costeans but no payable gold was found (Woodward, 1886).

In June 1933, a gold nugget weighing 20 oz (620 gm) was found by Kollosche while plant­
ing young pine trees 2 km east of Watts Gully (Fig. 1). The area was worked by Kollosche,
Barron and others, and a small amount of gold was recovered from near the surface. Some
fine specimens of gold were found associated with quartz crystals but the best values were
from nodules of schist and kaolin held together by wire gold (Cornelius, 1935).

The historical accounts suggest enrichment and concentration of gold in the regolith with
nugget formation in shallow soil and alluvium, particularly in the upper part of Watts Gully.
Unlike the alluvial deposits at Echunga and Barossa, the Gumeracha Goldfield has no signifi­
cant Tertiary fluviatile sediments to complicate possible interpretations of gold redistribution
and concentration. As such. Watts Gully was a logical choice for a preliminary investigation
in search of microbial pseudomorphs to confirm near surface chemical accumulation of gold.

PROCEDURES
Samples
Four sand-sized grains of gold were recovered by panning fine sand and clayey alluvium
to a depth of 0.5 m on the drainage line of Watts Gully, 50 m down stream of the top fork
(Fig. 1). Three of the grains were relatively clean and these were selected for examination.

Pre-treatment o f gold particles


Pre-treatment followed closely that described by Watterson (1992). Samples were
immersed in concentrated hydrofluoric acid (40% HF) for six days at around 25°C, then
rinsed and dried. They were then immersed in concentrated nitric acid (70% F1N0 3 ) in a
Teflon-lined stainless steel acid-digestion bomb and heated to 190°C for six hours. The gold
was removed, rinsed in distilled water and ethanol, and dried. This treatment removed
surface oxide coatings, quartz and clay but, according to Watterson (1992), does not
perceptibly etch the gold, at the scale of investigation.

14
Figure 1. Gumeracha alluvial gold diggings.

15
Samples were mounted onto an aluminium stub that had been coated with plastic
conductive carbon cement ‘Leit-c-Plasf. This provided good electrical conduction and
allowed the gold particles to be examined directly without the need for any surface coating.
Samples were examined using a Cambridge Stereoscan S250 SEM fitted with a Link System
energy dispersive X-ray analyser (EDX).

RESULTS
Apparent bacterioform pseudomorphs in the form of open clusters of micron-sized,
spheroidal, cell-like particles, hereafter called cells, were observed on two of the three gold
grains. Morphology of the gold grains was blocky, subrounded to subangular, irregular and
vuggy (Plate 1). Dimensions of the two grains showing bacterioforms were determined from
micrographs as 600x530 pm and 780x450 pm respectively. Gold-coated cells, or cells
replaced by gold, were best preserved as filling or lining cavities and in surface depressions
(Plates 2, 3). Cells were observed also as small clusters on surface projections. The
proportion of bacterioform gold was similar for both grains and was estimated at about
25-30% of the surface area. Bacteria-like cells were not present on the surface of the third
gold grain, although in one surface depression, numerous rod-shaped particles, 1-8 pm long
and 0.1-0.2 pm diameter, of unknown origin, were observed.

Bacterioform gold comprised spheroidal, oval and composite cells. These were clustered
together in irregular-shaped colonies that included single cells attached to the surface of gold
and, more commonly, linked cells that projected outward from the surface in an open
branching network (Plates 2 and 4). Mature cells were slightly elongated and generally in the
size range 0.7-1.5 pm diameter. Smaller, terminal cells were spheroidal and typically 0.3-0.5 pm
diameter. In some instances, these were clearly attached to larger cells by a short slender rod
hypha (Plate 5). Size, shape and structure of the bacteria pseudomorphs closely resembled
those described by Watterson (1992) from Alaskan placer gold and assigned by him to
Pedomicrobium-like budding bacteria. On the basis of morphological similarities, the Watts
Gully gold-coated bacteria are probably of the same genus.

Spot EDX analysis of the surface of bacteria-like cells and on adjacent gold surfaces
showed predominantly gold with only trace amounts of silver. Analysis of the grain that
showed no obvious bacterioform gold had slightly higher silver content.

DISCUSSION AND CONCLUSIONS


The presence of a substantial proportion of bacterioform gold on two of only three gold
grains examined is significant in that it indicates that chemical-biological processes played a
role in the precipitation and probable concentration of gold at Watts Gully. The almost
identical nature of the bacterial forms with those reported from Alaska was unexpected, given
the differences in the present environment between Alaska and southern South Australia.
Further work in isolation and identification of the species involved in gold precipitation is
indicated and could provide useful information on the conditions for gold accumulation and
the nature of gold transport.

The bacteria genus Pedomicrobium are small soil microbes comprising individual oval or
spherical cells 0.4-2.0 x 0.4-2.5 pm with up to five or more hyphae per cell body.
Multiplication is by budding at the hyphal tips and mature buds either separate from the
hyphae as uniflagellated swarmers or remain attached (Staley and Fuerst, 1989). The bacteria

16
Plate 1. Alluvial gold grain from Watts Gully Plate 2. Detail o f lower right-hand corner o f
with vuggy surface partially covered by Plate 1 showing open branching bacterioform
bacterioform gold. Scale bar 200 pm. gold. Scale bar 10 pm. (Photo no. 41843)
(Photo no. 41842)

Plate 3. Surface o f gold grain extensively Plate 4. D etail o f bacteria-like cells attached
covered with bacterioform gold. Scale bar to the surface o f grain. Note partial burying
20 pm. (Photo no. 41844) o f cells, lower left, by accumulating gold.
Scale bar 10 p m . (Photo no. 41845)

Plate 5. Gold pseudomorphs o f bacteria-like Plate 6. D etail o f branching network o f


cells. Note sm all rounded bud attached by rounded and oval cells that appear to have
slender hyphal stalk, centre left. Scale bar grown by direct budding and possibly cell
4 p m . (Photo no. 41846) division. Scale bar 2 pm. (Photo no. 41847)

17
are characterised by their ability to accumulate iron and manganese oxides on both the mother
cells and hyphae. This accumulation most probably takes place on or in the cell walls (Mann,
1992) The presence of hyphae, in both lateral positions and at the cell poles are distinctive
for this group of bacteria. Reproduction from hyphal stalks may be a useful mechanism that
enables sustained growth of the colony as older cells become encrusted by accumulated metal
or metal oxides.

Only rare examples of single hypha attached to cells were observed in the Watts Gully
samples. These were typically short, about 0.15-0.2 pm length by 0.1-0.15 pm width
(Plate 5). Cells appeared to have grown mostly by direct budding of the mother cell or, in
some instances, possibly by division of single mother cells (Plate 6). Both forms of
reproduction have been reported as occasionally observed in Pedomicrobium (Staley and
Fuerst, 1989). Watterson (1992) recorded evidence of lateral hyphae on cells of bacterioform
gold but conceded that these were not common and suggested that preferential survival (or
preservation) of the more robust metallised direct-budding cells over cells attached to slender
hyphal stalks might be the reason. In some instances, short hyphae between cells might be
obscured by subsequent metal encrustation.

The extent to which bacteria controlled gold precipitation and growth of gold particles was
not clear from the micrographs. Bacterioform gold covered less than one third of the grain
surface with the remaining surface being extremely fine-grained and varying from smooth to
lumpy. Bacteria colonies showed areas with gold infilling between cells that in places
obscured individuals and groups of cells (Plates 2 and 5). An irregular, lumpy gold surface
adjacent to some sites of attached colonies (Plate 4) gave the appearance of cells buried below
the surface. It might be inferred therefore, that bacterioform gold underpinned substantially
more of the grain structure than was apparent on the surface.

The presence of pseudomorphs of bacteria-like cells in gold from Watts Gully concurs
with Watterson’s suggestions that bacteria have a role in the formation and growth of gold
grains, and that bacterioform gold is widespread. The results also add support to early
theories of chemical processes involved in the formation of placer gold. The overall
contribution of chemical-bacterial gold to the accumulation of gold in alluvial deposits in the
Adelaide Hills remains to be tested, as does the range of mineralising environments in which
bacterioform gold might be found. The observation and interpretation by Watterson (1992)
of bacteria-like pseudomorphs intimately associated with gold grains is a significant
breakthrough that could provide the basis for a substantial reassessment of the dominant
processes in the geochemistry of gold in supergene environments.

ACKNOWLEDGMENTS
Colleagues at CSIRO Soils provided enthusiasm and encouragement for this preliminary
investigation. In particular, acknowledgment is made of M. Raven for initiating and assisting
with collection of gold samples, S. McClure for the SEM micrographs and for critical
discussion, and R. Foster and R. Moen for background on the nature and role of soil bacteria.

KEYWORDS: ECONOMIC GEOLOGY'/Geochemistry/Ore genesis/Gold placer deposits/Supergene deposits/


BacterialPseudontorphism/Adelaide Geosyncline/Burra Group/Gumeracha Goldfield/Watts Gully/Mount Lofty
Ranges!SI 54-091ADELA1DE/6628.

18
REFERENCES

Baker, W.E., 1973. The role of humic acids from Tasmanian podzolic soils in mineral degradation and metal
mobilization. Geochimica et Cosmochimica Acta, 37:269-281.
Baker, W.E., 1978. The role of humic acid in the transport of gold. Geochimica et Cosmochimica Acta,
42:645-649.
Beveridge, T.J., Meloche, J.D., Fyfe, W.S. and Murray, R.G.E., 1983. Diagenesis of metals chemically complexed
to bacteria: laboratory formation of metal phosphates, sulfides and organic condensates in artificial sediments.
Applied and Environmental Microbiology, 45:1094-1108.
Bowell, R.J., Gize, A.P. and Foster, R.P., 1993. The role of fulvic acid in the supergene migration of gold in
tropical rain forest soils. Geochimica et Cosmochimica Acta, 57:4179-4190.
Boyle, R.W., 1987. Gold: history and genesis o f deposits. Van Nostrand Reinhold, New York.
Brooks, R.R. and Watterson, J.R., 1992. The noble metal biogeochemistry of microorganisms. In: Brooks R.R.
(Ed.), Noble metals and biological systems metals. CRC Press, Florida, pp. 159-196.
Cloke, P.L. and Kelly, W.C., 1964. Solubility of gold under inorganic supergene conditions. Economic Geology,
59:259-270.
Cornelius, H.S., 1935. Mount Crawford gold prospecting. Mining Review, Adelaide, 61:79.
Egleston, T., 1881. The formation of gold nuggets and placer deposits. American Institute of Mining Engineers.
Transactions, 9:633-646.
Emmons, W.H., 1917. Gold. In: The enrichment of ore deposits. United States. Geological Survey. Bulletin,
625:305-324.
Freise, F.W., 1931. The transportation of gold by organic underground solutions. Economic Geology, 26:421 -431.
Hallbauer, D.K. and van Warmelo, K.T., 1974. Fossilized plants in thucholite from Precambrian rocks of the
Witwatersrand, South Africa. Precambrian Research, 1:199-212.
Hallbauer, D.K., 1975. The plant origin of the Witwatersrand ‘carbon’. Minerals Science and Engineering,
7:111-131.
Krupp, R.E. and Weiser, T., 1992. On the solubility of gold-silver alloys in the weathering environment.
Mineralium Deposita, 21:268-215.
Lakin, H.W., Curtin, G.C., Hubert, A.E., Shacklette, H.T. and Doxtader, K.G., 1974. Geochemistry of gold in the
weathering cycle. United States. Geological Survey. Bulletin, 1330.
Lindgren, W., 1911. The Tertiary gravels of the Sierra Nevada of California. United States. Geological Survey.
Professional Papers, 73.
Lindgren, W., 1933. Mineral deposits. 4th edition. McGraw Hill, New York.
Liversidge, A., 1893. On the origin of gold nuggets. Royal Society o f New South Wales. Journal, 27:303-343.
MacKay, B.R., 1921. Beauceville map-area, Quebec. Canada. Geological Survey. Memoirs, 127.
Maclaren, J.M., 1908. Gold: its geological occurrence and geographical distribution. Mining Journal. London.
Mann, A.W., 1984. Mobility of gold and silver in lateritic weathering profiles: some observations from Western
Australia. Economic Geology, 79:38-49.
Mann, S., 1992. Bacteria and the Midas touch. Nature, 357:358-360.
Mills, K.J., 1973. The structural geology of the Warren National Park and the western portion of the Mount
Crawford State Forest, South Australia. Royal Society o f South Australia. Transactions, 97:281-315.
Newbery, J.C., 1868. Introduction of gold to, and the formation of nuggets in, the auriferous drifts. Royal Society
o f Victoria. Transactions and Proceedings, 9:52-60.
Ong, H.L. and Swanson, V.E., 1969. Natural organic acids in the transportation, deposition, and concentration of
gold. Colorado School o f Mines. Quarterly, 64:395-425.
Robertson, R.S., 1991. Major South Australian gold deposits — summaries. South Australia. Department o f Mines
and Energy. Report Book, 91/66.
Staley, J.T. and Fuerst, J.A., 1989. Budding and/or appendaged bacteria. In: Staley, J.T., (Ed.), Bergey’s manual o f
systematic bacteriology. Williams and Wilkins, Baltimore, 3:1890-1993.
Watterson, J.R., 1992. Preliminary evidence for the involvement of budding bacteria in the origin of Alaskan
placer gold. Geology, 20:315-318.
Webster, J. and Mann, A.W., 1984. The influence of climate, geomorphology, and primary geology on the
supergene migration of gold and silver .Journal o f Geochemical Exploration, 22:21-42.
Webster, J., 1986. The solubility of gold and silver in the system Au-Ag-S-02-H20 at 25°C at 1 atm. Geochimica
et Cosmochimica Acta, 50:1837-1845.
Wilson, A.F., 1984. Origin of quartz-free gold nuggets and supergene gold found in laterites and soils — a review
and some new observations. Australian Journal o f Earth Sciences, 31:303-316.
Wilkinson, C., 1867. On the theory of the formation of gold nuggets by drift. Royal Society o f Victoria.
Transactions and Proceedings, 8:11 -15.
Woodward, H.P., 1886. Notes on the geological map of Gumeracha and Mount Crawford Goldfields. South
Australia. Parliamentary Paper, 62.

19
The Sandison Subgroup,
north of Buckaringa Gorge, Flinders Ranges
D.T. Miller* and J. Schuring**

INTRODUCTION
The Sandison Subgroup comprises the Nuccaleena For­
mation, Seacliff Sandstone, Brachina Formation and ABC
Range Quartzite (Dyson, 1992). Regional cycles identified
by Dyson in the subgroup have not been previously utilised
in mapping the subgroup in the southern Flinders Ranges,
in an area 30 km north of Quorn between Warren Gorge
and ‘Partacoona’ (Fig. 1). In the vicinity of ‘Partacoona’,
the Sandison Subgroup contains four distinct upward-coars­
ening cycles, each including several sub-cycles. Each cycle
has been mapped for 10 km of strike length.

PREVIOUS WORK
Plummer (1978) described the stratigraphy in the
Buckaringa Gorge and Middle Gorge (Fig. 1) areas as a suc­
cession of grey to purple siltstone of the Brachina Forma­
tion, overlain by ABC Range Quartzite. The ABC Range
Quartzite intertongues with thin shale and quartzite beds.
Plummer (1978, 1989) also recognised the cyclic nature of
the Brachina Formation and ABC Range Quartzite and identified four or five major upward-
coarsening cycles. It is these cycles that Dyson (1992) has recognised regionally in the San­
dison Subgroup.

CYCLES WITHIN THE SANDISON SUBGROUP


Cyclic sedimentation in the Sandison Subgroup is well displayed along Willochra Creek
near ‘Partacoona’ (Fig. 2). A transect through this area (Fig. 2, A-A) shows four well-defined
upward-coarsening successions. Each cycle can be traced laterally and correlates with the
cycles referred to but not mapped by Plummer (1978) in Middle Gorge. A schematic section
(Fig. 3) summarises the nature of the four cycles.

The first cycle is composed of purple to red-brown shale and siltstone and a thick quartzite
succession which comprises a basal massive purple quartzite overlain by white laminated and
cross-bedded quartzite beds. The characteristics of the shale are similar to shale in the lower
to middle Brachina Formation (Plummer, 1978). Units two and three are characterised by a
succession of thinly bedded red-brown siltstone interbedded with progressively thicker and
more dominant beds of light brown to white quartzite. The quartzite beds within each
subcycle thicken from 5 mm to several metres. Plummer (1978) has correlated similar strata
in the Buckaringa area with the upper Brachina Formation. Units 2 and 3 typically contain
several 20-30 m thick sub-cycles (Plate 1). Smaller cycles can be identified within each of
the sub-cycles. Unit 4 is a sandier sequence as it is dominated by shale and quartzite
interbeds about 50-300 mm thick. The shale units progressively thin up-section and are
replaced by several thick laminate and cross-bedded quartzite beds.

* Earthsciences Department, Flinders University of SA


** Eberhard—Karls—Universitat, Tubingen, Germany

20
The units defined here represent four major regressive cycles within the Sandison
Subgroup. The characteristics of the four cycles identified in the Willochra Creek transect
can be traced for about 10 km to the north.

Similar cycles are exposed in Buckaringa and Middle Gorges but the lack of distinct
outcrop between the gorges and the Willochra Creek area has not established reliable
correlation. However, subtle contrasts on aerial photography suggest that the cycles do occur.
Rollings (1986) identified several shale-quartzite repetitions approximately 10 km northwest
of ‘Partacoona’ but attributed these to structural repetition. Re-examination of this area
shows that the cycles identified in the Willochra Creek area are preserved between the faults
mapped by Rollings (1986). The cycles identified in the Sandison Subgroup north of
Buckaringa Gorge correlate with the cycles identified by Dyson (1992).

21
Plate 1. Paracycles in the lower
h a lf o f Cycle 2 in Warrakimbo
Gorge, Willochra Creek.
Quartzite beds become more
dominant in each successive
paracycle. (Photo no. 41848)

BUNYEROO-CAMBRIAN BOUNDARY
Exposed Bunyeroo Formation lies conformably above the Sandison Subgroup. The strata
are composed of laminated purple shales typical of the Flinders Ranges (Plummer, 1978). An
unusual feature within the map area, south of Willochra Creek (Fig. 2, *) is a thin 0.5-3 m
thick massive quartzite conformable with the top of the Bunyeroo Formation. This quartzite
has been mapped previously as an extremely thin unit faulted against the overlying Cambrian
strata (Webb and von der Borch, 1962; Binks, 1968). Close examination of the outcrop
suggests that the quartzite is conformable with the underlying Bunyeroo Formation and also
with the overlying Cambrian carbonates. No obvious structural evidence was found to
support fault movement. This quartzite unit is interpreted to be a stratigraphically thin
remnant of the Wilpena Group.

KEYWORDS: Stratigraphy/Stratigraphic definition/Adelaide Geosyncline/Sandison Subgroup/ Buckaringa


Gorge/Willochra Creek/ ‘Partacoona’/Flinders Ranges/SI 54-09 ORROROOI6533-IVI6534-III.

22
REFERENCES

Binks, P.J., 1968. ORROROO map sheet.


South Australia. Geological Survey/
Geological Atlas 1:250 000 Series, sheet
SI 54-1.
Dyson, I.A., 1992. Stratigraphic
nomenclature and sequence stratigraphy
of the lower Wilpena Group, Adelaide
geosyncline: the Sandison Subgroup.
South Australia. Geological Survey.
Quarterly Geological Notes, 122:2-13.
Plummer, P.S., 1978. Stratigraphy of the
lower Wilpena Group (Late
Precambrian), Flinders Ranges, South
Australia. Royal Society o f South
Australia. Transactions, 102( 1):25-28.
Plummer, P.S., 1989. Late Precambrian
wave-to tide-dominated delta evolution
in the west-central Adelaide Geosyncline,
South Australia. In: Jago, J.B. and
Moore. P.S. (Eds). The evolution of a
Late Precambrian-early Palaeozoic rift
complex: the Adelaide Geosyncline.
Geological Society o f Australia. Special
Publication, 16:164-176.
Rollings, N.M., 1986. Micaceous haematite
mineralisation and geology of the
Warrakimbo Gorge, southwest Flinders
Ranges, South Australia. South Australia.
Department o f Mines and Energy. Report
Book, 86/19.
Webb, B.P. and von der Borch, C.C., 1962.
Orroroo map sheet. South Australia.
Geological Survey. Geological Atlas
1:63 360 Series, sheet SI 54-1.

Subgroup, ‘Partacoona’ area: the section line is shown


on Figure 2.

23
191 Greenhill Road,
Parkside,
South Australia 5063

© Department of Mines and Energy, South Australia, 1993.

Prepared by the Publication Drafting Section


S.A. Departm ent o f Mines and Energy
All articles have been peer reviewed.
Printed by Kitchener Press Pty Ltd

View publication stats

You might also like