You are on page 1of 288

Erratum

There are several instances of a typographic error in


Figure 4.2 (page 79). The references to 'Y/Zr
!
in the
caption and 'Y/Zr' in labels on the diagram are
incorrect. They should be 'Zr/Y' in all cases.
Altered Volcanic Rocks
A guide to description and interpretation
Cathryn Gifkins
Walter Herrmann
Ross Large
Published by the Centre for Ore Deposit Research
University of Tasmania, Australia
UTAS
Published by CODES
Centre for Ore Deposit Research,
University of Tasmania,
Private Bag 79,
Hobart, Tasmania, Australia 7001
An ARC Special Research Centre
Centre for Ore Deposit Research, 2005
National Library of Australia Cataloguing-in-Publication Data
Gifkins, Cathryn.
Altered volcanic rocks : a guide to description and interpretation.
Bibliography.
Includes index.
ISBN 1 86295 219 1.
1. Rocks, Igneous. 2. Hydro thermal alteration. I.
Herrmann, Walter, 1951- . II. Large, Ross R. III.
University of Tasmania. Centre for Ore Deposit Research.
IV. Title.
552.2
another Pongratz Production 2005
Copy editing: Im'press: clear communication
Index: Word Wise and Im'press: clear communication
Printed in Australia by the Printing Authority of Tasmania
Ill
I CONTENTS
PREFACE v . v i i
ACKNOWLEDGEMENTS v i i i
INTRODUCTION ix
1 | ALTERATION IN SUBMARINE VOLCANIC SUCCESSIONS 1
1.1 Submarine volcanic successions 1
Volcanic facies 1
Volcanic facies associations 2
Evidence for submarine environment of emplacement 2
1.2 Alteration in submarine volcanic successions 2
Devitrification 4
Alteration processes 4
Characteristics inherited from volcanic facies 6
1.3 Geology of the Mount Read Volcanics 7
Stratigraphy of the Mount Read Volcanics 9
Submarine facies associations and architecture 10
Post-depositional alteration processes 11
Mineral deposits and prospects 11
1.4 Geology of the Mount Windsor Subprovince 12
Stratigraphy of the Seventy Mile Range Group 12
Submarine facies associations and architecture 13
Post-depositional alteration processes 14
Mineral deposits and prospects 14
2 | DESCRIBING ALTERED VOLCANIC ROCKS 15
2.1 Frequently asked questions 15
2.2 Alteration nomenclature 19
Mineral-based alteration nomenclature 19
Compositional alteration nomenclature 20
Generic alteration nomenclature 20
Descriptive nomenclature alteration facies 20
2.3 Alteraction facies the recommended method 22
2.4 Alteration mineral assemblage 23
Tools for mineralogical determination 24
2.5 Alteration intensity 25
Qualitative estimates of alteration intensity 25
Quantitative estimates of alteration intensity 26
An integrated approach to alteration intensity 33
2.6 Alteration data sheets 36
3 | COMMON ALTERATION TEXTURES AND ZONATION PATTERNS 37
3.1 Alteration textures 37
Replacement textures 37
iv | CONTENTS
Infill textures 41
Dissolution textures 41
Static recrystallisation textures 52
Dynamic recrystallisation textures 52
Deformation textures 52
3.2 Pseudotextures 54
Pseudoclastic textures 54
False polymictic texture 63
False matrix-supported texture 63
False coherent textures 63
3.3 Alteration distribution 63
3.4 Alteration zonation patterns 64
Regional diagenetic zones 64
Regional metamorphic zones 64
Regional, deep, semi-conformable altered zones 66
Local contact metamorphic or hydrothermally altered halos 66
Local hydrothermally altered halos around ore deposits 67
Vein and fracture altered halos 67
3.5 Overprinting relationships and timing of alteration 69
Method 70
Overprinting textures 70
4 | GEOCHEMISTRY OF ALTERED ROCKS 73
4.1 Lithogeochemistry 73
Sampling and analytical methods 73
Closure in composition data 78
Chemostratigraphy 79
- Mass transfer techniques 81
Rare-earth-element geochemistry related to alteration 87
4.2 Mineral chemistry 87
Principles 87
Applications 88
4.3 Stable isotopes 92
Theoretical background 92
Isotopic applications in alteration studies 92
5 | SEA F L OOR - A ND BURIAL-RELATED ALTERATION 97
5.1 Alteration related to sea-floor processes and burial 97
Physical conditions 98
Definitions 98
5.2 Hydration 98
Palagonite 99
Perlite 100
5.3 Diagenesis (glass to zeolite facies) 102
Diagenetic minerals 102
Diagenetic zones 105
Genesis of diagenetic minerals and zones 108
5.4 Regional metamorphism (zeolite to amphibolite facies) 115
Transition from diagenesis to regional metamorphism 115
Burial metamorphism 115
Burial metamorphic facies 115
Burial metamorphic zones 115
Zeolite facies 116
Genesis 116
5.5 Diagenetic alteration in the Hokuroku Basin 118
Geological setting 118
Alteration facies and zones 119
Genesis of altered zones 120
Data sheets 122
CONTENTS | V
5.6 Diagenetic alteration in the Mount Read Volcanics 128
Geological setting 128
Alteration fades and zonation 128
Genesis of alteration fades 128
Data sheets 133
6 | SYNVOLCANIC INTRUSION-RELATED ALTERATION 139
6.1 The role of intrusions in generating hydrothermal systems 140
Subseafloor regional hydrothermal systems 140
6.2 Regional altered zones assodated with intrusions 141
Recharge zones 141
Discharge zones 141
Deep, semi-conformable altered zones 142
Altered zones as part of a regional hydrothermal system 147
6.3 Altered zones within intrusions 148 '
Deuteric alteration 148
Hydrothermal alteration 148
6.4 Contact altered halos around intrusions 149
Contact altered zones 149
Genesis of contact altered zones 153
6.5 Contact altered zones associated with the Darwin Granite 154
Geological setting 155
Alteration fades and zonation 155
Genesis of the alteration system 156
Data sheets 157
7 | LOCAL HYDROTHERMAL ALTERATION RELATED TO
VHMS DEPOSITS 163
7.1 Common features of VHMS deposits 163
7.2 Hydrothermal alteration halos associated with VHMS deposits 164
Footwall alteration pipes 164
Stratabound altered footwall zones 166
Altered hanging wall zones 167
Chemical reactions and mass changes 168
Alteration box plot trends in altered footwall zones 169
The genesis of footwall alteration pipes 170
Significance of pyrophyllite and kaolinite in VHMS systems 174
Metamorphism of altered zones 174
7.3 The spectrum of volcanic-hosted deposits and associated alteration patterns 174
Hydrothermal alteration related to the spectrum of deposits 176
7.4 Comparisons between Archaean, Palaeozoic and Cainozoic VHMS
alteration systems 178
Australian Palaeozoic VHMS alteration halos 178
Japanese Cainozoic VHMS alteration halos 179
Canadian and Australian Archaean VHMS alteration halos 179
Comparisons : 180
7.5 Hellyer: a massive elongate polymetallic lens 181
Geological setting 182
Alteration fades and zonation 182
Ore genesis 183
Data sheets 184
7.6 Rosebery: a polymetallic sheet-style deposit 194
Geological setting 194
Alteration facies and zonation 195
Genesis of the ore lenses and alteration system 195
Data sheets 196
7.7 Western Tharsis: a hybrid Cu-Au VHMS deposit 202
Geological setting 202
Alteration facies and zonation 202
V i | CONTENTS
Ore genesis 203
Data sheets 204
7.8 Henty: a volcanogenic gold deposit 212
Geological setting 212
Alteration fades and zonation 212
Ore genesis 213
Data sheets 214
7.9 Thalanga: a polymetallic sheet-style deposit 221
Geological setting 221
Alteration facies and zonation 222
Ore genesis 222
Data sheets 223
7.10 Highway-Reward: a pipe style Cu-Au VHMS deposit 232
Geological setting 232
Alteration facies and zonation 232
Ore genesis 232
Data sheets 233
| FINDING ORE DEPOSITS IN ALTERED V OLCANIC ROCKS : 2 4 1
8.1 Principles of discriminating between diagenetic, hydro thermal
and metamorphic alteration facies 241
Diagenetic facies 241
Metamorphic facies 242
Hydrothermal alteration facies 242
8.2 Exploration vectors and proximity indicators 243
Mineral zonation 243
Major element lithogeochemistry 243
Alteration indices 244
Mass change vectors 245
Mineral chemistry vectors 245
Isotopic vectors 246
| REFERENCES 2 5 1
| INDEX 2 7 1
I vii
PREFACE
Altered volcanic rocks is principally for hands-on geologists,
our fortunate colleagues who practise in mineral exploration
and mining geology, and the students who may in the future
play in those professional fields.
We began designing and writing this book in mid 2001
after struggling for several decades to come to terms with a
variety of alteration styles in ancient submarine volcanic
successions. We realised that although a large number of
company and research geologists were working on similar
rocks there was no existing text to help guide us through the
complexity of altered volcanic rocks. The so-called volcanic
rocks we deal with in ancient volcanic successions and around
ore deposits frequently bear little resemblance to their fresh
counterparts, which are studied in undergraduate igneous
petrology and volcanology courses. It is typically only with
long experience that geologists develop the confidence and
skills to be comfortable working with altered volcanic rocks,
to interpret the original volcanic facies, unravel complex
alteration histories and determine their significance in terms
of mineral deposit prospectivity, particularly in ancient and
deformed successions.
The topic and content of the book were inspired by
problems that we have faced, and in many cases overcome,
while working on industry-related volcanic facies, alteration
geochemistry and economic geology research projects,
particularly in the Mount Read Volcanics. Many of the ideas
presented in this book come from the results of CODES
research projects, which have been run in collaboration
with industry partners and the Australian Research Council
(ARC) over the last 15 years. In particular, AMIRA-ARC
Linkage project P439 (Studies of VHMS-related alteration:
geochemical and mineralogical vectors to ore) provided an
enormous amount of data, case studies and expertise. Some
of the results of this project have previously been published as
a special issue in Economic Geology (Gemmell and Herrmann,
eds., A special issue devoted to alteration associated with volcanic-
hosted massive sulfide deposits, and its exploration significance,
August 2001, v. 96, no. 5).
We were encouraged by the wide acceptance and success
of the CODES publication by Jocelyn McPhie, Mark
Doyle and Rod Allen (1993) Volcanic textures: a guide to the
interpretation of textures in volcanic rocks, which has been a
major factor in improving the description and interpretation
of volcanic facies over the last decade. The advance we have
made in Altered volcanic rocks is to integrate observations
and data on volcanic facies and textures with alteration
mineralogy and geochemistry at both regional and local scales
in order to provide a multidisciplinary method for the study
and discrimination of different alteration types: diagenetic,
metamorphic and hydrothermal alteration.
We hope that this book will help to equip geologists
working in altered and deformed successions with the skill
and confidence to interpret the original volcanic facies and
encourage the use of altered rocks as discriminants and
vectors in mineral exploration. This book may not provide
all the answers, but if it gives readers the courage to tackle the
study of altered rocks, embrace the problems and pursue the
answers it will have been worthwhile.
Cathryn C. Gifkins
Wally Herrmann
Ross R. Large
viii |
| ACKNOWLEDGEMENTS
While preparing this book, we were fortunate to have valuable
support, assistance and advice from many people.
We extend our sincere thanks to those people whose
discussions and/or reviews of various chapters have helped
shape this book. Chapters were peer-reviewed by Stuart Bull,
David Cooke, Mark Doyle, Kim Denwer, Allan Galley, Bruce
Gemmell, Anthony Harris, Jocelyn McPhie, Andrew Rae,
Mike Solomon and Fernando Tornos. Valuable discussions
were also held with Ron Berry, Stuart Bull, Jocelyn McPhie,
Phil Robinson and Mike Solomon.
Although samples and photographs used herein are
principally from the authors' collections, we also made use of
hand specimens and thin sections from the School of Earth
Sciences rock catalogue at the University of Tasmania, and
samples and photographs from colleagues. Thank you to those
people who contributed: Sharon Allen, Stuart Bull, Kate Bull,
Tim Callaghan, Cari Deyell, Bruce Gemmell, George Hudak,
Karin Orth and Jocelyn McPhie. We also thank Izzy von
Lichtan, Curator at the School of Earth Sciences, for her help
in finding and returning hundreds of catalogue samples.
Andrew McNeill very kindly provided a long projection
of the Rosebery ore lenses. Tim Callaghan assisted with core
specimens and whole-rock geochemical data from Mount
Julia. Jon Huntington and Melissa Quigley at CSIRO
provided HyMap images of the Mount Lyell field.
We are infinitely grateful for the hard work of the
production team. Karin Orth and Simon Stephens helped with
sample preparation. Mike Blake and Karin Orth assisted with
photography. Rose Pongratz and Izzy von Lichtan prepared
the bibliography and checked references. June Pongratz
provided expert drafting, design and desktop publishing, and
was incredibly tolerant of the endless revisions. Final editing
was by Impress: clear communication and indexing by Word
Wise and Impress: clear communication.
We also appreciate our families, friends and colleagues
who have been very understanding of our commitment to
this project over the last three years. Thank you for your
support and patience.
r
I ix
INTRODUCTION
This book is about the processes and products of alteration
in submarine volcanic successions, although many of the
concepts presented here can be applied to altered volcanic
rocks from almost any environment. Its emphasis is on
hydrothermal alteration associated with volcanic hosted
massive sulfide (VHMS) deposits.
Few volcanic rocks in submarine settings are entirely
unaltered, and in hydrothermal environments all rocks
are altered to some degree. Recognising, describing and
interpreting altered volcanic rocks is not always easy, but
the results can have important implications for volcanology,
petrology and ore deposit studies, and can improve and
accelerate success in mineral exploration. Determining pre-
alteration characteristics and discriminating between primary
volcanic, magmatic and secondary alteration features requires
knowledge of the alteration processes and their products.
Valuable base-metal, gold and silver deposits exist
in a variety of modern and ancient submarine volcanic
successions. Many of these deposits are surrounded by, or
spatially related to, extensive altered zones that record the
passage of mineralising hydrothermal fluids and fluid-wall
rock reactions. Research into the textural, mineralogical and
compositional effects of alteration around VHMS deposits
has shown that they can be quantified and used as effective
exploration tools for discriminating deposit styles and guiding
exploration towards mineralised zones.
An introduction to alteration
Guilbert and Park (1986) defined alteration as any change in
the mineralogical composition of a rock brought about by
physical or chemical means, especially by the interaction with
hot or cold aqueous solutions or gases. Alteration typically
encompasses mineralogical changes and changes in the rock
texture and composition. Components of rocks, including
ore metals, can be dissolved, replaced or recrystallised. New
minerals may precipitate and isotopic ratios may change.
Porosity and permeability may be reduced or increased.
Primary volcanic textures are overprinted, and may be
destroyed and replaced by new 'false' textures, or enhanced.
The resulting altered rock is described as the 'alteration fades'
(e.g. Riverin and Hodgson, 1980).
Thus, alteration involves complex modification of a rock.
Furthermore, a rock may undergo several episodes of syn- to
post-depositional alteration, not all of which are related to
mineralising hydrothermal systems. Each alteration episode
is influenced by the existing texture and composition of the
rock, and may also overprint and modify that texture and
composition. As a result the characteristics of altered rocks
are highly variable. In ancient volcanic rocks it is a challenge
to determine host volcanic facies, unravel complex alteration
processes and interpret their significance in terms of mineral
prospectivity. That challenge is the focus of this book.
How the book is organised
Altered volcanic rockshzs two main themes, which are organised
into eight chapters: (1) it describes the basic principles behind
recognising and describing altered volcanic rocks; and (2) it
discusses the different alteration processes that are common
in submarine volcanic successions and their products.
Chapter 1 introduces the concepts of alteration in
submarine volcanic successions and summarises the main
alteration processes and volcanic facies. It outlines the regional
geology of two of the most productive Australian submarine
volcanic successions: the Cambrian Mount Read Volcanics
in western Tasmania and the Cambro-Ordovician Mount
Windsor Subprovince in Queensland. This book principally
employs examples from these two successions, and includes
descriptions of other ancient submarine volcanic successions
for comparison.
Chapter 2 discusses alteration nomenclature, mineralogy,
intensity and indices, and the principles of alteration facies.
It proposes an integrated multidisciplinary approach to
description and classification. The main elements of alteration
facies mineral assemblage, intensity, texture, distribution,
zonation and timing are described in Chapters 2 and 3.
Chapter 4 outlines geochemical methods used in alteration
studies and their applications. It emphasises whole-rock
lithogeochemistry, mineral chemistry and stable isotope
analysis.
Chapter 5 concentrates on regional alteration styles
including hydration, diagenesis and metamorphism associated
with burial, Chapter 6 on intrusion-related alteration styles,
X I INTRODUCTION
and Chapter 7 on hydrothermal alteration and mineralisation
associated with VHMS deposits. We present short case studies
for these different alteration styles, emphasising hydrothermal
alteration associated with a variety of VHMS deposits
including Rosebery, Hellyer, Henty and Thalanga. These
case studies incorporate pictorial data sheets that present the
mineralogical, textural and compositional characteristics of
each of the main alteration facies. They combine volcanic
facies analysis and alteration mineral assemblages, textures,
intensity and geochemistry to interpret the features of
different alteration styles.
Chapter 8 outlines the methods for discriminating
between the products of mineral deposit-related hydrothermal
alteration and other alteration processes, and identifying
favourable altered zones for mineral exploration. It also
discusses geochemical vectors that may guide explorers
towards mineralised rock within these altered zones.
Significance of altered volcanic rocks to mineral
exploration
Economic geologists are particularly interested in alteration
because hydrothermal mineral deposits are commonly hosted
by altered rock. Hydrothermally altered zones around mineral
deposits provide much larger targets for mineral exploration
than the deposits themselves. The mineral assemblages,
and in some cases the chemical composition, of the altered
rocks may provide indications of the proximity of an ore
deposit, and thus vectors towards mineralised rock. In
addition, mineralogical, textural and compositional studies
of alteration facies can provide important constraints on
the timing, physical and chemical conditions, and origins
of hydrothermal systems and related mineralisation (Barnes,
1979). The texture and distribution of alteration facies can
also be used to infer changes in porosity, permeability and
fluid pathways in the host succession. The results of alteration
studies are commonly incorporated into ore deposit models
used in mineral exploration. Thus, the identification and
interpretation of alteration facies is, and should be, a routine
part of exploration for hydrothermal mineral deposits.
.
11
1 | ALTERATION IN SUBMARINE VOLCANIC
SUCCESSIONS
This chapter describes submarine volcanic successions,
the common processes of alteration that occur in these
successions, and provides two examples of ancient submarine
volcanic successions, that have been variably altered and
mineralised: the Mount Read Volcanics and the Mount
Windsor Subprovince.
In submarine volcanic environments, the coincidence
of magmatic fluids, heat and abundant seawater generates
hydrothermal convection. As a consequence, submarine
volcanic successions may host important hydrothermal
mineral deposits, commonly referred to as volcanic-hosted
massive sulfide (VHMS) deposits.
VHMS deposits are a significant source of zinc, copper,
lead, silver and gold, and continue to be a target for base-metal
exploration. They range in size from less than one million
tonnes to over 200 million tonnes, and commonly contain
high metal grades. For example, the Hellyer deposit in western
Tasmania produced 16.2 Mt at 13.9% Zn, 7.1% Pb, 0.4%
Cu, 168 g/t Ag and 2.5 g/t Au in its nine years of operation.
VHMS deposits occur mainly in submarine rift environ-
ments particularly back arc and mid ocean rifts; however,
they can occur in a variety of other submarine environments
including continental rifts, oceanic basins or plateaux, and
arc-continent or continent-continent collision zones. They are
one of the few classes of ore deposits that exist throughout the
geological record from early Archaean to Recent.
1.1 | SUBMARINE VOLCANIC
SUCCESSIONS
Submarine volcanic successions are significantly different from
subaerial volcanic successions, as the processes of eruption,
transport, emplacement, and post-emplacement alteration
may be strongly affected by the presence of water. Typically,
submarine volcanic successions comprise a wide variety of
coherent and volcaniclastic facies intercalated with mixed
provenance and non-volcanic sedimentary facies (Fig. 1.1).
The volcanic facies may be derived from intrabasinal, extra-
basinal or basin-margin eruptions in submarine or subaerial
settings. Eruption styles may be effusive or explosive and the
products may remain in situ or be redeposited or reworked
by sedimentary processes. In addition, volcanic units may be
emplaced into the succession as synvolcanic intrusions.
This section summarises the main volcanic facies that
occur in submarine volcanic successions. For a more detailed
discussion of submarine volcanism, volcanic textures, facies
and their interpretation, readers are referred to McPhie et al.
(1993) Volcanic textures: a guide to the interpretation of textures
in volcanic rocks.
Volcanic facies
For descriptive purposes, volcanic facies are divided into two
main textural types: coherent and volcaniclastic. Coherent
facies consist of solidified magma and are commonly
characterised in volcanic rocks by aphyric (fine grained or
glassy) or porphyritic textures, where porphyritic refers to
evenly distributed euhedral crystals (phenocrysts) in a fine-
grained or glassy groundmass (McPhie et al., 1993).
Volcaniclastic facies are those composed mainly of volcanic
particles (Fisher, 1961). Volcanic particles are crystals, crystal
fragments, shards, pumice clasts, scoria clasts and dense
volcanic clasts, which may be produced by primary volcanic
(pyroclastic and autoclastic) or sedimentary (weathering and
erosion) processes. Volcaniclastic facies include a spectrum
of facies: primary volcanic facies, syneruptive volcanic
facies generated by coeval eruptions and deposited from
sedimentary processes, and volcanogenic sedimentary facies
that show evidence of temporary storage and reworking prior
to deposition (McPhie et al., 1993).
Primary volcaniclastic facies result from volcanic processes
of clast formation, transport and deposition and include
hydroclastic, pyroclastic and autoclastic facies. Hydroclastic
facies is a general term for facies, typically comprising blocky
glassy particles, produced by magma-water interactions,
whether by explosive steam generation or by non-explosive
quench fragmentation of magma (Fisher and Schmincke,
1984; Hanson, 1991). Pyroclastic facies comprise volcanic
particles (pyroclasts) that were generated by explosive
eruptions and deposited by primary volcanic processes, by
fallout, flow or surge. Autoclastic facies comprise volcanic
particles generated by in situ non-explosive fragmentation of
lava or magma (autobrecciation and quench fragmentation).
2 | CHAPTER 1
Autobrecciation occurs when the more viscous parts of a
moving lava respond in a brittle fashion to locally higher
strain rates, and fragment into blocky clasts (Fisher, I960).
Quench fragmentation occurs in situ where hot lava or
magma comes into contact with water, ice or water-saturated
sediment (Rittmann, 1962; Pichler, 1965; Yamagishi, 1987).
The resulting autoclastic deposits - autobreccia, hyaloclastite
or peperite - typically comprise dense blocky or splintery
clasts, but they may be pumiceous and have fluidal shapes
(Fisher, 1960; Pichler, 1965; Busby-Spera and White, 1987;
Gifkins et al., 2002).
Syneruptive volcaniclastic fades are composed dominantly
of unmodified volcanic clasts that were fragmented by
volcanic process such as explosive eruptions, autobrecciation
or hydration, but were transported and deposited by
sedimentary processes (McPhie et al., 1993; McPhie and
Allen, 2003). They can occur directly from eruption when
clasts bypass initial deposition as primary deposits and are
delivered directly to sedimentary transport and deposition
systems, such as subaqueous eruption-fed water-supported
gravity currents or water-settled fall (e.g. White, 2000;
McPhie and Allen, 2003). They may also occur indirectly
by rapid remobilisation and redeposition during or shortly
after the eruption (Fisher and Schmincke, 1984; Cas and
Wright, 1987; McPhie et al., 1993). Unconsolidated volcanic
debris may be remobilised by: the slumping and sliding of
gravitationally unstable rapidly accumulated clastic debris;
explosive eruptions; local uplift; syn-depositional faulting;
and extrusion and intrusion of magma.
Volcanogenic sedimentary fades (epiclastic volcanic, Fisher,
1960) contain volcanic particles derived from the post-
eruptive erosion and reworking of pre-existing volcanic facies
and may include a significant proportion of non-volcanic
particles (McPhie et al., 1993).
In submarine volcanic successions, volcaniclastic facies are
dominated by in situ autoclastic and syneruptive volcaniclastic
facies where the particles were derived from either autoclastic
fragmentation or explosive eruption. Most volcanic and non-
volcanic clastic deposits were emplaced by water-supported
density currents (i.e. high- and low-concentration turbidity
currents, debris flows and grain flows) and as fallout from
suspension in the water column.
Evidence for submarine environment of
emplacement
VHMS deposits occur in submarine volcanic successions,
thus exploration for new deposits is restricted to submarine
successions. However, there are few volcanic or sedimentary
facies that unequivocally constrain the host depositional
environment.
A subaqueous setting (marine or lacustrine) may be
interpreted based on the presence of: water-supported mass-
flow deposited facies; hemi-pelagic, biogenic, biochemical
and chemical sedimentary facies; pillow lavas; and quench
fragmented volcaniclastic facies. Also seawater-related dia-
genetic alteration facies (e.g. widespread albite alteration
facies) can suggest a submarine environment.
Without fossil evidence, differentiating between marine
and lacustrine settings is difficult as few facies are restricted
to either environment. Facies with tidal and wave tractional
structures, such as bimodal-bipolar ripples, are submarine,
whereas lacustrine settings may be indicated by the presence
of evaporites. Hummocky cross stratification is more common
in, although not restricted to, marine settings. Carbon-oxygen
isotope signatures of carbonates can be used to support marine
or lacustrine environments.
Although bedforms, sedimentary structures and some
sedimentary deposits help us to interpret a marine environ-
ment, they are of little help in constraining the water depth.
Water depth may be an important consideration for mineral
exploration as recent research suggests that Au-rich VHMS
deposits are restricted to shallow water environments (e.g.
Hannington et al., 1999; Hannington and Herzig, 2000;
Herzig et al., 2000). Shallow water environments are typically
dominated by the tractional processes of tidal and wave
action and result in characteristic sedimentary structures
and bedforms. In contrast, deep water environments, below
storm wave base, generally lack tractional currents: sediment
distribution and deposition mainly occurs through the
actions of turbidity currents, debris flows and the process
of suspension sedimentation. Water depth and depositional
setting may be more accurately constrained by the presence
of fossiliferous limestone or sedimentary facies that contain
marine fossils intercalated in the volcanic succession.
Volcanic facies associations
A facies association is a collection of facies that are spatially,
mineralogically, compositionally or texturally related, and
that may also be genetically related (Cas and Wright, 1987).
There are three common types of volcanic units represented
by facies associations in submarine volcanic successions
(Fig. 1.1): lavas, synvolcanic intrusions (cryptodomes, sills
and dykes) and syneruptive volcaniclastic facies. Lavas and
synvolcanic intrusions comprise associations of coherent and
autoclastic facies. The syneruptive volcaniclastic facies can be
divided into two principal categories, those dominated by
non- to poorly-vesicular blocky lava clasts and related to the
submarine emplacement of lavas and lava domes, and others
that contain abundant pumice or scoria clasts produced by
explosive eruptions. In addition, there is a wide variety of
volcanogenic sedimentary facies.
1.2 | ALTERATION IN SUBMARINE
VOLCANIC SUCCESSIONS
After emplacement, volcanic facies are commonly subjected
to a variety of alteration processes (Fig. 1.2). Alteration occurs
when existing components become unstable under changing
physical and chemical conditions, and alter to more stable
minerals. Volcanic glass, which is the main component of
many volcanic facies, is a metastable solid with the structure
of a liquid (Carmichael, 1979). It is undercooled to the point
where extreme viscosity has prevented crystallisation. As a
result, volcanic glass readily devitrifies to minerals that are
more stable under surface conditions; generally clay minerals,
zeolites, carbonates, feldspar, quartz and oxides (Carmichael,
1979; Henley and Ellis, 1983; Fisher and Schmincke, 1984;
ALTERATION IN SUBMARINE VOLCANIC SUCCESSIONS | 3
FIGURE 1.1 | Facies model of a submarine basin in which a variety of coherent and clastic volcanic facies are intercalated with sedimentary facies. The volcanic
facies include primary coherent and autoclastic facies, syneruptive and post-eruptive volcaniclastic facies derived from submarine and subaerial eruptions. Many of
the volcanic facies associations are laterally discontinuous. Common facies associations represent (A) lavas and lava domes composed of coherent and autoclastic
facies; (B) synvolcanic sills and cryptodomes; (C) syneruptive volcaniclastic facies derived from explosive and effusive submarine eruptions; (D) volcanogenic
sedimentary or resedimented volcaniclastic facies derived from pre-existing deposits; (E) syneruptive volcaniclastic facies derived from subaerial explosive eruptions;
(F) mixed provenance sedimentary facies; and (G) marine sedimentary facies.
FIGURE 1.2 | Facies model showing the distribution of different styles of altered zones in a submarine volcanic succession that hosts VHMS deposits. See Figure
1.1 legend for the patterns denoting volcanic and sedimentary facies.
4 | CHAPTER 1
Friedman and Long, 1984; Cerling et al., 1985). Alteration
of volcanic glass involves not only devitrification, but changes
in texture, composition, porosity and permeability, and may
simultaneously affect both the chemistry and circulation of
pore fluid in the volcanic succession (Noble, 1967; Dimroth
and Lichtblau, 1979; Fisher and Schmincke, 1984; Noh and
Boles, 1989; Torres et al., 1995).
Understanding alteration requires a range of skills that
include recognising alteration minerals, textures, paragenesis,
distribution, zonation, intensity, mineralogical and chemical
changes associated with alteration, pathways and mechanisms
for fluid migration, and fluid origin. These characteristics are
related to the alteration processes and to the characteristics of
the host volcanic succession.
Devitrification
The cooling history of volcanic facies may involve primary
crystallisation and later devitrification. Primary or high-
temperature crystallisation refers to crystallisation of magma
resulting in phenocrysts, microcrysts and microlites. In
contrast, devitrification refers to crystallisation of glass at low
temperatures (i.e. below the glass transition temperature:
Lofgren, 1971a). High-temperature devitrification accom-
panying first cooling may produce spherulites, lithophysae
and micropoikilitic or snowflake texture (e.g. Lipman, 1965;
Anderson, 1969; Lofgren, 1971b; Bigger and Hanson,
1992; McArthur et al., 1998) and is not considered to be
alteration. Low-temperature devitrification results in the
gradual conversion of glass to fine-grained granular crystalline
aggregates, which may happen over time as a result of
alteration during changing physical conditions or in response
to interaction with fluid. Devitrification may be accompanied
by changes in whole-rock composition (Lipman, 1965;
Lofgren, 1971a; Friedman and Long, 1984).
the presence of fluid (seawater, magmatic fluid or a mixture of
both). There are gradations from isochemical metamorphism
to metasomatism with increasing compositional changes.
The different alteration processes, hydration, diagenesis,
metamorphism and local hydrothermal alteration, are all part
of this continuum in submarine volcanic successions (Fig.
1.3).
The effects of each alteration process may be difficult
to distinguish. Hydrothermal alteration, diagenesis and
metamorphism can result in similar mineral assemblages
and textures. In addition, in many cases, different alteration
processes, such as diagenetic and hydrothermal alteration,
are contemporaneous and their products may be inseparable
(Iijima, 1974, 1978; Ohmoto, 1978; Reyes, 1990; Utada,
1991;Paradisetal., 1993).
In Chapters 5, 6, and 7 of this book, the common
alteration processes are grouped into those related to burial,
intrusions and VHMS deposits (Fig. 1.4). Thus, burial-
related alteration styles include hydration, diagenesis and
burial metamorphism. Alteration styles associated with
intrusions are hydrothermal alteration within intrusions,
contact metamorphism and hydrothermal alteration, and
regional hydrothermal alteration. Included below is a brief
introduction to each of the common alteration processes that
operate in submarine volcanic settings.
Hydration of volcanic glass
Hydration of glass involves the absorption of external
water into glass and modification of the glass structure,
either during cooling or at ambient temperatures (Ross and
Smith, 1955; Friedman and Long, 1984). Hydration does
not directly produce new minerals, but can form perlitic
fractures or palagonite in basaltic glass and it can facilitate
subsequent alteration (see Chapter 5). Compositional changes
Alteration processes
Alteration may result from regional or local processes. It
can occur as a result of the interaction with hydrothermal
fluid, as a result of changing physical (mainly temperature
and pressure) conditions during burial, in association with
the emplacement of intrusions, or a combination of all these
processes. Submarine volcanic facies, especially glassy facies,
are readily altered during hydration, diagenesis, hydrothermal
alteration, metamorphism and tectonic deformation.
Hydrothermal alteration is defined as the alteration of
rocks or minerals by the reaction of hydrothermal fluid
with pre-existing solid phases (Henley and Ellis, 1983).
Hydrothermal fluid is a hot aqueous solution or gas, with or
without demonstrable association with igneous processes.
Hydrothermal alteration usually results in significant changes
in rock texture, mineralogy and composition.
Alteration is either metasomatic or isochemical. Meta-
somatism involves changes in mineralogy, texture and
composition, whereas isochemical alteration (or meta-
morphism) involves mineralogical and textural changes
only. In submarine volcanic successions, almost all alteration
involves some degree of metasomatism, which is facilitated by
FIGURE 1.3 | This cartoon depicts the continuum between isochemical
and hydrothermal alteration and shows the alteration processes common in
submarine volcanic successions. They are positioned based on the relative
degrees of chemical exchange for each process.
accompanying hydration include gains in H
2
O, and minor
losses in silica and alkalis (Noble, 1967; Friedman and Long,
1984; Mungall and Martin, 1994).
Diagenesis
Diagenesis encompasses the changes that occur in response
to changing temperature and pressure during burial.
During diagenesis of volcanic facies, significant textural and
mineralogical changes can be produced by precipitation of
cement, dissolution and replacement of original components,
especially glass, and compaction (Fisher and Schmincke,
1984; Marsaglia and Tazaki, 1992; Torres et al., 1995). In
theory, diagenesis in submarine volcanic successions is a
metasomatic process involving minor chemical exchange
between the host facies and trapped modified seawater at low
temperatures (<250C). Transitions between diagenesis and
metamorphism, and diagenesis and hydrothermal alteration
have not been rigorously defined and are discussed in Sections
5.4 and 6.2.
Regional metamorphism
Regional metamorphism involves pervasive, mainly iso-
chemical, mineralogical and textural changes in response to
increasing pressure and temperature (Yardley, 1989). During
metamorphism, H
2
O and CO
2
-bearing fluids are generated
by dehydration and decarbonation reactions (Rose and Burt,
1979).
Contact alteration associated with intrusions
Contact alteration refers to the changes caused by the rise in
temperature of the host rock immediately surrounding an
intrusion, which may be accompanied by the circulation of
heated pore fluids around and within the intrusion. Contact
alteration can be isochemical (i.e. contact metamorphism)
or metasomatic (i.e. hydrothermal alteration). Contact
metamorphism typically results in recrystallisation of existing
minerals or components and minor remobilisation of elements
(Yardley, 1989). The effects of hydrothermal alteration
may include major changes in texture, mineral assemblage
and whole-rock composition on a scale of centimetres to
kilometres.
Hydrothermal alteration related to VHMS deposits
Two styles of hydrothermal alteration are commonly related
to VHMS mineralisation: (1) local alteration halos around
ore deposits; and (2) regional hydrothermally altered zones.
Regional hydrothermally altered zones are commonly spatially
and genetically associated with large intrusions and hence
in this book are discussed in Chapter 6 - intrusion-related
alteration styles.
Local hydrothermally altered halos around VHMS
deposits result from the reaction between the host facies and
the mineralising hydrothermal fluid (Sangster, 1972; Franklin
et al., 1975; Riverin and Hodgson, 1980; Green et al., 1983;
Urabe.et al., 1983). These altered halos are commonly zoned,
reflecting changes in the composition, pH and temperature
of the hydrothermal fluid with time, or the extent of reaction
with the host facies (Rose and Burt, 1979; Lydon and Galley,
1986; Schardt et al., 2001).
The nature of altered halos around VHMS deposits depends
on the host volcanic facies, host-rock composition, timing of
the hydrothermal alteration relative to the emplacement or
deposition of facies, structures, fluid pathways, distribution
pattern of the ore, and chemical and physical conditions of the
hydrothermal fluid (Large, 1992). Thus altered halos around
VHMS deposits exhibit a wide variety of geometries, sizes,
FIGURE 1.4 | Common alteration processes and their products.
ALTERATION IN SUBMARINE V OLCANIC SUCCESSIONS | 5
I CHAPTER 1 - "
alteration mineral assemblages, intensities, compositions, and
zones (Chapter 7).
Thick, extensive (up to 20 km), pervasive sub-horizontal
semi-conformable altered zones, referred to as regional
hydrothermal alteration zones or deep semi-conformable
alteration zones (Section 6.2), have been recognised in some
volcanic successions hosting VHMS deposits (Gibson et al.,
1983; Galley, 1993). Many of these regional hydrothermally
altered zones are spatially associated with, and possibly
genetically related to, intrusions (Galley, 1993; Brauhart et
al., 1998). They are interpreted to result from large-scale
convection of seawater through permeable volcanic successions
at elevated geothermal gradients (Spooner and Fyfe, 1973;
Munha and Kerrich, 1980; Baker, 1985). Reactions between
the volcanic successions and modified seawater have involved
Na-, Si-, Ca-Fe-, K-, or Mg-metasomatism, and the leaching
of ore-forming metals (Munha and Kerrich, 1980; Gibson et
al., 1983; Baker, 1985; Galley, 1993).
Syntectonic hydrothermal alteration
Hydrothermal alteration can also be synchronous with
tectonic deformation: syntectonic hydrothermal alteration. In
this case, the hydrothermal fluid may be modified seawater,
magmatic water or volatiles released during metamorphism,
or a combination of these, and migrates principally along
faults and shear zones. Contemporaneous deformation
may modify or destroy pre-existing textures and create new
textures or foliations.
Many VHMS deposits, such as Rosebery, Hercules and
Mount Lyell (western Tasmania), have been affected by
later tectonic deformation and modified by syntectonic
hydrothermal fluids (Walshe and Solomon, 1981; Khin Zaw
and Large, 1992). Although syntectonic hydrothermal fluids
were not responsible for the formation of these VHMS ores,
they can be critical to the subsequent formation of other styles
of ore deposits in the submarine volcanic successions, such as
mesothermal gold deposits. Detailed discussion of syntectonic
hydrothermal alteration is not dealt with in this book.
Characteristics inherited from volcanic facies
Volcanic components and facies with different compositions
and textures behave differently during the initial stages of
low-temperature alteration. Some components react more
rapidly than others and their composition may influence
the composition of the alteration mineral assemblage. For
example, Marsaglia and Tazaki (1992), in their study of
diagenetic trends in volcaniclastic sandstones of the Izu-
Bonin Arc, found that black mafic crystalline fragments were
unaltered, brown intermediate to mafic glassy fragments
showed evidence of dissolution, and colourless rhyolitic
fragments were altered to clay minerals. These differences
were due to the variable reactivity of the components and the
proportions of volcanic glass to crystalline facies.
Generally glass is the most reactive component, followed
by olivine > magnetite, titanomagnetite and ilmenite *
pyroxene and amphibole > biotite * Ca-plagioclase *
microcline, sanidine and orthoclase * quartz, apatite, rutile
and zircon (Browne, 1978; Reyes, 1990). Alteration rates
for different minerals vary considerably because of mineral
structure and composition. Silicate materials with an extensive
cross-linked (e.g. tetrosilicate) structure react slowly; whereas
those silicates with poorly connected fabric tend to react
rapidly and uniformly (Casey and Bunker, 1990).
Felsic volcanic facies typically have a higher proportion of
glassy to crystalline facies than mafic facies. This is because
the viscosity of silicic magmas (71-77% SiO
2
) inhibits
diffusive crystal growth and thus produces thick bodies of
glass, whereas the low viscosity of basaltic magmas favours
crystallisation (Friedman and Long, 1984).
Glassy facies or facies that contain glassy fragments are
likely to be more rapidly altered, and to form different mineral
assemblages, than those that are crystalline (Lee and Klein,
1986). In addition, mafic glasses are more rapidly altered than
felsic glasses (e.g. Whetten and Hawkins, 1970; Fisher and
Schmincke, 1984; Friedman and Long, 1984). The rate of
alteration is related to the glass's viscosity, which in turn is
a function of the composition (particularly SiO
2
and H
2
O)
and temperature (Friedman and Long, 1984). Increased SiO
2
decreases the rate of alteration, whereas increased MgO, CaO
and H
2
O increases the rate. Thus the higher SiO
2
content of
rhyolitic glasses retards reaction (Hawkins, 1981).
The primary composition can influence the alteration
mineralogy mainly because the dissolution of glass liberates
alkalis and silica, which are consumed by subsequent
reactions. Hence, highly silicic volcanic facies result in the
crystallisation of Si- and Na-rich minerals, such as opal,
quartz, tridymite, cristobalite and Na-zeolites (Sheppard et
al., 1988). In contrast, mafic glasses, such as those on pillow
rims, alter to Ca-, Fe-, Mg- and Mn-rich minerals such as
smectites, phillipsite, oxides and chlorite.
Volcaniclastic facies, particularly pumice-rich facies,
initially have very high porosities. In volcaniclastic facies, the
inter- and intra-particle pore space controls the porosity and
permeability and thus grain size, type and sorting influence
the distribution of early alteration facies. Early diagenetic
alteration in well-sorted pumice breccias, although commonly
patchy, is pervasive. In poorly sorted polymictic volcaniclastic
facies the porosity and permeability are initially much
more variable and diagenetic facies typically have complex
distribution patterns.
Coherent facies have much lower porosity and permea-
bility, factors that are controlled by fractures produced by
quenching, flowage and hydration. Alteration in coherent
facies typically progresses as fronts that move outward from
fractures into the less altered domains (e.g. the alteration of
perlite, Noh and Boles, 1989).
In addition, patchy or domainal alteration styles in
volcaniclastic facies may also be related to variations in the
quenching and hydration of glassy clasts (Surdam, 1973;
Boles and Coombs, 1977; Marsaglia and Tazaki, 1992).
Hydrothermal experiments on rhyolitic glass indicate that
at high temperatures (>200C), rhyolitic glass does not
recrystallise but instead acts as Na-K ion exchanger. Quenched
glass fixes K
+
, whereas slowly cooled glass fixes Na
+
. Therefore
variations in cooling history may explain why some glassy
fragments alter more readily to particular minerals than others
(Marsaglia and Tazaki, 1992).
6 | CHAPTER 1
1.3 | G E O L O G Y OF THE M O U N T R E A D
VOLCANICS
Many of the examples of altered volcanic rocks and alteration
systems discussed in the following chapters come from the
Middle to Late Cambrian Mount Read Volcanics in western
Tasmania. The Mount Read Volcanics are a submarine
succession of rhyolitic to basaltic volcanic and intrusive
rocks with variable proportions of intercalated sedimentary
rocks. They are interpreted as the products of post-collisional
volcanism associated with arc-continent collision (Berry and
Crawford, 1988; Crawford and Berry, 1992). The succession
occurs in a 200 x 20 km area that extends from Elliott Bay in
ALTERATION IN SUBMARINE V OLCANIC SUCCESSIONS | 7
the south through Queenstown and Rosebery to Deloraine in
the north (Fig. 1.5). The volcanic succession was deposited
in a series of troughs separated by areas of Proterozoic
basement (Corbett and Lees, 1987; Corbett, 1992; Crawford
and Berry, 1992). The Mount Read Volcanics host gold,
silver and base-metal massive sulfide (VHMS) ore deposits
at Hellyer, Que River, Rosebery, Hercules, Henty and Mount
Lyell (Fig. 1.5). The mineral district is referred to as the
Mount Read province.
The Mount Read Volcanics have undergone diagenetic
and hydrothermal alteration, metamorphism, at least two
phases of deformation, and intrusion by Cambrian and
Devonian granites (Corbett and Lees, 1987; Corbett, 1992).
FIGURE 1.5 | Distribution of the principal lithostratigraphic units and major massive sulfide deposits in the central Cambrian Mount Read Volcanics, western
Tasmania. Modified after Corbett (1992,2002).
8 I CHAPTER 1
FIGURE 1.6 | Stratigraphic correlation diagram showing the major lithostratigraphic units in the Mount Read Volcanics to the west (A) and east (B) of the Henty
fault. The sections are located at (1) Hellyer-Que River, (2) Pinnacles, (3) Hollway, (4) Mount Black, (5) Rosebery-Hercules, (6 ) White Spur, (7) Hall Rivulet Canal,
(8) Murchison Gorge, (9) Henty, (10) South Henty, (11) Anthony Road, (12) Comstock-Lyell, (13) Lynchford, (14) Jukes-Darwin. Sections are modified after Fitzgerald
(1974), Corbett (1979,1992, 2001), Cox (1981), Komyshan (1986 ), Coutts (1990), Allen (1991), Dugdale (1992), Waters and Wallace (1992), Jones (1993,1999),
McKibben (1993), Herrmann and MacDonald (1996 ), McPhie (1996 ), White and McPhie (1996 ,1997), Callaghan (2001), Gifkins (2001) and Wyman (2001).
(A) To the west of the Henty fault, the Central Volcanic Complex interfingers with, and is conformably and disconformably overlain by, the Dundas and Mount Charter
Groups of the western volcano-sedimentary sequences (Corbett and Lees, 1987; Corbett, 1992). Immediately overlying the Central Volcanic Complex is a variety
of small-volume sedimentary (Black Harry beds and Animal Creek greywacke) and volcanic units (rhyolite and pumice breccia). These units are overlain by the
Que-Hellyer Volcanics, a succession of calc-alkaline to shoshonitic, intermediate to mafic lavas and volcaniclastic units (Corbett and Komyshan, 1989; Waters and
Wallace, 1992). The Que-Hellyer Volcanics host the Que River and Hellyer ore deposits, and extend via Sock Creek to Burns Peak and Pinnacles (i.e. the Brown's
tunnel sequence). The Que River Shale overlies the Que-Hellyer Volcanics and is similar to mudstone in other lithostratigraphic units of the Mount Read Volcanics.
The Southwell Subgroup overlies the Que River Shale and is lithologically similar to the White Spur Formation and the Rosebery hanging-wall volcaniclastic units,
comprising quartz-bearing volcaniclastic mass-flow units interbedded with black mudstone and Precambrian basement-derived turbidites (Corbett, 1992; McPhie and
Allen, 1992). Overlying the Southwell Subgroup is the Mount Charter Group in which the upper Mount Cripps Subgroup is a correlate of the Tyndall Group (Corbett,
1992).
Although the primary textures, mineralogies and whole-rock
compositions have been modified to various degrees, volcanic
textures are generally well preserved. Locally, two regional
tectonic cleavages have been recognised; however, the axial
planar S
2
Devonian cleavage is the dominant cleavage. S
2
strikes north, dips steeply and varies from a weak, spaced
cleavage to an intense, pervasive, anastomosing cleavage in
the most strongly deformed rocks adjacent to faults and in
phyllosilicate-rich altered zones.
The geology of the Mount Read Volcanics has been
described in detail by Campana and King (1963), Corbett
(1981, 1986, 1992, 1994), Corbett and Lees (1987), Corbett
and Solomon (1989), Pemberton and Corbett (1992), McPhie
and Allen (1992) and Crawford et al. (1992).
ALTERATION IN SUBMARINE V OLCANIC SUCCESSIONS | 9
B
(B) To the east of the Henty fault, the Eastern quartz-phyric sequence overlies the Sticht Range Beds, interfingers with the Central Volcanic Complex and is
conformably overlain by the western volcano-sedimentary sequences (Farrell Slates). The southern Central Volcanic Complex is flanked to the west by the Yolande
River Sequence, part of the western volcano-sedimentary sequences. To the east, it interfingers with the Eastern quartz-phyric sequence and is overlain by the
Tyndall Group (Corbett, 1992) and locally by andesite and basalt lenses that occur between Henty and Queenstown (Anthony Road andesite, Crown Hill andesite,
Howards basalt, Spillway basalt, and Lynchford basalt). The Cambrian Murchison and Darwin granites intruded the succession in the Mount Murchison and Mount
Darwin areas (Corbett and Lees, 1987; Corbett, 1992). The Tyndall Group is the youngest lithostratigraphic unit. It extends north-south along the eastern margin of the
succession where it unconformably overlies Tyennan basement, Sticht Range Beds, southern Central Volcanic Complex, Eastern quartz-phyric sequence, western
volcano-sedimentary sequences, and the Darwin Granite (Corbett and Lees, 1987; White and McPhie, 1997). The Owen Conglomerate overlies the Mount Read
Volcanics both conformably and unconformably.
Stratigraphy of the Mount Read Volcanics
The stratigraphy of the Mount Read Volcanics can be divided
into (Figs 1.5 and 1.6): Sticht Range Beds, Eastern quartz-
phyric sequence, Central Volcanic Complex, western volcano-
sedimentary sequences, and the Tyndall Group and correlates
(Corbett, 1992). These lithostratigraphic units comprise
compositionally and texturally diverse coherent volcanic and
volcaniclastic facies intercalated with sedimentary rocks, which
are distinguished and mapped on the basis of the dominant
facies. The principal volcanic facies associations are lavas,
synvolcanic intrusions and syneruptive volcaniclastic units
(McPhie and Allen, 1992). Lavas and synvolcanic intrusions
are predominantly calc-alkaline rhyolites and dacites with
locally abundant andesites and basalts (Crawford et al.,
1992). The volcaniclastic facies associations include a variety
of primary and secondary volcaniclastic facies including
thick extensive syneruptive pumice- and crystal-rich units
and in situ and resedimented hyaloclastite (McPhie and
Allen, 1992). Sedimentary facies include black mudstone,
and graded, bedded sandstone of mixed volcanic and meta-
sedimentary Precambrian basement provenance (McPhie and
Allen, 1992).
Regional stratigraphic relationships between the litho-
stratigraphic units are complex and laterally variable
(Fig. 1.6). The Mount Read Volcanics are conformably and
10 | CHAPTER 1
unconformably overlain by the Owen Group, a thick sequence
of Late Cambrian-Early Ordovician siliciclastic, shallow
marine to fluvial conglomerate and sandstone (Corbett,
1992; White, 1996).
VHMS deposits occur in a variety of volcanic facies in
three of the main lithostratigraphic subdivisions of the
Mount Read Volcanics (McPhie and Allen, 1992; Pemberton
and Corbett, 1992; Waters and Wallace, 1992). In particular,
VHMS deposits are interpreted to occur: (1) at the top of the
Central Volcanic Complex, close to large felsic volcanic centres
(Rosebery, Hercules and Mount Lyell); (2) associated with
proximal facies of andesite-dacite volcanoes in the western
volcano-sedimentary sequences (Hellyer and Que River); and
(3) in the base of the Tyndall Group (Henty and Comstock)
(Corbett and Solomon, 1989; Halley and Roberts, 1997).
Sticht Range Beds
The Sticht Range Beds are a thin (<500 m) basal succession
of interbedded basement-derived sedimentary rocks and
volcaniclastic rocks that occur along the eastern margin of the
Mount Read Volcanics (Fig. 1.5: Baillie, 1989).
Eastern quartz-phyric sequence
The Eastern quartz-phyric sequence is a 2.5 km-thick
succession of quartz + feldspar-phyric lavas, synvolcanic
intrusions and volcaniclastic units limited to the eastern
margin of the Mount Read Volcanics (Fig. 1.5: Polya, 1981;
Polya et al., 1986; Pemberton et al., 1991; Corbett, 1992).
Central Volcanic Complex
The 3 km-thick Central Volcanic Complex dominates the
central part of the Mount Read Volcanics between Mount
Darwin and Mount Block (Fig. 1.5: Corbett, 1979). It consists
of feldspar-phyric rhyolitic and dacitic lavas, synvolcanic
intrusions and pumiceous volcaniclastic units (Corbett, 1979,
1992; Corbett and Lees, 1987; Corbett and Solomon, 1989;
Pemberton and Corbett, 1992; Gifkins, 2001). Andesites
and basalts are locally intercalated with the felsic succession
(Crawford et al., 1992). Quartz + feldspar-phyric intrusions
and tholeiitic basalt and dolerite dykes (Henty dyke swarm)
occur throughout the northern Central Volcanic Complex
(Corbett and Solomon, 1989; Crawford et al., 1992).
Western volcano-sedimentary sequences
The western volcano-sedimentary sequences include the
Yolande River Sequence, Dundas Group, Mount Charter
Group and Henty fault wedge sequence (Corbett, 1992).
These sequences are thick (>3 km) mainly sedimentary
successions of quartz + feldspar-phyric volcaniclastic facies,
mixed provenance sandstone and mudstone intercalated
with rhyolitic, andesitic and basaltic lavas and synvolcanic
intrusions, (Corbett and Lees, 1987; Corbett, 1989; McPhie
and Allen, 1992). The volcaniclastic facies contain a diverse
range of clasts including quartz + feldspar porphyry, feldspar-
phyric rhyolite, pumice, granite and massive sulfide clasts.
Tyndall Group and correlates
The Tyndall Group is the youngest lithostratigraphic unit
in the Mount Read Volcanics. It extends along the eastern
margin and locally along the western side of the succession
(Fig. 1.5). The Tyndall Group varies in thickness from 350 to
1300 m and comprises distinctive quartz + feldspar crystal-
rich sandstone, volcanic breccia and volcanic conglomerate
intercalated with minor rhyolitic welded ignimbrite, felsic
to intermediate lavas and intrusions, and non-volcanic
sedimentary rocks including limestone, mudstone and
sandstone (White and McPhie, 1996, 1997).
Cambrian granites
Five Cambrian granitoids (commonly referred to as 'granites')
have been recognised in western Tasmania: the Murchison,
Darwin, Elliott Bay, Dove and Timber Tops granites (Leaman
and Richardson, 1989). Cambrian granites may also occur
at depth in a belt that extends along the eastern margin of
the Mount Read Volcanics between Mount Darwin and
Mount Murchison (Large et al., 1996). They are typically
medium grained quartz + K-feldspar + plagioclase + biotite +
hornblende + apatite + zircon rutile granite or granodiorite
(McNeill and Corbett, 1992). They intrude the western
volcano-sedimentary sequences, Central Volcanics Complex
and Eastern quartz-phyric sequence and are unconformably
overlain by the Tyndall Group in the Murchison and Darwin
areas (Corbett, 1992). They are interpreted to be subvolcanic
intrusions genetically related to the host volcanic succession
(Solomon, 1981).
Submarine facies associations and architecture
The Mount Read Volcanics were deposited in a predominantly
below wave-base, moderate to relatively deep submarine
setting. This interpretation is supported by the presence of
trilobite and other marine fossils, fossiliferous limestone,
turbidites and black pyritic mudstone in the sedimentary
facies association (Jago et al., 1972; McPhie and Allen, 1992).
The presence of massive sulfide ore deposits, very thick
volcaniclastic mass-flow units, hyaloclastite, peperite and
pillow lava in the volcanic facies association are also consistent
with a subaqueous environment (Corbett, 1992; McPhie and
Allen, 1992; Waters and Wallace, 1992).
The essential elements of the facies architecture in
the Mount Read Volcanics are a variety of volcanic and
sedimentary facies associations that include lavas, lava
domes, synvolcanic intrusions and diverse volcaniclastic
units (McPhie and Allen, 1992). Lavas and sills occur
separately or in clusters in the succession (McPhie and
Allen, 1992). The four common types of volcaniclastic
facies associations in the Mount Read Volcanics are: (1)
very thick (tens of metres), massive to graded beds of
rhyolitic to dacitic pumice breccia; (2) very thick, massive
to diffusely stratified units of crystal-rich (feldspar, quartz,
clinopyroxene) sandstone; (3) thick to very thick, massive to
graded beds of polymictic volcanic conglomerate or breccia;
and (4) massive or laminated shard-rich siltstone (McPhie
and Allen, 1992). Many of these volcaniclastic facies contain
a high proportion of crystals, crystal fragments, shards and
pumice clasts, which are interpreted to be juvenile pyroclasts
transported by water-supported gravity flows (McPhie and
Allen, 1992). The sedimentary facies association comprises
conglomerate, sandstone, interbedded turbiditic sandstone
and mudstone, mudstone, and fossiliferous carbonate (Selley,
1997; Large et al., 2001a; McPhie and Allen, 2003). These
facies are of non-volcanic and mixed provenance, and include
pelagic marine sediment, meta-sedimentary and ultramafic
(bonninite, gabbro, peridote) rock fragments derived from
the Precambrian basement, and volcanic clasts and crystals.
There are regional variations in the proportion of volcanic
versus sedimentary facies, the types of volcanic facies and the
dominant magma composition. Volcanic facies associations
locally dominate the stratigraphy (e.g. at Rosebery) where-
as, elsewhere, volcanic facies are intercalated with, or sub-
ordinate to, sedimentary facies (e.g. in the hanging wall at
Hellyer). Parts of the volcanic succession are dominated
by the products of effusive, intrabasinal eruptions, such as
the andesitic lavas and domes in the footwall of the Hellyer
deposit (McPhie and Allen, 1992). In contrast, other areas
are dominated by volcanic facies generated by explosive
eruptions, such as the crystal and pumice-rich volcaniclastic
units of the White Spur Formation (McPhie and Allen, 1992;
McPhie and Allen, 2003). The volcanic facies associations also
display marked regional variations in composition. Rhyolite
and dacite dominate much of the succession (^90% of the
mapped area of the central Mount Read Volcanics: Gifkins
and Kimber, 2004); however, intermediate to mafic volcanic
facies are locally important at Hellyer and between Henty and
Queenstown (Corbett, 1992; Crawford et al., 1992; Large et
al., 2001a).
Post-depositional alteration processes
Formerly glassy or partly glassy volcanic rocks dominate
the Mount Read Volcanics. These rocks have textures and
compositions that reflect subsequent modification by a
variety of processes, which include: hydration, diagenesis,
hydrothermal alteration, regional and contact metamorphism,
and deformation.
ALTERATION I N SUBMARINE V OLCANIC SUCCESSIONS | 1 1
Two regional Cambrian diagenetic zones (albite zone
and epidote zone, Section 5.6) are preserved in the northern
Central Volcanic Complex (Gifkins, 2001). Locally, hydro-
thermal alteration and mineralisation was synchronous with
diagenesis. In addition, altered halos developed around thick
synvolcanic intrusions and Cambrian granites (Eastoe et al.,
1987; Large et al., 1996; Gifkins, 2001).
The Mount Read Volcanics were faulted during the
Middle to Late Cambrian and more extensively deformed,
folded and faulted during the Early to Middle Devonian
(Corbett and Lees, 1987; Crawford and Berry, 1992).
Regional metamorphism to lower greenschist facies produced
assemblages of quartz, albite, sericite, calcite, chlorite,
tremolite-actinolite, K-feldspar, epidote and biotite, and was
contemporaneous with the Devonian deformation (Corbett,
1981; Green et al., 1981; Walshe and Solomon, 1981;
Corbett and Solomon, 1989). During the Late Devonian to
Early Carboniferous, contact metamorphism was associated
with the intrusion of I- and S- type granites (Corbett and
Lees, 1987; Williams et al., 1989; Corbett, 1992).
Mineral deposits and prospects
In 2004, the Mount Read province contained two major
operating base-metal-sulfide mines (Rosebery and Mount
Lyell), one gold mine (Henty), and a number of exhausted
and smaller sub-economic deposits or prospects. Published
resource estimates are listed in Table 1.1.
The Hellyer polymetallic massive sulfide deposit was a
classic mound-shaped ore body discovered by a combination
of geophysical, geological and geochemical exploration
techniques in 1982 (Sise and Jack, 1984). Production
commenced in 1989 and mining was complete by 2000. The
high grade and simple geometry of the Hellyer ore body made
it a profitable operation although metal recoveries were low
due to the fine grainsize of the sulfides.
The Que River polymetallic massive sulfide deposit was a
small deposit comprising five steeply dipping ore lenses (four
Zn + Pb rich and one Cu rich). It was discovered in early 1974
by airborne electromagnetic and soil geochemical exploration
(Webster and Skey, 1979). Production from the Que River
deposit occurred from 1981 to late 1991.
Rosebery is the largest polymetallic massive sulfide deposit
in western Tasmania. It comprises at least 16 separate stacked
ore lenses over a strike length of 1.5 km. The deposit was
initially discovered in 1893 when prospectors traced gold and
Table 1.1 | Tonnages and grades of massive sulfide deposits in the Mount Read province (Data from Mineral Resources Tasmania, Pasminco
Mining and Exploration Goldfields P/L, and Gemmell and Fulton, 2001: in situ values based on average metal prices in 2000).
Deposit
Hellyer
Que River
Rosebery
Hercules
Henty-Mount Julia
Mount Lyell field
t x 1 0
6
16 .2
3.1
32.1
2.7
2.2
311
Zn
wt%
13.9
13.5
14.7
15.9
-
0.04
Pb
wt%
7.1
7.5
4.5
5.1
-
0.01
Cu
wt%
0.4
0.6
0.58
0.4
-
0.97
Ag
g/t
168
200
146
159
-
7
Au
g/t
2.5
3.4
2.3
2.54
12.1
0.31
In situ value
USS billion
3.99
0.81
7.76
0.71
0.23
6 .83
Form
Massive lens
Stratabound sheet
Stratabound sheets
Stratabound sheets
Sheet-like
Disseminated
Status
Past producer
Past producer
Current mine
Past producer
Current mine
Current mine
in stratigraphic order, the Puddler Creek, Mount Windsor,
Trooper Creek and Rollston Range formations (Henderson,
1986; Paulick and McPhie, 1999). These have a total thickness
of at least 14 km (Henderson, 1986) and generally young to
the south. The four formations respectively represent initial
continent-derived sedimentation and minor rift-related mafic
volcanism, succeeded by voluminous eruptions of crustally
derived rhyolitic magmas, abruptly followed by mixed mafic-
felsic volcanism derived from subduction-modified mantle,
and culminating in deep-water fine-grained sedimentation
(Stolz, 1995). The stratigraphic relationships and lithologies
are summarised in Figure 1.8.
ALTERATION IN SUBMARINE VOLCANIC SUCCESSIONS | 13
Rollston Range Formation
The uppermost formation of the Seventy Mile Range Group,
the Rollston Range Formation, is poorly exposed, except in
the southern central part of the belt where it has a minimum
thickness of 1 km. It consists of Early Ordovician, fossiliferous,
thinly bedded sandstone and siltstone of largely volcanic
provenance. Minor intervals of felsic lava and volcaniclastic
units exist locally (Berry et al., 1992).
Submarine facies associations and architecture
Puddler Creek Formation
The Puddler Creek Formation is the oldest formation in the
Seventy Mile Range Group and consists mainly of massive to
laminated lithic sandstone, greywacke and siltstone of mixed
continental and volcanic derivation. Minor altered trachy-
andesitic to trachytic coherent volcanic rocks are intercalated
with clastic rocks in the upper few hundred metres. The
volcanic rocks have geochemical signatures indicating an
alkali intraplate association related to lithospheric thinning
and incipient back-arc basin development (Stolz, 1995). The
formation is up to 9 km thick in the western part of the belt
and is partly stoped out by the Ravenswood Batholith in the
east.
The volcanic facies associations in the vicinity of the Highway-
Reward, Liontown and Thalanga base-metal sulfide deposits
are known from several detailed deposit scale studies (Hill,
1996; Miller, 1996; Doyle, 1997; Paulick, 1999; Paulick and
McPhie, 1999; Doyle and McPhie, 2000). Recent regional
studies of volcanic facies and lithostratigraphy contribute to
an improved understanding of the volcanic facies associations
(Simpson and McPhie, 1998; Simpson, 2001); however, much
of the regional data is currently confidential or unpublished.
In summary, the Mount Windsor and Trooper Creek
Formations comprise deep submarine volcanic facies that
include pyroclasts, probably from both subaerial and
submarine explosive eruptions. Lithofacies such as sparse
microbialitic ironstones (Simpson and McPhie, 1998)
indicate shallow marine settings for the Mount Windsor
Subprovince.
Mount Windsor Formation
The Mount Windsor Formation is a 0.4 to 5 km-thick
succession of subaqueous rhyolitic volcanic rocks dominated
by thick lavas, domes and high-level intrusions with sub-
ordinate volcaniclastic breccias. Isotopic evidence (Stolz,
1995) suggested the magmas were derived from the melting
of continental crust.
Trooper Creek Formation
The Trooper Creek Formation is a 0.5 to 2 km-thick
succession of highly variable basaltic-andesitic, dacitic and
rhyolitic coherent and brecciated volcanic rocks intercalated
with abundant volcanogenic siltstone, and minor calcareous
meta-sedimentary rocks. It is internally heterogenous and there
are major lateral variations in the proportion of volcanic and
sedimentary facies. Base-metal sulfide deposits and exhalative
siliceous ironstones occur at various stratigraphic levels (Fig.
1.8: Duhig et al., 1992). Stolz (1995) suggested that the
Trooper Creek Formation volcanic rocks were derived from
a melted subduction-modified, sub-arc mantle wedge that
was erupted during back-arc extension. Decreased volcanic
activity, or an increase in clastic sedimentation, appears to
have coincided with the change from exclusively rhyolitic
volcanism in the preceding Mount Windsor Formation.
FIGURE 1.8 | Simplified stratigraphic column for the Seventy Mile Range
Group. Modified after Large (1992) and Paulick and McPhie (1999).
14 I CHAPTER 1
Table 1.2 | Tonnages and grades of massive sulfide deposits in the Mount Windsor Subprovince (data from Berry et al., 1992; Large,
1992: in situ values based on average metal prices in 2000).
Deposit
Thalanga
Highway-Reward
Liontown
Handcuff
Waterloo/Agincourt
Magpie
t x 1 0
6
6.6
3.7
1.8
1
0.4
0.3
Zn
wt%
8.4
-
6.2
10
19.7
15
Pb
wt%
2.6
-
2.2
0.4
2.8
2
Cu
wt%
1.8
6.2
0.5
0.6
3.8
2
Ag
g/t
69
-
29
8
94
30
Au
g/t
0.4
1.5
0.9
0.2
2
1
In situ value
US$ billion
1.02
0.47
0.18
0.13
0.13
0.06
Form
Stratabound sheet
Subvertical pipes
Stratabound sheet
Stratabound sheet
Stratabound sheet
Stratabound sheet
Status
Past producer
Current mine
Prospect
Prospect
Prospect
Prospect
Post-depositional alteration processes
Some of the least-altered felsic coherent facies of the Seventy
Mile Range Group commonly have relict spherulitic or
perlitic textures due to devitrification and hydration (Paulick
and McPhie, 1999; Doyle, 2001). Pseudoclastic textures,
attributed to domainal devitrification and subsequent
diagenesis of coherent rhyolites, are prominent in the Mount
Windsor Formation at Thalanga and probably elsewhere.
Zones of intense hydrothermal alteration comprising quartz
+ sericite + chlorite + pyrite + carbonate assemblages partly
enclose the major sulfide deposits. They vary in style from
the broadly stratabound zone extending laterally beneath
the Thalanga deposit, to the discordant concentric zones
enveloping the Highway-Reward sulfide pipes. Early dia-
genetic and hydrothermal alteration facies were overprinted
by regional deformation coeval with extensive intrusion of
Mid-Late Ordovician gneissic granitoids (Berry et al., 1992).
This deformation produced relatively low-pressure regional
metamorphic assemblages that range from prehnite grade in
the east, to upper greenschist grade in the west, and a near
vertical axial planar cleavage. Subsequent intrusion of post-
kinematic Siluro-Devonian plutons in the central and eastern
parts of the subprovince produced contact metamorphic
aureoles with assemblages up to amphibole-hornfels grade.
Despite the multiple alteration processes, well preserved
primary volcanic textures that enable detailed interpretations
of facies associations occur away from zones of hydrothermal
alteration and mineralisation (e.g. Simpson and McPhie,
1998). Even in intensely hydrothermally altered rocks, the
existence of resistant primary components such as quartz
phenocrysts allow volcanic facies interpretation (e.g. Paulick
and McPhie, 1999).
Mineral deposits and prospects
The Mount Windsor Subprovince contains two major base-
metal sulfide deposits and several small sub-economic deposits
and historical prospects. Published resource estimates are
listed in Table 1.2. The known deposits are all in the Trooper
Creek Formation. The two largest deposits, Thalanga and
Highway-Reward, exist at the base and near the top of the
formation respectively (Fig. 1.8).
A gossanous outcrop led to the 1975 discovery of the
Thalanga deposit and its eventual development for open pit
and underground mining. Production between 1990 and
1998 amounted to 4.7 Mt from an estimated total resource
of 6.6 Mt. Thalanga mine was not highly profitable, mainly
because of ore dilution in underground mining and stability
problems related to the thin ore lenses.
The Highway-Reward Cu-Au deposit consists of two
discordant, vertical pipe-like bodies of massive pyrite about
200 m apart. Originally discovered in a surface road-metal
scrape in 1953 (Beams et al., 1998), the Highway-Reward
deposit has been the subject of intense but sporadic
exploration. Open pit mining of small oxide and supergene
high-grade Cu-Au resources occurred during the late 1980s
and from 1997 to the present. The deeper hypogene parts of
the sulfide pipes remain undeveloped.
I 1 5
2 I DESCRIBING ALTERED VOLCANIC ROCKS
This chapter addresses some of the common problems that
we face when studying altered volcanic rocks. Recognising
and describing the characteristics of the altered rocks is an
important step towards understanding the processes of
alteration. Alteration involves complex modifications of
the pre-existing rock and can encompass mineralogical,
textural and compositional changes. Resolving these complex
relationships is dependent on a systematic multidisciplinary
descriptive approach incorporating aspects of volcanology, ore
deposit geology, petrology and geochemistry. Unfortunately
relatively few studies adequately integrate these datasets.
Studies of ore deposits generally describe the characteristics
of the host rocks (i.e. lithology, petrology, geochemistry and
alteration) in separate sections of manuscripts. In many cases,
particularly in unpublished company reports, the geochemical
data and petrographic descriptions are in appendices,
discouraging integration and interpretation.
The integration of physical or textural observations and
geochemical data is a powerful tool in the study of altered
rocks. The physical characteristics and immobile element
concentrations of altered volcanic rocks can help to identify
the original rocks, where relict primary minerals and textures
are inconclusive (e.g. Paulick and McPhie, 1999; Barrett et al.,
2001). Physical and chemical changes that occurred during
alteration may help to determine the degree of alteration (i.e.
alteration intensity), the style of alteration (i.e. isochemical
versus metasomatic), and to discriminate between alteration
processes such as diagenesis, metamorphism and hydrothermal
alteration (e.g. Offler and Whitford, 1992; Gifkins and Allen,
2001). In addition, this integrated approach can lead to the
development of vectors to guide explorers toward ore deposits
(e.g. Large et al., 2001c).
This chapter compares alternative schemes of alteration
nomenclature and presents a multi-variable system for
describing and naming alteration facies. The different
elements of this descriptive approach to nomenclature are
explained in detail in subsequent sections and chapters (i.e.
alteration mineral assemblage in Section 2.4, alteration
intensity in Section 2.5, alteration textures in Section 3.1
and 3.2, distribution and zonation in Section 3.3, and timing
in Section 3.5). It also explains alteration indices, and the
physical and geochemical techniques for determining the
intensity of alteration. Alteration data sheets, which visually
combine mineralogical, textural and chemical data for altered
volcanic rocks, are introduced. These alteration data sheets are
used in Chapters 5, 6 and 7 to present examples of alteration
facies associated with various alteration processes and different
VHMS deposits.
2.1 | FREQUENTLY ASKED QUESTIONS
Most geologists are introduced to the basic principles of
hydrothermal alteration when they are students. However,
they typically have a limited knowledge of how to recognise,
characterise and interpret altered rocks. Common questions
are:
Was the rock altered?
What was the nature of the alteration?
Was the rock hydrothermally altered?
How do we name the altered rock?
What was the original rock?
To address these questions we need to make some simple
observations, which include the recognition of primary
minerals and textures, alteration colour, mineral assemblage,
texture and intensity, overprinting relationships, and alter-
ation distribution patterns. These observations can be made
at a variety of scales: map, outcrop, hand-specimen and thin-
section scales.
Was the rock altered?
Few volcanic rocks in submarine settings are entirely un-
altered and most altered rocks are easily recognised as such.
The most effective method of determining if a rock is altered
is by comparing it with other samples from the same unit.
Observed differences in mineral assemblage, texture and
colour may indicate a spectrum from fresh, or least-altered, to
significantly altered samples.
Some indicators of alteration in submarine volcanic facies
may be:
absence of glass
colour differences
presence of abundant minerals that typically form during
alteration, such as clays, zeolites, chlorite, micas, kaolinite,
1 6 | CHAPTER 2
tourmaline, apatite, alunite, epidote, carbonates and quartz
association between a distinctive mineral assemblage and
sulfides
presence of halos around veins, faults, intrusions and
mineralised rock
lack of, or only partial preservation of, primary textures
hardness of the rock: if not silicified, altered volcanic rocks
tend to be softer than unaltered volcanic rocks, which are
typically glassy or crystalline, hard and brittle
degree of deformation: clay- or phyllosilicate-altered rocks
are commonly more deformed than unaltered or least-
altered rocks because deformation-related strain is typically
partitioned into softer altered rocks.
What was the nature of the alteration?
Characterising the nature of the alteration can be challenging.
Nevertheless, systematic descriptions of alteration mineral
assemblages, alteration textures, preservation of relict minerals
and textures, patterns of distribution and overprinting
relationships combined with interpretations of alteration
indices and compositional changes provide important
information for subsequent classification and genetic
interpretation. To ensure that the data are meaningful,
systematic schemes for core logging and sample description
should be employed. Figures 2.1 and 2.2 are examples where
detailed observations in drill core and hand specimen have
established the characteristics of the altered rocks.
Was the rock hydrothermally altered?
Hydrothermal alteration can be discriminated from
metamorphism and diagenesis, which are typically regionally
extensive processes that result in weakly altered rocks and
preserve delicate volcanic textures. In contrast, hydrothermal
alteration styles, especially those associated with mineral-
isation, are local in their distribution, have variable intensity
(from weak to intense) and generally destroy primary
textures.
Discriminating accurately between different alteration
styles (Section 8.1) requires knowledge of: the host rock;
alteration intensity; distribution; timing; mineralogical,
textural and chemical changes; and comparison with changes
related to diagenesis, metamorphism and hydrothermal
alteration, which have been documented in well-preserved,
geologically young, submarine volcanic successions.
How do we name the altered rock?
Typically, rocks that are only weakly to moderately altered,
in which primary textures and minerals can be easily
recognised, are given precursor names (e.g. quartz-phyric
pumice breccia). In contrast, rocks that are intensely altered,
in which few primary textures or minerals are recognisable,
are given alteration names (e.g. quartz-augen schist or massive
chlorite rock). This is similar to metamorphic rocks where
low-grade rocks are given precursor names the prefix 'meta-'
is assumed and pervasively deformed and metamorphosed
rocks are given metamorphic names.
What was the original rock?
Outcrops and hand specimens of ancient volcanic rocks
rarely exhibit clear evidence of their modes of eruption and
emplacement. In many cases, the best we may hope for is
to recognise features that help distinguish coherent volcanic
facies from volcaniclastic facies.
The simplest approach to recognising the primary rock
is to move out of the altered zone and examine unaltered
rock. However, ancient volcanic successions rarely contain
unaltered rocks. As a result we rely on the preservation of
relict textures and minerals in altered rocks to provide a guide
to the interpretation of the primary volcanic facies. Relict
textures are original pre-alteration features that have not been
destroyed by alteration. Relict textures are most likely to be
visible in polished hand specimens with the aid of a hand lens
and in thin sections cut parallel to the tectonic foliation.
There are a small number of volcanic, devitrification
and hydration textures, and components or structures that
usually survive diagenesis, moderate hydrothermal alteration,
low-grade metamorphism and deformation - these are
particularly helpful in deciphering the primary volcanic facies.
For example, porphyritic texture, spherulites, lithophysae,
micropoikilitic texture, perlite, flow banding, columnar
joints and pillows are all characteristic of coherent volcanic
facies. Volcanic components such as pumice and scoria clasts,
glass shards, accretionary lapilli and non-vesicular volcanic
bombs or blocks, as well as bedding and cross stratification,
are characteristics of volcaniclastic facies. For a more detailed
discussion of volcanic, devitrification and hydration textures,
components and structures that help to determine the host
volcanic facies, readers are referred to McPhie et al. (1993)
Volcanic Textures: a guide to the interpretation of textures in
volcanic rocks.
Primary crystals and crystal fragments are found in a wide
variety of volcanic facies and can also be helpful indicators of
the host volcanic facies. Whole crystals and crystal fragments
in volcanic facies are mainly derived from porphyritic magmas.
Crystals may be liberated from magmas during volcanic
processes (explosive eruption or auto fragmentation) or by
surface sedimentary processes. The shape and distribution of
crystals in an altered volcanic rock can be used as a guide
to whether the primary facies was coherent or clastic. In
pyroclastic facies, angular and broken crystal fragments
are much more common than whole euhedral crystals. In
autoclastic facies, whole crystals and clusters of jigsaw-fit
crystal fragments are common. In contrast, coherent volcanic
facies typically, but not necessarily, contain very few broken
crystal fragments. The distribution of crystals and crystal
fragments in volcaniclastic facies may be random, related
to size or density sorting, or concentrated in particular
clasts, clusters or lenses. Crystal-bearing coherent facies are
porphyritic; they contain evenly distributed euhedral crystals
in a fine-grained or glassy groundmass.
Although relict textures and primary crystals can be used
as a guide, the discrimination of coherent and volcaniclastic
facies in altered volcanic rocks is not trivial. In originally
glassy volcanic rocks, alteration may produce convincing
pseudotextures such as pseudobreccia, false polymictic
texture, false thin-bedded texture and pseudomassive texture
(Section 3.2: Allen, 1988).
DESCRIBING ALTERED V OLCANIC ROCKS I 1 7
FIGURE 2.1 | Part of a drill core graphic log using a modified standard logging sheet, which incorporates volcanic and alteration fades descriptions.
This drill core, EHP319, is from western Tasmania and includes a thick interval of Central Volcanic Complex rocks. Abbreviations: S
o
= bedding, S, and
S
2
= tectonic foliations, LCA = long core axis, cc = calcite, chl = chlorite, fsp = feldspar, qtz = quartz, ser = sericite and gb = graded bedding.
1 8 | CHAPTER 2
FIGURE 2.2 I This sample description - for a rock sample from 245.2 m depth in drill core EHP319 - shows the main descriptive fields for alteration
studies. Abbreviations: S
2
= regional cleavage, chl = chlorite, fg = fine grained, fsp = feldspar, hem = hematite, plag = plagioclase, py = pyrite, qtz = quartz
and ser = sericite.
Crystal assemblages may reflect the primary volcanic
composition; they are relicts of original magmatic mineral
assemblages. For example, a rock containing abundant quartz
crystals was probably derived from a quartz-phyric magma,
which was likely of rhyolitic composition (Table 2.1).
In cases of low temperature (<200C), weak to moderate
intensity alteration, the alteration mineral assemblage may
also be a guide to the primary composition of the volcanic
fades. Alteration minerals rich in Fe, Mg and Ca are common
in mafic volcanic rocks; K- and Na-rich minerals in felsic
rocks. Typical alteration minerals in mafic rocks are chlorite,
epidote, calcite, palagonite, zeolites, albite, micas, actinolite-
tremolite and clays (Table 2.2). In contrast, common alteration
minerals in felsic rocks are quartz, micas, feldspars, zeolites,
cristobalite, opal and clays, especially montmorillonite and
kaolinite (Table 2.2). At temperatures above 200C and at
high water-rock ratios, the alteration mineral assemblage
formed is less dependent on primary host composition and
more on the fluid composition, temperature, permeability
and pressure (Browne, 1978; Henley and Ellis, 1983; Reyes,
1990).
Generally, consideration of a combination of field
relationships, relict textures and mineral assemblages will
enable interpretation of coherent versus clastic, and felsic
versus mafic volcanic facies, in all but the most intensely
altered volcanic rocks. Beyond that we must resort to
lithogeochemical techniques (Chapter 4).
DESCRIBING ALTERED V OLCANIC ROCKS | 1 9
2.2 I ALTERATION NOMENCLATURE
A variety of approaches have previously been taken to the
classification of alteration and altered rocks, particularly
hydrothermal alteration associated with different styles
of mineral deposits. Common methods of alteration
nomenclature are mineral based, compositional, generic, or
use terminology that reflects a combination of mineralogical
and textural characteristics (e.g. alteration facies). Discussions
of alteration nomenclature appear in Meyer and Hemley
(1967), Rose and Burt (1979), Beane (1982), Titley (1982),
Guilbert and Park (1986) and Thompson and Thompson
(1996).
Mineral-based alteration nomenclature
Classifying altered rocks in terms of mineral assemblage
was discussed in detail by Creasey (1959). Mineral-based
classification involves field and petrographic observations,
in some cases supported by other analytical techniques (e.g.
microprobe, X-ray diffraction and SWIR spectroscopy). It is
based on direct observations and provides the simplest non-
genetic approach to naming alteration and altered rocks.
There are two levels of mineral-based alteration
nomenclature: (1) terminology based on the dominant
TABLE 2.1 | Summary of the common volcanic rock compositions, their chemical classification (SiO2 content) and likely
phenocryst minerals. SiO2 contents for unaltered modern subduction-related volcanic rocks are from Ewart (1979).
Rhyolite
Dacite
Andesite
Basalt
>6 9 K-feldspar (orthoclase) quartz plagioclase biotite muscovite
amphibole pyroxene fayalite
quartz biotite amphibole pyroxene 6 3-6 9 Na-plagioclase
52-6 3 Na- or Ca-plagioclase + biotite quartz K-feldspar olivine
or amphibole or pyroxene
<52 Ca-plagioclase + pyroxene olivine hornblende
TABLE 2.2 | Common alteration minerals that replace glass and magmatic minerals in volcanic rocks. Alteration minerals are from
Schwartz (1959), White and Sigvaldason (196 2), lijima (1978), Hay (1978), Honnorez (1978), Brey and Schmincke (1980), Tucker
(1987) and Utada (1991).
Silicic volcanic glass
Mafic volcanic glass
Magnetite, ilmenite and titano-magnetite
Pyroxene, amphibole, olivine and biotite
Plagioclase
Anorthoclase, sanidine and orthoclase
Quartz
Zeolites (mordenite, clinoptilolite, laumonite, analcime, heulandite),
cristobalite, opaline silica, quartz, calcite, clays (montmorillonite, smectite,
mixed-layer clays)
Palagonite, nontronitic clays, smectite, calcite, chlorite, epidote, Ca-rich
zeolites, Fe/Ti/Mn-oxides
Pyrite, leucoxene, titanite, pyrrhotite, hematite
Chlorite, illite, quartz, calcite, pyrite, anhydrite
Calcite, albite, adularia, wairakite, quartz, anydrite, chlorite, illite, kaolin,
montmorillonite, epidote, sericite
Adularia, albite, sericite
Microcrystalline quartz
2 0 | CHAPTER 2
mineral; and (2) the use of the complete or abbreviated
alteration mineral assemblage. Some authors also use negative
mineral-based names, such as K-feldspar-destructive alteration
(e.g. Gustafson and Hunt, 1975).
The simplest method of alteration nomenclature uses the
dominant or most recognisable mineral phase in the altered
rock. Examples of this are albitic, which is dominated by
albite; silicic, dominated by quartz; chloritic, dominated by
chlorite; and sericitic, dominated by sericite. In addition,
the terms chloritisation, sericitsation, silicification and
carbonitisation are common in VHMS literature in reference
to the processes of alteration (e.g. Sangster, 1972; Paradis et
al., 1993). They, like the terms alteration and mineralisation,
are widely misused (Solomon, 1999).
Deciding which mineral is dominant in an altered zone
is not always a straightforward task. Several minerals may
be obvious and their proportions may vary. In addition,
common alteration minerals, such as sericite, can occur as the
dominant mineral in several different mineral assemblages
that have different origins, timing and economic significance.
In fact, sericitic assemblages are probably the most abundant
and widespread of all alteration assemblages. They are present
in aluminous rocks in nearly all types of hypogene alteration
associated with ore deposits (Meyer and Hemley, 1967).
Along with sericite, carbonates, chlorite and quartz are among
the most widespread alteration minerals (Meyer and Hemley,
1967).
Alternatively, more detailed alteration mineral assemblages
can be used; either complete assemblages of all the visible
alteration minerals, or abbreviated assemblages of the most
abundant and distinctive minerals. Minerals are usually listed
in order of decreasing abundance; thus a mixture of 60%
sericite, 35% quartz and 5% pyrite becomes the sericite +
quartz + pyrite alteration assemblage. This nomenclature has
the advantage of clearly defining the alteration assemblage.
However, some confusion may exist where mineral assemblages
contain identical or similar minerals in different abundances.
For example, sericite + quartz + pyrite is easily confused with
sericite + chlorite + quartz + pyrite.
Dana's Textbook of Mineralogy (Dana, 1957) defined
sericite as 'fine scaly muscovite united in fibrous aggregates
and characterized by its silky lustre'. The term has since
been widely used to refer to all fine-grained pale-coloured
micas, and indeed almost any fine-grained aggregates of
pale-coloured layer-lattice minerals (Whitten and Brooks,
1972), particularly in hydro thermally altered and low-grade
metamorphic rocks. White mica is the preferred term to avoid
the ambiguity in sericite where compositional differences such
as sodic muscovite, muscovite and phengite, may be important
(e.g. Yang, 1998). In this book, we always use sericite in the
loose sense, referring to fine-grained pale-coloured micas
of undetermined composition. We use the non-specific
alternative term: white mica, where compositions are known
(e.g. in discussions of mineral chemistry in Sections 4.2 and
8.2).
Compositional alteration nomenclature
Chemical methods of assessing hydrothermally altered
rocks (i.e. lithogeochemistry) have been applied in mineral
exploration, especially around VHMS and porphyry
Cu deposits, leading to the classification of alteration by
compositional changes that occurred during alteration
(Hemley and Jones, 1964). Examples of this include Na-
metasomatism or soda-metasomatism, Mg-metasomatism
and K-metasomatism (e.g. Hemley and Jones, 1964) or K-
enriched alteration, Ca-enriched alteration and Mg-enriched
alteration (e.g. Elliott-Meadows and Appleyard, 1991), and
Na-depleted alteration (e.g. Date et al., 1979, 1983).
There are several problems with compositional alter-
ation nomenclature: (1) it becomes increasingly complicated
where more than one element is mobilised during alteration,
which is almost always the case in the alteration of volcanic
rocks; (2) it cannot be applied in the field, as it requires a
detailed knowledge of the addition and removal of elements;
and (3) the character of the chemical alteration can only
be accurately determined if a least-altered protolith can be
unequivocally identified (Section 4.1).
Generic alteration nomenclature
A number of generic terms, such as advanced argillic,
intermediate argillic, phyllic or sericitic, potassic, propylitic,
skarn and greisen, have been applied to common alteration
mineral assemblages or groups of assemblages (Table 2.3:
Meyer and Hemley, 1967). These terms are widespread in
the geological literature, however they are not always clearly
defined or uniformly applied by different authors, and are
less precise than alteration assemblages. Many workers apply
generic terms based on the occurrence of indicator minerals
rather than complete alteration assemblages, with the result
that the terms are not always distinguishable in their usage
(Rose and Burt, 1979). To apply these terms rigorously,
alteration mineral assemblages for specific host rocks need to
be identified and correlated.
Generic alteration nomenclature tends to reflect detailed
work on altered rocks associated with particular deposit
types or geothermal systems, specifically porphyry, skarn,
mesothermal vein and epithermal deposits (Table 2.4). In
each case, the generic classification conveys a sense of the
mineralogical composition and implies knowledge of alteration
processes and environment of formation. Although generic
classification of altered rocks surrounding an ore deposit can
be useful, the reliance on understanding the environment
of formation can cause problems and may incorrectly imply
genetic processes. Thus, generic classification is best avoided
during the early stages of recognition, description, mapping
and interpretation of altered rocks in favour of a more rigorous
and descriptive classification.
Descriptive nomenclature alteration facies
The term alteration facieswas first proposed by Creasey (1959)
in an attempt to standardise the subdivision of hydrothermally
altered rocks in a similar manner to metamorphic facies,
which are assemblages of co-existing metamorphic minerals
that characterise particular pressure and temperature regimes
during metamorphism (Yardley, 1989). Creasey's concept of
three chemically and mineralogically distinctive alteration
DESCRIBING ALTERED VOLCANIC ROCKS | 21
TABLE 2.3 | Generic alteration terms based on common alteration mineral assemblages. Modified after Creasey (1959), Meyer and Hemley (196 7),
Lowell and Guilbert (1970), Rose (1970), Gustafson and Hunt (1975), Rose and Burt (1979), Beane and Titley (1981), Guilbert and Park (1986 ), Beane
(1982), and Thompson and Thompson (1996 ). Forsimplic,*, skarn implies a limestone or dolomite host rock.
Argillic
Advanced argillic
Intermediate
argillic
Phyllic (or sericitic)
Sericitic (or phyllic)
Propylitic (or
saussuritization)
Potassic
Greisen
Skarn
Calcic skarn
(or tactite)
Kaolinite (or halloysite, metahalloysite or dickite) +
montmorillonite sericite (or muscovite) chlorite
Pyrophyllite + kaolinite (or dickite) quartz sericite
andalusite diaspore alunite topaz zunyite
enargite tourmaline pyrite chalcopyrite hematite
Chlorite + sericite kaolinite montmorillonite illite-
smectite calcite epidote biotite pyrite
Sericite + quartz + pyrite biotite chlorite rutile
leucoxene chalcopyrite illite
(Note: K-feldspar absent)
Sericite + quartz + pyrite K-feldspar biotite calcite
dolomite chlorite andalusite chloritoid albite
pyrrhotite
Epidote (or zoisite or clinozoisite) + chlorite + albite
carbonate sericite montmorillonite septachlorite
apatite anhydrite ankerite hematite pyrite
chalcopyrite
K-feldspar (orthoclase) + biotite + quartz magnetite
sericite (or muscovite) albite chlorite anhydrite
apatite rutile epidote chalcopyrite bornite pyrite
Muscovite (or sericite) + quartz + topaz tourmaline
fluorite rutile cassiterite wolfranite magnetite
zunyite K-feldspar
Porphyry Cu, high-sulfidation epithermal, low-
sulfidation epithermal, geothermal
Porphyry Cu, high-sulfidation epithermal, low-
sulfidation epithermal, geothermal
Porphyry Cu, high-sulfidation epithermal
Porphyry Cu
Porphyry Cu, low-sulfidation epithermal,
geothermal, VHMS , sediment hosted massive
sulfide
Porphyry Cu, high-sulfidation epithermal, low-
sulfidation epithermal, geothermal
Porphyry Cu
Porphyry Cu, porphyry Sn
Pyroxene + garnet + wollastonite epidote (or zoisite) Porphyry, skarn
actinolite-termolite vesuvianite pyrite chalcopyrite
sphalerite
Magnesian skarn Forsterite + diopside + serpentine + talc actinolite-
tremolite calcite magnetite hematite chalcopyrite
pyrite sphalerite
Retrograde skarn
Jasperiod
Calcite + chlorite hematite pyrite
Quartz + pyrite + hematite
Porphyry, skarn
Porphyry, skarn
Sedimented-hosted Au, VHMS
fades propylitic, argillic and potassium silicate facies
was abandoned for a wide variety of generic and non-generic
terms.
Subsequently, Riverin and Hodgson (1980) proposed that
alteration facies be used as a descriptive term to refer simply
to altered rocks that could be identified during the course
of mapping or in hand specimen. They described a 'spotted
facies' characterised by a well-developed spotted texture due
to large, strongly altered, cordierite porphyroblasts, and a
'silicified facies' that lacked spots and was typically grey in
colour and siliceous in appearance.
More recently, the concept of alteration facies has
been expanded to incorporate other descriptive elements,
particularly alteration mineral assemblages (e.g. Gibson et al.,
1983; Elliott-Meadows and Appleyard, 1991; Paradis et al.,
1993; Tiwary and Deb, 1997; Brauhart et al., 1998; Gifkins
and Allen, 2001). Examples are 'mottled quartz-epidote' and
'silicifi cation alteration facies' described by Gibson et al. (1983)
in the Amulet Rhyolite of Noranda, Canada, and 'domainal
feldspar-quartz-sericite', 'fracture-controlled chlorite-sericite'
and 'stylolitic chlorite-sericite-hematite alteration facies'
described by Gifkins and Allen (2001) in a regional study of
alteration in the Mount Read Volcanics, western Tasmania.
The advantage of characterising alteration in terms of
alteration facies is that it is a purely descriptive scheme in
which the basic criteria used to classify the alteration can be
recognised and established in the field or in hand specimen.
More importantly, by using a combination of textural and
mineralogical terms, alteration facies convey the general
appearance of an altered rock. Also, the descriptive variables
in the alteration facies provide information that is critical to
subsequent genetic interpretations of the alteration process
(e.g. diagenetic, metamorphic, or hydro thermal).
22 | CHAPTER 2
TABLE 2.4 | Examples of different alteration nomenclature (i.e. dominant mineral, abbreviated mineral assemblage, compositional and generic terminology)
applied to altered rocks in a variety of ore deposit environments.
VHMS deposits
Silicic
Chloritic
Sericitic
Albitic
Carbonate
Porphyry deposits
Kaolinitic
Pyrophyllitic
Kaolinitic
Sericitic
Feldspathic
Biotitic
Chloritic
Epithermal deposits
Silicic
Al unite
K-mica or kaolinite
Chloritic
Sericitic
Sediment-hosted
deposits
Silicic
Silicic
Tourmaline
Carbonate
Sericitic
Albitic
Quartz + sericite + pyrite chlorite K-feldspar
Chlorite + pyrite + sericite quartz
Sericite quartz chlorite pyrite
Albite + sericite quartz
Dolomite/siderite/ankerite quartz sericite chlorite
pyrite
Kaolinite + montmorillonite sericite + chlorite
Pyrophyllite + kaolinite quartz sericite
Kaolinite + chlorite + sericite montmorillonite illite-
smectite calcite epidote biotite
Sericite + quartz + pyrite chlorite biotite
K-feldspar biotite quartz sericite albite
anhydrite epidote
Biotite + K-feldspar + magnetite quartz albite
anhydrite
Chlorite + epidote + albite carbonate + sericite
montmorillonite pyrite
Quartz chalcedony alunite barite pyrite
Alunite + kaolinite/dickite + quartz/cristobalite
pyrophyllite diaspore pyrite topaz andulusite
Kaolinite/dickite + illite-smectite quartz pyrite
Chlorite + calcite + epidote + albite pyrite
Sericite + illite-smectite quartz calcite dolomite
pyrite
Quartz + pyrite + hematite
Quartz muscovite carbonate + pyrite + pyrrhotite
Tourmaline muscovite quartz pyrrhotite
Ankerite/siderite/calcite + quartz muscovite pyrrhotite
Sericite + chlorite + quartz pyrrhotite pyrite albite
Albite + chlorite + muscovite biotite
Si-metasomatism
Mg-metasomatism
K-enrichment
Na-depletion
Ca, Mg, or Mn-metasomatism
K, Ca, Mg-metasomatism
K, Ca, Mg-metasomatism
K, Ca, Mg-metasomatism
Na, Ca, Mg-metasomatism
K-metasomatism
K-metasomatism
Ca-Mg-metasomatism
Si-enrichment
Ca, Mg, Na-depletion
K, Ca, Mg, Na-metasomatism
Ca, Mg-metasomatism
K-metasomatism
Not used in VHMS
literature
Argillic
Advanced argillic
Intermediate argillic
Phyllic
Potassic
Potassic
Propylitic
Silicic
Advanced argillic
Intermediate argillic
Propylitic
Argillic
Jasperiod
Tourmalinite
2.3 | ALTERATION FACIES -
THE RECOMMENDED METHOD
We advocate a multi-faceted, descriptive approach to studying
altered volcanic rocks. Different alteration facies can be
defined not only on the basis of their mineral assemblage and
texture, but also on distribution, intensity and composition
(or compositional changes). This approach to describing
and naming alteration facies is similar to the nomenclature
scheme adopted by McPhie et al. (1993) for volcanic facies
and the descriptive scheme for diagenetic calcite used by Folk
(1965).
The four alteration variables are: mineral assemblage,
texture, distribution and intensity (Fig. 2.3). However, because
it is not always practical to provide information on all four
variables, we suggest that the alteration mineral assemblage
and at least one other variable be used. Ideally, descriptive
names for alteration facies follow the formula: intensity +
distribution + texture + mineral assemblage.
The intensity variable (Section 2.5) provides information
on the degree of mineralogical, compositional and textural
modification (i.e. subtle, weak, moderate, strong or intense). It
is determined from petrographic descriptions in combination
with compositional data (e.g. Na
2
O) and alteration indices.
The distribution variable (Sections 3.3 and 3.4) refers to
the mappable extent of the alteration facies and its relationship
to host facies or components, structures, mineralised rock,
veins and other alteration assemblages (i.e. local or regional;
footwall or hanging wall; stratabound, pipe or plume).
The texture variable (Sections 3.1 and 3.2) refers to the
alteration texture that is superimposed on the rock, and is
typically described in hand specimen and/or thin section. It
may incorporate the shape, form, grainsize or fabric in the
altered rock (e.g. pervasive, selective or vein halo).
The alteration mineral assemblage (Section 2.4) is expressed
as an abbreviated alteration mineral assemblage in which the
minerals are listed in order of decreasing abundance (e.g. the
assemblage feldspar > quartz > sericite becomes feldspar +
quartz + sericite).
This approach produces alteration facies names such as
weak, regional, selective, chlorite + sericite alteration facies or
strong, massive, footwall, quartz + sericite alteration facies.
DESCRIBING ALTERED VOLCANIC ROCKS | 2 3
FIGURE 2.3 | Descriptive names for alteration facies.
2.4 | ALTERATION MINERAL
ASSEMBLAGE
Mineral assemblage refers to specific, and usually characteristic,
observed mineral associations that may be in equilibrium or
disequilibrium. An equilibrium mineral assemblage is a group
of minerals formed at the same time, lacking any indication
of disequilibrium, such as replacement or veining textures,
and hence interpreted to have formed due to the same process
and under the same fluid-rock conditions (Hemley and Jones,
1964). Disequilibrium or metastable mineral assemblages
are common and caution must be exercised in equating co-
existence with stable equilibrium (Meyer and Hemley, 1967;
Rose and Burt, 1979).
In general, alteration is a process of re-equilibration.
The pre-existing mineral constituents in a rock become
unstable under changed physicochemical conditions (e.g.
the addition of hydrothermal fluid) and progressively
alter to a new stable mineral assemblage, with or without
24 | CHAPTER 2
metasomatic chemical changes. The alteration process may be
only partially completed and may result in a disequilibrium
assemblage containing a mixture of the pre-existing and new
alteration minerals. Indeed, disequilibrium assemblages are
typical of altered volcanic rocks. Common examples, at low
metamorphic grade, are domainal devitrification of felsic glass
and incipient sericitisation of feldspar crystals.
Subsequent overprinting alteration may complicate dis-
equilibrium assemblages. Volcanic rocks commonly retain
relicts of primary minerals (especially as phenocrysts)
and alteration minerals from several stages of diagenetic,
metamorphic and/or hydrothermal alteration (e.g. Fig. 2.4).
Equilibrium assemblages may be attained in zones of intense
hydrothermal alteration or metamorphism, but primary
equilibrium assemblages are rarely preserved in ancient
volcanic rocks. This is true even in least-altered rocks.
When mapping altered rocks it is important to recognise
disequilibrium assemblages and correctly attribute minerals
to the various processes of formation. Equally important is
an understanding of the effects and constraints that earlier
alteration facies, at various scales, may impose on subsequent
processes.
Tools for mineralogical determination
The first steps in the identification of alteration minerals are
to apply the three essential field tools: the practised geological
eye, the hand lens and the scriber. These are frequently
adequate for useful descriptions of alteration mineral
assemblages, mapping of altered zones and interpretation of
styles or processes of alteration. Large-scale features must not
be overlooked; these provide the geological context that is
critical for interpretation of processes.
Simple chemical field tests, such as the use of dilute hydro-
chloric acid for discriminating carbonates, and sodium cobalt
nitrite for staining K-feldspar, can also be useful. However,
when alteration minerals occur as fine-grained masses
additional instrumental techniques, such as microscopic
petrography, short wavelength infrared spectrometry, X-ray
diffraction and micro-analyses, may be necessary to identify
them. In many situations, such as mineral exploration, the
practising geologist must rely largely on field skills, perhaps
augmented with limited laboratory work to substantiate
and assist in developing an 'eye' for particular mineral
assemblages.
Polished slabs
Stage 1: Hydration
Glassy plagioclase-phyric
coherent rhyolite with perlitic
fractures. Fracture surfaces
are coated with clay
minerals.
Stage 2: Diagenetic
alteration
Partly clay + zeolite-altered
plagioclase-phyric coherent
rhyolite. The alteration
facies distribution is
controlled by the perlitic
fracture pattern.
Stage 3: Hydrothermal
alteration
Moderately sericite + quartz-
altered plagioclase-phyric
coherent rhyolite.
Pervasively developed
sericite + quartz has
replaced all glass and
previously altered domains.
Some clay-altered relicts
have been altered to sericite.
Plagioclase phenocrysts are
partly altered to sericite.
Stage 4: Hydrothermal
alteration
Intense chlorite + pyrite-
altered plagioclase-phyric
coherent rhyolite. Vein-halo
chlorite + pyrite associated
with cross-cutting chlorite +
carbonate veins has
overprinted and destroyed
earlier clay and sericite +
quartz alteration
assemblages and textures.
FIGURE 2.4 | Cartoons of the microscopic textural and mineralogical evolution
of an originally glassy plagioclase-phyric coherent rhyolite. Overprinting
hydration, diagenesis and two stages of hydrothermal alteration are visible in the
final rock.
The identification of primary and alteration minerals, textures,
and overprinting relationships can often be facilitated by the
careful examination of polished slabs using a hand lens or
simple binocular microscope. Polished slabs can be made from
drill core or hand specimens that have been sawn to produce
a relatively flat surface, which is subsequently ground smooth
using a diamond lap. At this stage many coarser minerals and
textures will be visible on the wet surface. The resolution of
finer features can be improved by polishing the slab surface on
a rotating metal lap with 220 to 400-grit zinc or iron powder
and then finer powder (with water) on a glass plate. In the
absence of polishing equipment, it is sometimes beneficial to
buff the sawn surface with wet sandpaper. Steel wool can be
used to clean tarnished sulfides. Polished slabs are the cheapest
and most readily available tools for the field geologist.
Petrography
Examination of standard 75 x 25 mm thin sections or
polished thin sections with a polarising microscope is an
excellent and relatively inexpensive method of mineral
identification. It is the best way of resolving small-scale
spatial relationships between minerals to assist determination
of alteration reactions, paragenesis and likely processes.
DESCRIBING ALTERED VOLCANIC ROCKS | 2 5
Petrography is most effective if carried out by the person
who maps and samples the rocks. This requires access to
preparation facilities and a polarising microscope. This is not
a practical solution for mineral explorers, but is still widely
applicable in academia. The alternative is to send selected
samples to a consultant petrographer with complete details
of the geological context and the underlying objectives. Many
professional petrographers are unapologetic petrologists,
principally interested in petrogenesis and not enthusiastic
about the obscuring effect of alteration. Therefore, it is
imperative that the client informs the petrographer of the
importance of alteration mineral assemblages.
Short-wavelength infrared spectroscopy
The development of portable field instruments like the
PIMA (portable infrared mineral analyser), has increased the
use of short-wavelength infrared (SWIR) spectroscopy in
mineral exploration and related research (Thompson et al.,
1999). The technique identifies phyllosilicates, hydroxylated
silicates, carbonates and sulfates in most types of dry geological
samples and can also provide information on crystallinity and
compositional variations in some minerals, such as clays, white
mica and chlorite. These minerals, particularly phyllosilicates,
are common constituents of alteration mineral assemblages
and may be difficult to discriminate by other field or optical
methods.
Portable SWIR analysis has significant limitations in
resolving complex mineral assemblages, analysing dark
samples with significant opaque components and in
identifying aspectral anhydrous minerals, such as quartz.
It is an empirical method and does not supersede precise
determinative methods such as X-ray diffraction. Nevertheless,
portable SWIR has practical advantages including rapid in-
field analyses of up to 30 samples per hour and no sample
preparation other than drying. It has many applications in
the recognition and mapping of altered zones in a variety of
mineral deposit styles. Thompson et al. (1999) listed recently
published SWIR studies in epithermal, Archaean greenstone,
VHMS, uranium, evaporite and regolith environments.
Electron microprobe
Micro-analysis of mineral grains by electron microprobe has
become the standard tool for studies of mineral chemistry
over the past few decades. It has fine resolution, down to a
few microns diameter, and provides quantitative analyses
of elements with atomic numbers greater than four (Be) at
concentrations of greater than about 0.01 wt% (Berry et
al., 1983). Major element data can be used to estimate the
molecular formulae of unidentified minerals and investigate
spatial variations in mineral composition. Non-destructive
analyses are made on standard polished petrographic thin
sections or polished grain mounts. However, the electron
microprobe is an expensive laboratory instrument; it
requires a skilled operator and the sample throughput is low.
Consequently, it is essentially a research tool. It is rarely applied
in alteration studies, mineral exploration or mapping, but is
potentially useful for the verification of mineral identification
and spatial compositional variations interpreted by other
means.
X-ray diffraction
X-ray diffraction (XRD) is the definitive method for the
identification of all crystalline minerals, including opaque
minerals and structural polymorphs with similar chemical
compositions. Modern powder diffractometers can provide
semi-automated analysis and computerised semi-quantitative
mineral identifications from a small amount of powdered
sample (Berry et al., 1983). Like the electron microprobe,
these machines are mainly used as research tools. Commercial
quantitative XRD is not widely available and is relatively
expensive, currently around $75-95 per sample (AMDEL)
in Australia.
2.5 | ALTERATION INTENSITY
Alteration intensity is an indication of how completely a
rock has reacted to produce new minerals and textures,
and is independent of the alteration process. The alteration
intensity does not reflect the new mineral species, only their
abundance. It is closely linked to textural and compositional
changes because it reflects the extent to which pre-existing
textures and minerals (relicts of the original volcanic facies)
are preserved, and the degree of metasomatism (Rose and
Burt, 1979). Alteration intensity can be estimated both
qualitatively and quantitatively.
Qualitative estimates of alteration intensity
Qualitative estimates of the alteration intensity summarise
the textural and mineralogical changes that occurred during
alteration. They are based on the abundances of new alteration
minerals, the degree of destruction of pre-existing minerals,
the pervasiveness of alteration textures, and/or the degree of
preservation of pre-existing textures. Although these features
can be estimated to some extent in hand specimen, they are
commonly estimated petrographically. Many geoscientists
apply terms such as least altered, weakly altered, moderately
altered, strongly altered and intensely altered to describe
alteration intensity; however, these terms are subjective and
are rarely well defined.
Simmons and Christenson (1994) determined alteration
intensity by measuring the percentage conversion of primary to
secondary minerals, such that a weakly altered rock contained
0-33% alteration minerals; moderately altered 33-67%; and
strongly altered 67-100%. Alteration intensity can also be
measured by independently estimating the addition of new
minerals in the groundmass and the destruction of primary
phenocrysts such as plagioclase.
In contrast, Guilbert et al. (in Guilbert and Park, 1986)
proposed that alteration intensity be described in terms of both
the growth of new alteration minerals and the destruction of
pre-existing textures. Their two-part alteration intensity scale
incorporates estimates of the susceptibility of minerals to
alteration and the pervasiveness of alteration minerals. Mineral
susceptibility is the degree to which minerals in the rock are
altered, S1-S10 (vol%), whereas pervasiveness is the degree to
Despite these attempts to quantify alteration intensity,
it is still applied subjectively by most geologists. For this
reason we prefer to avoid a numerical system and retain the
descriptive terms subtle, weak, moderate, strong and intense.
The term least altered is reserved for rocks that are less altered
than their counterparts in the same environment. Least-
altered rocks may be weakly to moderately altered, especially
in hydrothermal environments where all rocks are altered to
some degree.
Here we define subtle, weak, moderate, strong and intense
alteration based on the extent of growth of new alteration
minerals, the destruction of primary minerals and textures,
and pervasiveness of alteration textures (Table 2.5). Typically
with increasing intensity of alteration, primary minerals are
progressively replaced, new minerals are more pervasively
distributed, primary textures are less consistently preserved,
and new textures are developed (Fig. 2.5). For example,
Gustafson and Hunt (1975) noted that with increasing
intensity of hydrothermal K-silicate alteration at the El
Salvador porphyry deposit in Chile, there is an increasing
degree of replacement of plagioclase phenocrysts by K-
bearing phases until the phenocrysts are obliterated. With
progressively more intense alteration, the mafic phenocrysts
are replaced, the groundmass becomes coarser grained with
K-feldspar overgrowths, magnetite and hematite disappear,
and the abundance of veins increases.
Estimates of alteration intensity that incorporate textural
changes are biased towards texturally destructive alteration
styles such as feldspar-destructive hydrothermal alteration.
It is important to recognise that under some circumstances
alteration, particularly carbonate alteration and some forms
of silicification, can enhance some primary or pre-existing
textures (e.g. Fig. 2.6: Titley, 1982; Allen, 1988). For
example, carbonate nodules preserve delicate shard textures
in the Hercules footwall, western Tasmania (Fig. 2.6A: Allen,
1997), and shards are preserved in quartz nodules and quartz
+ chlorite ( muscovite) altered zones in the Gossan Hill
footwall, Western Australia (Fig. 2.6B and C: Sharpe and
Gemmell, 2001). Although these alteration styles preserve
pre-existing textures, they may still be recognised as intensely
altered because of the pervasiveness of the new mineral
assemblage.
Colour contrasts related to overprinting alteration
assemblages or different mineral habits within an assemblage
can enhance textures, such as clast margins, whereas another
alteration assemblage with lower colour contrast and of equal
intensity may preserve textures just as well but textures may
be less discernable. The pervasiveness of alteration textures
and the degree of preservation of pre-existing textures are
dependent on the resilience of the pre-existing textures, the
intensity and style of alteration (Doyle, 2001; Gifkins and
Allen, 2001).
Quantitative estimates of alteration intensity
Alteration indices
Alteration indices are simple, multi-component or normalised
ratios of lithogeochemical composition data. They are usually
calculated from composition data expressed as weight
percentages (wt%) or parts per million (ppm), although in
some cases molar proportions are used. They are geochemical
representations of hydrothermal mineral assemblages designed
to facilitate discrimination of alteration styles, quantification
of alteration intensity, and exploration vectors. Alteration
indices have been widely applied in research and exploration
forVHMS deposits (Ishikawaetal., 1976; Large etal., 2001a).
They have also been used to a lesser extent in sediment hosted
Zn-Pb-Ag deposits (Large et al., 2001a) and Archaean lode
Au deposits (Eilu et al., 1997; Bierlein et al., 2000).
Simple ratio indices, especially of molar proportions, are
generally easily related to mineralogical changes (Eilu et al.,
1997). However, that is not the case for some more complex
indices where changes in the index value could be due to
changes in one or more of three or four components, and thus
related to several mineral phases. Stanley and Madeisky (1996)
noted that some empirically determined alteration indices are
not universally effective outside the district where they were
initially developed, tend to generate many false anomalies, or
may fail to identify significant altered zones, because losses of
one component may cancel out gains in another.
Alteration indices are formulated by placing proportions
of components that were gained during alteration in the
numerator and components that were lost in the denominator,
thus producing the highest values in the most intensely altered
rocks. In developing new indices, it is therefore useful to first
apply mass transfer techniques to determine the components
gained and lost.
Because alteration indices are ratios, they are less affected
by closure than composition data (closure is discussed
in Section 4.1). They respond only to changes in the
concentrations of those components used in the index, but
not to all other components of the rock. Nevertheless, and
contrary to the opinion of Eilu et al. (1997), alteration indices
are not independent of closure because each component
of composition data is affected by closure. Hence, major
components that dominate igneous rock compositions, such
as SiO
2
and A1
2
O
3
(which is also relatively immobile), are
rarely used in alteration indices for volcanic rocks.
Simple indices are ratios formulated from two components
of analytical data. For example, the S/Na
2
O ratio of Large et
al. (2001a). S/Na
2
O shows high contrast in VHMS alteration
systems, typically with values less than 0.1 in least-altered
rocks and values several orders of magnitude greater in sulfide-
bearing, intense, proximal altered footwall zones (Fig. 2.7).
Multi-component and normalised indices have two or more
components added together in either or both the numerator
and denominator of the index. The alkali based K
2
O/
Na
2
O + K
2
O + CaO index of Date et al. (1983) is a typical
example. It has a common structure for alteration indices: the
components of the numerator are also in the denominator.
This has a normalising effect of limiting the possible range
of values from zero to one. The normalisation in some
indices involves multiplication by a factor of one hundred to
2 6 [ CHAPTER 2
which alteration minerals permeate the entire rock, PIP10
(vol%).
DESCRIBING ALTERED V OLCANIC ROCKS | 2 7
TABLE 2.5 | Descriptive alteration intensity terms (subtle, weak, moderate, strong and intense) defined on the extent of growth of new alteration
minerals, the destruction of primary minerals and textures, and pervasiveness of alteration textures.
Subtle Phenocrysts and free New minerals have Primary volcanic, devitrification Minor replacement/recyrstailisation
crystals of feldspar, coated the surfaces of and hydration textures are clearly (micro- or cryptocrystalline,
quartz, and mafic existing phenocrysts, visible with little or no modification, overgrowths, poikilitic, microlitic,
minerals (amphiboles, fractures and clasts, spherulitic, variolitic and perlitic) and
pyroxenes etc.) were and infilled open space infill textures,
unaffected by alteration, (fractures, vesicles,
Plagioclase may have pore space, etc.). Glass
been dusted with has been devitrified.
sericite, carbonate or
hematite.
Weak Feldspar has been Patchy or domainal and Good preservation of most Replacement, dissolution,
partly replaced by albite, disseminated selective textures (original groundmass, recrystallisation, deformation and
sericite, carbonate, alteration styles. matrix textures and phenocrysts). infill textures. Most common textures
hematite and/or epidote. Alteration commonly Delicate textures such as shards, include: pseudomorphs, cleavage
Mafic minerals have nucleated on existing pumice clasts, perlite and the fine and rim texture, core and zonal
been partly replaced minerals, clasts or fibrous textures in spherulites texture, core and rim texture, skeletal
by Mg- and Fe-rich fractures and interstitial show some modification. - texture, overgrowths, micro- or
minerals, such as in glomerocrysts. - cryptocrystalline, dissolution vugs,
chlorite, epidote and J stylolites, poikilitic, foliations, fiamme,
Fe-oxides. and infill textures.
Moderate Feldspar has been partly Patchy or domainal Most textures modified and/or Replacement, dissolution,
to completely replaced and disseminated destroyed by alteration. Delicate recrystallisation, deformation, infill
by feldspar, sericite, alteration styles. textures commonly destroyed or and pseudotextures. In particular:
carbonate, epidote, Individual domains may substantially modified. Coarser pseudomorphs, partial pseudomorphs,
quartz and/or magnetite, have been texturally groundmass textures (perlite, overgrowths, disseminated nodules,
with outlines still destructive (i.e. chlorite spherulites, amygdales and flow spheriods, micro- or cryptocrystalline,
visible. Mafic minerals alteration of pumice banding) and clasts partially to dissolution vugs, stylolites, fiamme,
commonly completely clasts). Selective completely recrystallised but still porphyroblasts, poikiloblasts, poikilitic,
pseudomorphed. Minor alteration of individual clearly visible in domains. hornfelsic and augen textures, and
recrystallisation or clasts, groups of clasts foliations and lineations.
replacement of quartz. or minerals. V ein-halo
alteration.
Strong Feldspar has been Domainal selective to Primary volcanic, devitrification Replacement, dissolution,
completely replaced pervasive. V ein-halo and hydration textures almost recrystallisation, deformation, infill
by chlorite, sericite, alteration. completely destroyed (regardless and pseudo textures. Including:
carbonate and/or of grainsize). Pervasive pseudomorphs, nodules, spheroids,
opaques (although , replacement of groundmass, micro-or cryptocrystalline,
outlines still partly matrix and phenocrysts. Sparse dissolution vugs, stylolites, fiamme,
visible) and quartz relict fiamme, amygdales and clast porphyroblasts, poikiloblasts, poikilitic,
partly replaced or outlines preserved. granoblastic, decussate, hornfelsic
recrystallised. and augen textures, and foliations and
lineations.
Intense No primary minerals Transgresses textural All original rock textures including Replacement, dissolution,
remain. Sparse outlines facies and unit contacts phenocrysts have been destroyed, recrystallisation, deformation, infill and
after primary minerals and primary textures. Weak pseudomorphs or outlines pseudo textures. Including: nodules,
may still be visible. Pervasive, typically of coarse phenocrysts may spheriods, micro- or cryptocrystalline,
homogenous, alteration be visible. Primary rock type rare stylolites, granoblastic, decussate
on a local scale. V ein- indeterminate. and hornfelsic textures, and foliations
halo alteration. and lineations.
2 8 | CHAPTER 2
DESCRIBING ALTERED VOLCANIC ROCKS | 29
A. Bubble-wall shards
Delicate bubble-wall and platy shards (S) have been preserved
within a carbonate nodule in the proximal, carbonate zone
beneath the Hercules VHMS deposit. The carbonate nodule
comprises quartz + calcite + chlorite-altered pumice breccia.
Plane polarised light.
Sample MR96-57, Cambrian Hercules Pumice Formation,
Central Volcanic Complex, Mount Read Volcanics, Hercules
footwall, western Tasmania.
B. Pumice shards
Delicate tube pumice clasts (P) have beeen preserved in intensely
quartz + chlorite ( muscovite)-altered pumice breccia from the
footwall to the Gossan Hill VHMS deposit. The tube vesicles have
been coated in thin films of chlorite and filled with quartz, and
vesicle walls have been altered to quartz. Plane polarised light.
Sample 138752, Archaean Golden Grove Formation, Gossan
Hill footwall, Western Australia.
C. Shards
This quartz nodule (Q) from the footwall, quartz + chlorite
( muscovite) zone contains delicate shard textures. Plane
polarised light.
Sample 138795, Archaean Golden Grove Formation, Gossan
Hill footwall, Western Australia.
FIGURE 2.6 | Photographs of intensely altered pumice breccias with delicate primary textures.
FIGURE 2.5 | Pairs of hand-specimen and thin-section photographs of increasing intensity of alteration in rhyolitic feldspar-phyric pumice breccia in the Hercules
footwall, northern Central Volcanic Complex, western Tasmania. (A) Hand-specimen and (B) thin-section photographs of subtle, domainal, albite + sericite- and
sericite + chlorite-altered pumice breccia (sample MR96 -6 3) showing excellent preservation of volcanic textures. Plagioclase crystals are partly replaced by albite.
In albite-rich domains, tube vesicles and clast margins are lined with sericite and albite + quartz altered. In contrast, pumice clasts and shards in the chlorite-rich
domains are pervasively sericite + chlorite altered. The Al = 40 and CCPI = 26 . (C) Hand-specimen and (D) thin-section photographs of weak, domainal, albite +
sericite- and sericite + chlorite-altered pumice breccia (sample MR96 -54). Volcanic textures are well preserved in the albite-rich domains and poorly preserved in the
chlorite-rich domains. Plagioclase crystals (P) are sericite albite opaques altered and have albite overgrowths or nodules (alb), which locally preserve delicate
vesicular textures. Elsewhere vesicles are coated in sericite and filled with albite. Pumice walls are albite + quartz altered and sericite chlorite + hematite fiamme
and stylolites are abundant. The Al = 58 and CCPI = 37. (E) Hand-specimen and (F) thin-section photographs of moderate, pervasive, albite + sericite-altered pumice
breccia (sample MR96 -48) with partly preserved pumice textures and plagioclase crystals. Sericite fiamme (F) and sericite + hematite stylolites are abundant.
Nodules or overgrowths of albite occur around calcite and albite + hematite-altered plagioclase crystals (P). The Al = 70 and CCPI = 38. (G) Hand-specimen and (H)
thin-section photographs of strong, pervasive, quartz + sericite + pyrite-altered pumice breccia (sample MR96 -50). Primary volcanic textures are faint, with sparse
sericite-altered pumice clasts or fiamme (F). Plagioclase crystals (P) are polycrystalline-quartz pyrite altered. The Al = 98 and CCPI = 6 4. (I) Hand-specimen and
(J) thin-section photographs of intense, schistose, quartz + sericite + pyrite-altered pumice breccia (sample MR96 -46 ). No relict plagioclase or volcanic textures are
preserved in thin section: in hand specimen irregular lenses of sericite resemble fiamme (F). This alteration facies is pervasive and strongly foliated. The Al = 99 and
CCPI = 30.
30 | CHAPTER 2
FIGURE 2.7 | West-east 1700mN section through the K-lens of the Rosebery VHMS deposit, western Tasmania, showing geology and
contoured S/Na2O data.
produce a potential range from zero to one hundred, which
is convenient for quantification of alteration intensity. The
classic example is the Alteration Index (AI) of Ishikawa et al.
(1976):
AI =
100(MgO + K
2
O)
MgO K
2
O CaO + Na
2
O
Originally devised as a measure of intensity of sericite and
chlorite alteration associated with the Kuroko-VHMS
deposits, it is useful in many types of plagioclase-destructive
hydrothermal alteration systems.
In some cases where there are large differences in
magnitudes between components, some components are
multiplied by appropriate factors to adjust their effect in the
index. An example is the AI mark 4 index,
AI mark 4 =
100(FeO + lOMnO)
FeO + lOMnO + MgO + (SiO
2
/10)
which quantifies alteration in siliciclastic dolomites (Large et
al., 2000).
Molar proportion alteration indices are said to be more easily
related to the stoichiometry of alteration reactions and hence
to alteration assemblages (e.g. Eilu et al., 1997). The extra
step in converting composition data to molar proportions of
oxides or elements is easily achieved in computer spreadsheets
but it significantly complicates manual calculations. Some
examples of molar indices are the 3K/A1 sericitisation index
and the CO
2
/CaO carbonation index used in exploration
for lode Au deposits (Davies et al., 1990). The ACNK
index of Hodges and Manojlovic (1993) used the molecular
proportions of Al
2
O
3
/(CaO + Na
2
O + K
2
O) to quantify
intensity of alteration related to metamorphosed massive
sulfide deposits at Snow Lake, Manitoba.
DESCRIBING ALTERED V OLCANIC ROCKS | 31
The AI-CCPI alteration indices and box plot
The well-known Alteration Index (AI) was developed in the
Kuroko VHMS deposits, Japan, to represent the principal
components gained (MgO and K
2
O) during chlorite and
sericite alteration, and those lost (Na
2
O and CaO) during
the breakdown of Na-plagioclase and volcanic glass (Ishikawa
et al., 1976). The AI has since been widely used in VHMS
mineral exploration to provide quantitative estimates of the
intensity of alteration. It typically increases to maximum values
in the proximal hydrothermal zones beneath massive sulfide
lenses (e.g. Saeki and Date, 1980). The AI ranges from 0 to
100. High (> 60) values reflect high MgO and K
2
O contents
relative to CaO and Na
2
O, and may be related to intense
hydrothermal sericite and chlorite alteration. In contrast, low
(< 30) AI values reflect high CaO or Na
2
O contents that may
be due to intense albite or calcite alteration more typical of
regional diagenetic alteration or metamorphism. For example,
at Hellyer AI increases from 35 to 95 from the margin to the
centre of the alteration pipe directly below the ore deposit
(Fig. 2.8A and B: Gemmell and Large, 1992; Large et al.,
2001a).
There is a strong inverse relationship between AI and Na
concentration (e.g. Fig. 2.8A and D) because loss of Na, and
sometimes loss of Ca, is the major chemical change involved
in the breakdown of plagioclase. In many studies Na depletion
is used instead of AI as the principal measure of alteration
intensity (Date et al., 1983).
The Ishikawa alteration index has two major limitations
(Large et al., 2001a). Firstly, it does not take carbonate
alteration into account, even though this type of alteration
can be significant in some VHMS alteration systems. Where
Ca-carbonates are present they cause a decrease in AI, even
where plagioclase destruction is extreme, because CaO is
in the denominator. Secondly, the AI effectively measures
plagioclase destruction but does not differentiate chlorite-
from sericite-altered rocks. Variations in relative proportions
of chlorite and sericite or spatial relationships between chlorite
and sericite zones may be important guides to exploration
in some VHMS alteration systems. A geochemical index to
quantify the variation would be an improvement on subjective
visual estimates.
The chlorite-carbonate-pyrite index,
CCPI = 100(FeO +MgO)
FeO+MgO+Na
2
O+K
2
O
was developed to reflect the prominence of chlorite, Fe-
Mg carbonates, and pyrite, which are common minerals in
the proximal altered zones of many VHMS deposits (Large
et al., 2001a). High values of CCPI reflect high FeO and
MgO contents, suggesting intense alteration to Fe- or Mg-
rich minerals such as chlorite, Fe-Mg-bearing carbonates
(dolomite, ankerite or siderite), pyrite, magnetite or hematite.
However, the CCPI of least-altered rocks is dependent on
primary composition and magmatic fractionation. Mafic
rocks with high primary FeO and MgO contents typically
have CCPI values greater than 50, whereas more evolved felsic
rocks have lower CCPI values between 10 and 50. Thus the
CCPI is not well suited to the study of altered mafic rocks.
hangingwall Y
volcaniclastic unit \
FIGURE 2.8 | Alteration intensity in the altered footwall zones at the Hellyer
VHMS deposit, western Tasmania. (A) Schematic cross-section of the altered
footwall zones and variations in alteration intensity in these zones as measured
by (B) Alteration Index (Al), (C) Chlorite-carbonate-pyrite index (CCPI), and
(D) Na2O. Modified after Gemmell and Large (1992) and Large et al. (2001a).
Used in conjunction with the AI, particularly graphically
on x-y bivariate plots with AI as the x-axis, the CCPI provides
an effective means of discriminating sericite-, chlorite- and
carbonate-rich altered zones. Furthermore, the AI-CCPI
bivariate plot, termed the Alteration box plot by Large et al.
(2001a), discriminates these VHMS-related hydrothermal
alteration assemblages from diagenetic albite- or albite + K-
feldspar-bearing assemblages.
Feldspar, phyllosilicate, carbonate and several other
alteration mineral compositions plot around the margins of
the Alteration box plot (Fig. 2.9). Albite plots at the lower
left, K-feldspar and pure muscovite at the lower right, chlorite
at the top right, and carbonates along the upper margin et
cetera. Calcite plots at the top left corner (although CCPI is
indeterminate for pure calcite, the merest trace of Fe or Mg
will result in CCPI = 100), magnesite at the top right and
the Ca-Mg carbonates spread between them according to AI
32 | CHAPTER 2
0 1 0 2 0 30 4 0 5 0 6 0 7 0 8 0 90 1 00
Al (Ishikawa Alteration Index)
0 10 20 30 40 50 6 0 70 80 90 100
Al (Ishikawa Alteration Index)
FIGURE 2.9 | Al - CCPI Alteration box plot for least-altered samples from the
Mount Read V olcanics, western Tasmania (modified after Large et al., 2001a).
The data are classified according to Ti/Zr ratios, where rhyolites have Ti/Zr <10,
dacites 1 0-2 0 and andesites and basalts >20, and show the effect of magmatic
fractionation on the CCPI .
determined by Mg/Ca ratios. Similarly, Mn-carbonates (except
pure rhodochrosite, which is indeterminate in both indices)
plot along the top margin of the Alteration box plot; their
positions determined by the inevitable minor concentrations
of Ca, Mg and Fe. Siderite, pyrite and Fe-oxides have CCPI
values of 100, but are indeterminate for Al and thus plot as a
line, rather than a point, along the top.
Large et al. (2001a) found that least-altered rocks in the
Mount Read Volcanics have an Al range of 20 to 65 and a
CCPI range of 15 to 85 (Fig. 2.9). A compilation of 1734
geochemical analyses for unaltered volcanic rocks from
various modern volcanic arcs shows a slightly smaller Al range
of 20 to 60 and a slightly greater CCPI range of 10 to 90
(Fig. 2.10). Hence, least-altered volcanic rocks plot within a
rectangle near the middle (somewhat left of centre) of the
AI-CCPI bivariate plot. This is the least-altered 'box' that
inspired the term box plot. The extent and position of the
least-altered box may vary for data from different districts,
according to the diversity of primary compositions.
Fluid-dominated pervasive hydrothermal alteration
tends to produce simple equilibrium assemblages of only a
few phases. Therefore, intensely altered samples tend to plot
outside the least-altered box and towards the positions of
the dominant alteration minerals. For example, unaltered
calc-alkaline rhyolites plot in a box towards the centre of the
Alteration box plot; with increasing intensity of alteration,
altered samples plot progressively further away from the
unaltered box (Fig. 2.11). The relative direction of movement
away from the unaltered box is controlled by the alteration
assemblage and hence by the alteration process (Large et al.,
2001a).
Large et al. (2001a) defined 10 different mineralogical
trends on the Alteration box plot. Six of these trends relate to
common VHMS hydrothermal alteration styles and four are
associated mainly with diagenetic alteration (Fig. 2.12).
FIGURE 2.10 | AI -CCPI Alteration box plot for 1734 analyses of rocks from
modern volcanic arcs; most are assumed to be unaltered. Geochemical data
are from Aleutian, Andean, Indonesian and Scotian volcanic arcs, and were
compiled by A.J. Stolz (electronic communication, 1998). These data (classified
by SiO2 content) show the effects of magmatic differentiation on CCPI , and to a
lesser degree on Al. Mafic rocks have high CCPI and low to moderate Al values
because of their high Fe, Mg and high Ca contents, respectively. In contrast,
felsic rocks have low CCPI because of their low Fe, Mg and high K contents, and
high Al due to their low Ca and high K contents.
It is important to note that neither Al nor CCPI includes
SiO
2
; thus the Alteration box plot does not provide a direct
measure of the intensity of quartz or silica alteration. As
outlined in Section 7.2, silica-altered rocks are important
around some VHMS deposits, including the silicic core zone
in the Hellyer alteration pipe (Gemmell and Large, 1992), the
stockwork zones of some Kuroko deposits (Shirozu, 1974),
and in some Cyprus-type alteration pipes (Lydon, 1984).
The Alteration box plot is a powerful tool for relating
lithogeochemical data to mineral assemblages and alteration
intensity in VHMS systems, particularly in felsic volcanic
rocks. It has obvious applications in mineral exploration for
recognising favourable alteration styles, delineating altered
zones and providing vectors to ore within large altered systems.
A similar dual index approach may be useful for other deposit
types, with different indices designed to highlight specific
alteration assemblages.
Element concentrations and mineral abundances
Element concentrations can also be used as guides to alteration
intensity. Many alteration studies have used Na depletion as
a measure of hydrothermal alteration intensity (e.g. Franklin
et al., 1975; Date et al., 1979; 1983; Hashiguchi et al., 1983;
Ashley et al., 1988). Typically unaltered modern arc calc-
alkaline rhyolites have Na
2
O values between 3 and 5 wt%
(Barrett et al., 1993; Stolz et al., 1996). Rhyolites with greater
than 5 wt% Na
2
O are normally albitised, whereas rhyolites
with less than 3 wt% Na
2
O reflect feldspar-destructive
alteration styles (e.g. sericite, chlorite, pyrite, K-feldspar and
carbonate). The lower the Na
2
O content the more intensely
hydrothermally altered the rock; thus, Na
2
O typically
DESCRIBING ALTERED V OLCANIC ROCKS I 33
0 10 2 0 30 4 0 5 0 6 0 7 0 8 0 90 1 00
Al (I shikawa Alteration I ndex)
FIGURE 2.11 | Al - CCPI Alteration box plot for rhyolites in the northern Central
Volcanic Complex, western Tasmania. Samples are of rhyolitic pumice breccias
from the Rosebery and Hercules footwalls. With increasing intensity of footwall
chlorite + sericite pyrite alteration, Al and CCPI values increase and samples
plot in the upper right of the Alteration box plot.
decreases towards the centre of VHMS alteration systems
(e.g. Fig. 2.13).
Alteration mineral abundances quantitatively estimated
from whole-rock composition data can also provide measures
of alteration intensity. For example, Large et al. (2001b) found
that calculated mineral percentages closely approximated the
petrographic estimates of alteration mineral abundances in
samples from Rosebery. By plotting the calculated mineral
abundances down hole they showed that diagenetic minerals,
such as albite, decrease in abundance and hydrothermal
minerals, such as sericite, quartz, chlorite and Mn-carbonate,
increase in abundance with proximity to ore (Fig. 2.14).
Mineral abundances can be calculated as percentages from
the whole-rock analyses by the least-squares method outlined
in Herrmann and Berry (2002). A free copy of the MINSQ
(least-squares spreadsheet method for calculating mineral
proportions from whole-rock major element analyses) is
available to download from the University of Tasmania's
Centre for Ore Deposit Research website <www.codes.utas.
edu.au>.
An integrated approach to alteration intensity
A combined compositional and descriptive approach to
estimating alteration intensity can also be used. In fact,
the Alteration box plot is most powerful when used in
combination with petrographic and/or other instrumental
mineralogical studies, such as X-ray diffraction (XRD) or
short wavelength infra-red spectral analysis (e.g. PIMA). Used
in this way the box plot reveals trends in the data, from least
altered to intensely altered, which can be related to alteration
processes and hence exploration targets.
FIGURE 2.12 | Schematic AI-CCPI Alteration box plots showing the 10
alteration trends recognised by Large et al. (2001a). These provide a tool for
graphically discriminating prospective from non-prospective altered zones and/or
systems. (A) The six trends marked by arrows on this box plot are typical of
hydrothermally altered rocks associated with VHMS deposits. Trend 1: sericite
alteration at the margins of the hydrothermal alteration halo in felsic volcanic
rocks. Trend 2: footwall sericite + chlorite pyrite alteration in felsic and mafic
volcanic rocks. Trend 3: chlorite sericite pyrite alteration, typical of footwall,
chlorite-dominated zones in either felsic or mafic volcanic rocks. Trend 4: chlorite
+ carbonate alteration typically developed proximal to massive sulfide lenses
in the footwall of either felsic or mafic host rocks. Trend 5: sericite + carbonate
alteration in the proximal hanging wall to ore deposits or along strike in the host
rocks. Trend 6 : K-feldspar + sericite, an uncommon trend developed locally
within footwall felsic volcanic rocks. (B) The four trends marked by arrows on
this box plot are mainly attributed to diagenetic processes and are unrelated to
mineralisation. Trend 7: albite + chlorite alteration, typical of low temperature
seawater-volcanic rock interaction. Trend 8: epidote + calcite albite alteration
common in intermediate and mafic volcanic rocks. Trend 9: K-feldspar + albite
alteration. Trend 10: paragonitic sericite + albite alteration.
34 | CHAPTER 2
FIGURE 2.13 | Contoured Na2O data for
the west-east 1700 mN section through
the K-lens of the Rosebery VHMS
deposit, western Tasmania. Modified after
Large etal. (2001b).
In the northern Central Volcanic Complex (Mount Read
Volcanics), detailed petrographic descriptions were combined
with compositional data to assess the range of AI and CCPI
values for subtly to intensely altered rhyolites (Table 2.6 and
Fig. 2.15). The least-altered rhyolites were subtly altered and
have comparable alteration indices to unaltered modern arc
rhyolites (AI = 30-60 and CCPI = 10-40). Intensely altered
rhyolites have mid to high alteration indices (AI = 40-100 and
CCPI = 28-100). AI ranges for subtly, weakly, moderately,
strongly or intensely altered rhyolites, dacites, andesites and
basalts are similar. In contrast, the CCPI is influenced by the
primary composition. For this reason the Alteration box plot
should never be used independently as a method of classifying
the alteration system, but should be integrated with the
primary geochemical and/or petrographic data.
Using a combination of alteration mineral assemblage and
composition data also enables separation of rock samples into
least-altered, diagenetically altered and hydrothermally altered
samples (Gifkins and Allen, 2001; Large et al., 2001a). The
trend from subtly to intensely hydrothermally altered rocks
associated with VHMS deposits is characterised by increases
in both CCPI and AI, and decreases in Na
2
O (Table 2.7).
TABLE 2.6 | Alteration indices for altered rhyolites in the northern Central
Volcanic Complex, western Tasmania. The broad range in Al and CCPI values
reflects different alteration styles. For example, strongly altered rhyolites with low
Al values probably reflect diagenetic alteration, whereas high Al values reflect
hydrothermal alteration.
TABLE 2.7 | Alteration indices, Na2O contents and approximate mass changes
for hydrothermally altered rhyolites from the footwalls to the Rosebery and
Hercules VHMS deposits, western Tasmania. AI and CCPI increase and Na2O
decreases with increasing intensity of alteration.
nyuruinermai aiieiauon.
Alteration intensity
Subtle
Weak
Moderate
Strong
Intense
AI
30-55
25-6 0
10-75
5-90
40-100
CCPI
10-32
1 5 ^5
10-55
28-90
28-100
Na2O (wt%)
3.5-5
2-5.5
1-6
0.5-4.5
0-3
Alteration
intensity
Subtle
Weak
Moderate
Strong
Intense
AI
30-55
40-6 0
40-75
70-90
90-100
CCPI
10-32
15-45
10-55
30-90
28-100
Na 2 O
(wt%)
3.5-5
2-A
1-2
0.5-1
0-0.5
Mass changes
(g/100g)
<1
<10
5-30
15-6 0
15-100
DESCRIBING ALTERED VOLCANIC ROCKS | 35
FIGURE 2.14 | Variations in calculated mineral abundances in samples from DDH 120R through K lens of the Rosebery VHMS
deposit, western Tasmania. Increasing intensity of hydrothermai alteration towards the ore lens corresponds with increasing proportions
of chlorite, Mn-carbonate and calcite, and decreasing concentrations of quartz and albite.
36 | CHAPTER 2
0 1 0 2 0 30 4 0 5 0 6 0 7 0 8 0 90 1 00
Al (I shi kawa Al terati on I ndex)
intense A moderate subtle
strong o weak
FIGURE 2.15 | Bivariant plots of rhyolite samples from the northern Central
Volcanic Complex, western Tasmania, which have been classified qualitatively.
(A) AI-CCPI Alteration box plot for subtly, weakly, moderately, strongly and
intensely altered rhyolites. With increasing intensity of alteration rhyolites
plot away from the subtle box in all directions depending on the alteration
composition and hence on the processes. (B) Al versus Na2O for subtly, weakly,
moderately, strongly and intensely altered rhyolites. With increasing intensity of
alteration rhyolites plot away from the subtle field.
2.6 | ALTERATION DATASHEETS
A practical way of integrating different alteration data,
constructing descriptive names and defining alteration
facies is to use alteration data sheets. These visually display
a combination of different data types for a specific rock or
alteration facies on one page. They present all the relevant
data for a particular rock sample or alteration facies together,
and graphically illustrate relationships between texture,
mineral assemblage, composition and alteration zonation,
particularly with respect to an ore body. Alteration data sheets
act as 'flash cards', incorporating the distinctive physical and
chemical characteristics of the altered rock or alteration facies,
providing quick reference to the data collected and facilitating
interpretation. Data sheets are used in Chapters 5, 6 and 7 to
illustrate the dominant alteration facies or zones associated
with each of the case studies. The information that is included
on the data sheets may vary because the relevant or available
data varies in different volcanic successions and in different
deposits. Where appropriate, data sheets may incorporate:
sample number
location information
geographical or geological feature
formation or group
succession
coordinates
map, cross-section, or alteration zonation model showing
the location of the sample or alteration facies
volcanic facies characteristics
descriptive name for the volcanic facies (see McPhie et al.
(1993) for guidelines)
relict primary minerals
composition (e.g. rhyolitic, dacitic, andesitic or basaltic)
estimated from relict primary minerals and/or geochemical
data
lithofacies characteristics
relict textures
interpretation of the volcanic facies and application of
genetic nomenclature (e.g. volcanogenic sedimentary
deposit, resedimented mass-flow or turbidite deposit,
syneruptive mass-flow deposit, autobreccia, hyaloclastite,
peperite, pyroclastic-flow deposit, pyroclastic-fall deposit
or pyroclastic-surge deposit)
alteration facies characteristics
descriptive name for the alteration facies
alteration mineral assemblage
alteration textures
distribution or zonation of alteration facies
alteration intensity
relative timing
interpretation of the alteration process (i.e. diagenetic,
metamorphic, hydration, intrusion-related hydrothermal
alteration, proximal or regional hydrothermal alteration
and mineralisation, syntectonic hydrothermal alteration)
photographs of distinctive features of the alteration facies
in outcrop, drill core, hand specimen or thin section
composition data (whole-rock, mineral-chemistry or
isotope analyses)
chemical characteristics such as important mass changes,
alteration indices (e.g. Al and CCPI) and immobile
element ratios (e.g. Ti/Zr)
Alteration box plot, with the sample highlighted
other significant compositional plots, such as Ti/Zr-SiO
2
bivariant plot, mass change bar graph, SWIR spectra et
cetera.
I 37
3 | COMMON ALTERATION TEXTURES AND
ZONATION PATTERNS
This chapter describes common alteration textures,
pseudotextures, and alteration distribution and zonation
patterns in submarine volcanic successions, which can be
observed at a variety of scales: map, outcrop, hand specimen
and thin section. It also discusses the use of overprinting
relationships in determining the paragenetic sequence.
Alteration textures, patterns of distribution and zonation,
and overprinting relationships are fundamental elements in
describing and interpreting alteration facies (e.g. Fig. 2.3).
Alteration textures can aid determination of equilibrium
mineral assemblages, alteration intensity, and overprinting
relationships. Alteration facies distribution and zonation
patterns can be used to interpret patterns of fluid flow, changes
in physicochemical conditions and development of alteration
systems. Superimposed alteration patterns and overprinting
textures are important for determining paragenesis involving
multiple stages of alteration and hence for understanding
evolution of the system over time.
3.1 | ALTERATION TEXTURES
Typically alteration encompasses mineralogical and textural
changes. Textural changes are changes in the shape, form,
grainsize and orientation of grains within the rock and can be
texturally destructive, preserve relicts of pre-existing textures
or enhance textures (McPhie et al, 1993; Doyle, 2001).
Alteration textures are those that are superimposed on the rock
by the processes of alteration (i.e. by hydration, dissolution,
diagenesis, hydrothermal alteration, metamorphism and
deformation).
Changes in texture during alteration may involve: the
precipitation of minerals along fluid pathways; creation
or infilling of pore space; the dissolution and replacement
of earlier minerals and glass by subsequent minerals; and
recrystallisation. There are five common types of alteration
textures that occur in volcanic facies (Tables 3.1 and 3.2):
(1) replacement textures, (2) infill textures, (3) dissolution
textures, (4) recrystallisation textures, and (5) deformation
textures. In addition, the combined effects of a number of
different overprinting alteration facies can result in false or
pseudotextures (De Rosen-Spence et al., 1980; Allen, 1988).
Furthermore, there are two types of textures that are
common in and unique to volcanic rocks, which, although
not alteration textures, influence subsequent alteration,
especially the development of pseudotextures. These are high-
temperature devitrification textures (i.e. spherulites, varioles,
lithophysae and micropoikilitic texture) and perlitic fractures
(Figs 3.1 and 3.2). The formation and alteration of perlite is
described in detail in Section 5.2.
Replacement textures
Most alteration forms by replacement, because pre-existing
mineral phases and glass become unstable during changed
geothermal conditions and are readily substituted by new,
more stable minerals. Replacement is the process of practically
simultaneous solution and deposition of a new mineral of
partly or completely different composition either in a pre-
existing mineral or an aggregate of minerals (Lindgren, 1933).
Although mineral exchange is essentially simultaneous,
replacement may occur in stages, where intermediate
products form, at least temporarily, before the final alteration
TABLE 3.1 | Types of textural changes that occur during alteration.
Replacement
(metasomatism)
Infill
Dissolution
Static
recrystallisation
Dynamic
recrystallisation
Deformation
Existing minerals or glass are replaced by one
or more new mineral species
A mineral or minerals are precipitated from
solution into open space
Existing minerals or glass are leached
and removed by solution with or without
replacement
Recrystallisation of existing minerals to new
grains, and/or a change in morphology of the
same mineral species or composition
Recrystallisation of existing minerals to new
grains and/or a change in morphology and/or
orientation of the same mineral species or
composition
Existing component or texture is rotated,
milled, broken, compressed, modified,
distorted or fractured
38 | CHAPTER 3
TABLE 3.2 | Common macroscopic and microscopic alteration textures in volcanic rocks.
minerals. For example, relict radiating fibrous textures locally
preserved in feldspar-altered pumice and perlite clasts in the
Mount Read Volcanics, western Tasmania, suggest that an
intermediate phase between felsic glass and feldspar, possibly
fibrous zeolites, occurred (Fig. 5.11: Gifkins and Allen,
2001).
Replacement can range from the conversion of specific
mineral phases or domains to new minerals (selective
alteration, Fig. 3.3B, C and D), to complete replacement of
a rock to a completely new mineral assemblage (pervasive
alteration, Fig. 3.3A). Where alteration occurs dominantly
by diffusion, it may affect a large volume of rock. Elsewhere
it may occur along well-defined fluid pathways (vein-halo
alteration, Fig. 3.3E) with its effects restricted to a scale of
millimetres to metres (Titley, 1994). It is worth noting that
the terms pervasive, selective or vein-halo depend on the scale
of observation. For example, vein-halo alteration can appear
pervasive when viewed in thin section.
Pervasive
Pervasive alteration is extensive alteration that has completely
changed the rock composition and texture at scales that range
from millimetres to kilometres (Rose and Burt, 1979; Titley,
1982). Pervasive alteration is distributed without regard for
pre-existing textures and can result in disseminated, massive
microcrystalline or cryptocrystalline microscopic textures
(Fig. 3.4A).
Selective
Selective alteration converts only specific pre-existing phases
to new mineral phases (Titley et al., 1978; Titley, 1982). The
original rock texture may be only slightly modified during
selective alteration because only certain components in the
host (e.g. minerals, volcanic glass or clasts: Fig. 3.3B, C and
D) are preferentially replaced, and others are left relatively
unaltered (Rose and Burt, 1979).
Replacement Pervasive Pervasive, selective, massive, disseminated, microcrystalline,
cryptocrystalline
Selective Disseminated Disseminated, pseudomorph, overgrowth, cleavage and rim
texture, core and zonal replacement texture, microcrystalline,
cryptocrystalline, spheroid, nodule, concretion
Domainal
Vein halo Pervasive, selective, disseminated, pseudomorph, overgrowth,
cleavage and rim texture, core and zonal replacement texture,
microcrystalline, cryptocrystalline, spheroid, nodule, concretion
Infill Incomplete infill Crustiform, fibrous, prismatic, spherulitic
Massive infill Microcrystalline, prismatic
Layered or banded infill Crustiform, colloform, comb, botryoidal
Dissolution Stylolitic foliation Stylolites, solution seams
Corrosion vug Open pore space infill textures (prismatic, fibrous and
massive)
Static Pervasive (hornfels) Equigranular, granoblastic, granophyric, decussate
recrystallisation Seledive Porphyroblastic, idioblastic, xenoblastic, poikiloblastic,
intergrowths, overgrowths, reaction rims, polycrystalline grains
Dynamic Foliation Slaty cleavage Cleavage, mineral alignment, granoblastic, porphyroblastic,
recrystallisation poikiloblastic
Schistosity
Layering (gneissosity) Differential layering, microcrystalline, granoblastic, granophyric
Lineation Aligned, strained, bent, kinked, flattened, twinned and broken
grains (crystals or clasts), cleavage
Cataclasite No foliation, porphyroblastic, microcrystalline
Mylonite Foliation, granular
Deformation Foliation Cleavage, aligned, strained, bent, kinked, flattened, twinned and
broken grains (crystals or clasts), fiamme, eutaxitic
Lineation Aligned, strained, bent, kinked, flattened, twinned and broken
grains (crystals or clasts), cleavage
Augen structure Cleavage
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 3 9
A. Spherulites and obsidian in rhyolite
Pink, isolated spherulites and densely microspherulitic
flow bands are enclosed in black obsidian in this flow-
banded rhyolite. Spherulites are radiating aggregates
or bundles of acicular and fibrous crystals. They vary
in shape from spherical to bow-tie shaped sheafs and
axiolitic bundles, and are commonly composed of
feldspar or intergrowths of alkali feldspar, plagioclase,
cristobalite or tridymite and clinopyroxene (Lofgren,
1971b). Spherulites are typically the product of high-
temperature (above the glass-transition temperature)
devitrification of silicic glass (Lofgren, 1971a).
Sample NG4, recent Ngongotaha lava dome, Hendersons
quarry, Rotorua, New Zealand.
B. Lithophysae in rhyolite
This red albite + quartz + hematite-altered, flow-banded
quartz + plagioclase-phyric rhyolite contains abundant
spherulites and lithophysae. The lithophysae are filled
with layered quartz.
Sample from the Lower Devonian Snowy River Volcanics,
Flukes Knob area, Victoria.
C. Varioles in basalt
Dark spots in this basalt outcrop are varioles: radial or
sheaf-like aggregates of plagioclase and pyroxene, olivine
or iron oxides, and are similar to spherulites, but only
occur in mafic facies (cf. Fowler et al., 1987; Williams
etal., 1982).
Shirakawa quarry, Miocene Green Tuff Belt, Odate,
Japan.
D. Micropoikilitic texture in thin section
The groundmass of this rhyolite is densely micropoikilitic;
comprising patches of optically continuous quartz,
which enclose variably oriented laths of sericitised albite.
Poikilitic and micropoikilitic texture (snowflake texture)
comprise an optically continuous crystal enclosing
numerous randomly oriented inclusions of a different
composition (Anderson, 1969). The boundaries between
the micropoikilitic quartz domains in this sample are
highlighted by concentrations of sericite. Cross polarised
light.
Sample 133921, Cambrian Mount Black Formation,
Central Volcanic Complex, Mount Read Volcanics, Mount
Black, western Tasmania.
FIGURE 3.1 | Examples of high-temperature devitrification textures.
40 | CHAPTER 3
A. Altered macroperlite
Relict macroperlitic factures in this coherent dacite are
enhanced by dark grey sericite + chlorite-altered zones
along and adjacent to the perlitic fractures. The arcuate
shape of the fractures is preserved in some areas. The
perlite cores are pink albite + quartz + sericite altered.
Cambrian Mount Black Formation, Central Volcanics
Complex, Mount Read Volcanics, Pieman Road, western
Tasmania.
B. Relict perlite in thin section
The formerly glassy groundmass of this rhyolite preserves
perlitic fractures. Perlitic fractures are a network of fine
typically concentric, arcuate fractures that enclose glassy
or originally glassy cores. Here, the perlitic fractures are
filled with dark, mixed-layer smectite/chlorite and the
groundmass adjacent to the fractures is clinoptilolite
altered. The perlitic cores are partly glassy and partly
smectite altered. Plane polarised light.
Sample J6-737 m, Miocene Nishikurosawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
C. Banded perlite
This finely flow-banded, plagioclase-phyric rhyolite
contains an intersecting fracture network of sub-parallel
long fractures linked by short cross fractures (banded
perlite) superimposed on the flow-banded texture.
Sample KB257, Siluro-Devonian rhyolite, Ural Volcanics,
Ural Ridges area, New South Wales.
D. Banded perlite in thin section
In thin section, concentrations of sericite hematite
mark the relict perlitic fractures. The pale flow bands
comprise a fine-grained mosaic of feldspar + quartz,
whereas the darker bands consist of sericite + feldspar
+ quartz + chlorite. Disseminated fine-grained hematite
occurs throughout the groundmass. Plane polarised
light.
Sample KB257, Siluro-Devonian rhyolite, Ural Volcanics,
Ural Ridges area, New South Wales.
FIGURE 3.2 | Examples of perlite.
Two textural types of selective alteration occur: disseminated
alteration (or selective-pervasive alteration), which refers to
the replacement of selective pre-existing phases throughout
the host rock; and domainal alteration, which refers to the
alteration of patches, pods, or groups of clasts within the host
rock (Fig. 3.3F, G and H). In addition, selective alteration may
result in concentrically zoned alteration facies within clasts or
alteration halos around clasts (Fig. 3.31, J, K and L). Selective
alteration can result in a patchy or mottled appearance (e.g.
Allen, 1988).
Common microscopic selective replacement textures
are pseudomorphs, partial pseudomorphs (cleavage and rim
texture, core and zonal texture, core and rim texture and
skeletal texture), overgrowths on pre-existing components,
and spheroids or nodules (Fig. 3.4B to L: Dimroth and
Lichtblau, 1979; Craig and Vaughan, 1981; Ineson, 1989).
Carbonate and zeolite nodules are common in submarine
volcaniclastic facies and can have a wide variety of grainsizes
from 0.2 to greater than 20 mm (Fig. 3.41 to L: Franklin et
al., 1975; Lees, 1987; Khin Zaw and Large, 1992; Hill and
Orth, 1994; Allen, 1997).
Vein halo
Vein-halo alteration involves the replacement of either the
whole rock (pervasive alteration: Fig. 3.3E) or specific pre-
existing phases (selective alteration) in restricted areas, such as
the halos around veins, intrusion contacts, or at stratigraphic
contacts. Alteration progresses in fronts, moving out from
fractures or contacts into the adjacent wall rock. Vein-halo
alteration has also been termed infiltration metasomatism,
vein-veinlet, reaction rim, vein-wall-rock, vein-envelope,
veinlet-controlled and fracture-controlled alteration (e.g.
Titley et al., 1978; Titley, 1982; Thompson and Thompson,
1996; Doyle, 2001).
Infill textures
Infill or open space-filling textures result from the precipitation
of new mineral phases from solution into open spaces or
cavities such as pore spaces, vesicles, inter-clast space, vugs
and fractures (Taylor, 1992). Infill textures are characterised
by well-developed crystal faces, zoned crystals and mineral
banding (Craig and Vaughan, 1981). Silicate, carbonates,
oxides, sulfates and sulfides all occur as void fill in altered
volcanic rocks.
Infill results from the precipitation of minerals from
solution. The first mineral to be deposited forms a crust
on the cavity walls and grows inwards, generally with the
development of inward facing crystal faces. Common infill
textures include incomplete infill, massive infill, and layered
or banded infill (Fig. 3.6: Taylor, 1992). These textures can
include fibrous, prismatic, spherulitic or equant crystal shapes
and exist on a range of scales from micrometres to metres
(Dimroth and Lichtblau, 1979; Taylor, 1992).
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 4 1
Incomplete infill
Incomplete or partial infilling of veins and cavities or
dissolution of void fill can leave an open vug in the centre (e.g.
Fig.3.5A). In many cases, the resulting infill texture contains
well-formed crystals that project into this vug.
Massive infill
Massive infill textures result from the continuous deposition
of a mineral or aggregate of minerals until the cavity is filled
(e.g. Fig. 3.5B). Massive infill commonly contains well-formed
crystals, especially quartz, feldspar, fluorite, cassiterite, galena,
sphalerite and chalcopyrite crystals. Massive, microcrystalline
forms also exist (Taylor, 1992).
Layered infill
Layered or banded infill textures result from the deposition of
a succession of minerals inwards from the cavity or fracture
wall (Bateman, 1951). Layered infill textures do not generally
contain well-formed crystals, such as comb texture (e.g. Fig.
3.5H), but vary from thin layers of individual minerals to
crustiform bands or colloform textures.
Dissolution textures
Dissolution textures are common in altered volcanic rocks
(Allen, 1990; Allen and Cas, 1990; Marsaglia and Tazaki,
1992; Gifkins and Allen, 2001). They form from the corrosion
or leaching of pre-existing phases (either glass or mineral
phases), with or without minor replacement by new mineral
phases (Fisher and Schmincke, 1984). For example, leaching
of rhyolitic glass is commonly accompanied by crystallisation
of muscovite or clay minerals that absorb leached ions from
solution (Karkhanis et al., 1980).
Dissolution textures include corrosion vugs or dissolution
pits, stylolites and solution seams (e.g. Fig. 3.6: Pettijohn,
1957).
Corrosion vugs
Dissolution or corrosion of volcanic glass or pre-existing
minerals can create open cavities or oversized pores (Fig. 3.6A
to F) in which infill can occur synchronous with dissolution or
after dissolution (Hay, 1963; Sheppard et al., 1988). In some
cases pseudomorphs of minerals or originally glassy particles,
such as glass shards, form by dissolution and precipitation
(Riech, 1979; Sheppard etal., 1988). Riech (1979) recognised
infill textured zeolites and calcite within clinopyroxenes,
and proposed that clinopyroxenes were corroded during
diagenesis, creating an open void that was subsequently filled
with zeolites and calcite. Similarly, Hay (1963) recognised
partial to complete dissolution of glass shards followed by the
precipitation of authigenic minerals, especially zeolites, in the
new cavities as well as in original pore space.
42 | CHAPTER 3
A. Pervasively altered rhyolite
Intense, pervasive, fine-grained K-feldspar + quartz
alteration has completely replaced the groundmass and
plagioclase phenocrysts in this rhyolite.
Sample 143286, Central Volcanic Complex, Mount Read
Volcanics, Mount Darwin, western Tasmania.
B. Selectively altered phenocrysts
Sericite has selectively altered the coarse prismatic
feldspar phenocrysts (F) in this latite. The pale green-
grey, fine-grained groundmass is moderately and
pervasively phengite + chlorite + ankerite altered, and
the amygdales are quartz filled.
Sample 144369, Ordovician Lake Cowal Volcanics, Junee-
Narromine Volcanic Belt, Endeavour 42 prospect, New
South Wales.
C. Selectively altered pumice clasts
Large pumice clasts (P) in this sample of crystal- and
pumice-rich volcaniclastic breccia have been selectively
altered to orange albite + quartz, whereas the finer
grained matrix has been altered to green sericite +
chlorite + albite. The domainal alteration style enhances
its clastic appearance.
Sample 131993, Cambrian Mount Julia Member,
Tyndall Group, Mount Read Volcanics, Comstock, western
Tasmania.
D. Selectively altered matrix
In this andesitic volcaniclastic breccia, the matrix is
moderately and selectively epidote altered. In contrast,
the plagioclase-phyric clasts (C) are weakly chlorite +
sericite altered.
Sample 144805, Ordovician Mingelo Volcanics, Junee-
Narromine Volcanic Belt, Peak Hill, New South Wales.
FIGURE 3.3 | Examples of replacement textures in hand specimen.
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS I 4 3
E. Vein halo
Red albite altered zones are restricted to 5 mm halos or
selvages adjacent to quartz + actinolite + pyrite veinlets
in this feldspar porphyritic dacite.
Sample TH386 271.1 m, Cambro-Ordovician Trooper
Creek Formation, Seventy Mile Range Group, Mount
Windsor Subprovince, Thalanga, Queensland.
F. Banded, domainal alteration facies
Diffuse and discontinuous pink and green bands in this
massive crystal-rich volcaniclastic sandstone are defined
by domains of albite + quartz chlorite, and chlorite
+ sericite + magnetite alteration facies, respectively. The
bands are not obviously consistent with grainsize or
component variations; they alternate on a 210 cm scale,
are laterally extensive (1020 m) and are commonly, but
not exclusively, bedding parallel.
Sample 131982, Cambrian Mount Julia Member, Tyndall
Group, Mount Read Volcanics, Lyell Comstock, western
Tasmania.
G. Patchy, domainal alteration facies
The domainal, green epidote + quartz and grey albite
+ quartz + hematite alteration facies are distributed in
irregular patches with diffuse margins in this coherent
plagioclase-phyric dacite.
Sample M142, Cambrian Mount Black Formation,
Central Volcanic Complex, Mount Read Volcanics, Tullah,
western Tasmania.
H. Domainal alteration facies in pseudobreccia
Domainal red albite + quartz and dark green epidote +
sericite + albite alteration facies in this sample of macro-
perlite gives it a pseudo-polymictic and -clastic texture.
However, the red apparent clasts have diffuse margins
and identical phenocryst populations to the apparent
matrix.
Sample 147550, Cambrian Mount Black Formation,
CentralVolcanics Complex, Mount Read Volcanics, Pieman
Road, western Tasmania.
FIGURE 3.3 | Examples of replacement textures in hand specimen, cont.
44 | CHAPTER 3
I. Zonation within clasts
The andesite and basalt clasts in this polymictic volcanic
breccia are concentrically zoned, with sericite + quartz
+ calcite-altered rims, and chlorite-altered cores. Some
of the larger clasts have an additional quartz + sericite
+ chlorite-altered core zone. The matrix has been
moderately and pervasively quartz + sericite + calcite
chlorite altered.
Cambrian Que-Hellyer Volcanics, Mount Charter Group,
western volcano-sedimentary sequences, Mount Read
Volcanics, Hellyer, western Tasmania.
J. Zonation within clasts
Clasts in this basaltic pebble conglomerate display
heterogenous alteration facies, and some clasts are
internally zoned. The basalt clast (B) has a fine-grained,
pale green sericite-rich rim, and darker sericite + chlorite-
altered core.
Sample 134632, Cambrian Red Lead Formation correlate,
Dundas, Kapi Creek, western Tasmania.
K. Zonation within pillows
This metamorphosed, amphibolite-grade lava-pillow has
a typical triangular, draped shape and is concentrically
zoned. The central red zone is coarse-grained, scapolite-
poor and albite + hematite + sericite epidote altered.
The average grainsize decreases, and scapolite grainsize
and abundance increases, in consecutive zones towards
the rim. Biotite + calcite + hornblende + microcline +
scapolite + epidote + quartz comprise the inter-pillow
matrix.
Proterozoic Corella Formation, Mary Kathleen Group,
Malbon River, northwest Queensland.
L. Altered halos around clasts
Orange albite + quartz alteration facies is distributed in
a halo around a massive, albite-altered dacite clast (C) in
this crystal- and lithic-rich volcaniclastic sandstone. The
more pervasive green-grey domain is sericite + chlorite +
quartz + albite altered.
Sample 132090, Cambrian Mount Julia Member, Tyndall
Group, Mount Read Volcanics, Anthony Road, western
Tasmania.
FIGURE 3.3 | Examples of replacement textures in hand specimen, cont.
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 4 5
A. Microcrystalline texture in thin section
The groundmass of this rhyolite is a microcrystalline
mosaic of quartz + feldspar + sericite. Quartz phenocrysts
(Q) have been recrystllised. Microcrystalline texture
(aphanitic) is a fine-grained granular texture where the
individual crystals can be distinguished in thin section.
In contrast, cryptocrystalline texture (phaneritic) is
where the crystals are too minute to be distinguished
even with the aid of a microscope (Williams et al.,
1982). Cross polarised light.
Sample 133318, Cambro-Ordovician Mount Windsor
Formation, Seventy Mile Range Group, Thalanga,
Queensland.
B. Pseudomorphs in thin section
The plagioclase phenocrysts in this sericite + quartz
+ tourmaline-altered andesite were pseudomorphed
by tourmaline, and subsequently almost completely
replaced by blue-green chlorite. Pseudomorphs are
crystals or aggregates of crystals that preserve the shape
of a pre-existing mineral or particle (e.g. glass shard or
pumice clast) (Spry, 1976). Plane polarised light.
Sample 145199, Ordovician Forest Reefs Volcanics, Molong
Volcanic Belt, Black Rock, New South Wales.
C. Pseudomorphs in thin section
This thin section of plagioclase + quartz + pyroxene-
phyric rhyolite shows an illite pseudomorph after
pyroxene. Plagioclase phenocrysts have been altered to
K-feldspar and the groundmass comprises a fine-grained
mosaic of K-feldspar + quartz + chlorite + smectite.
Cross polarised light.
Sample KB495, Siluro-Devonian Coan rhyolite, Mount
Hope Volcanics, Coan Gonn Peak, New South Wales.
D. Cleavage and rim texture in thin section
The plagioclase crystals in this basalt have been selectively
altered by sericite along cleavage planes. Cleavage and
rim textures occur by selective alteration of mineral grain
boundaries and cleavages. It is common in plagioclase,
in which montmorillonite, sericite or calcite form along
the cleavage planes (Sales and Meyer, 1948). Plane
polarised light.
Sample SVD87a-104.9 m from the Cambrian Sterling
Valley Volcanics, Mount Read Volcanics, Sterling Valley,
western Tasmania.
FIGURE 3.4 | Examples of replacement textures in altered volcanic rocks.
4 6 I CHAPTER 3
E. Core and zonal texture in thin section
Zones within plagioclase phenocrysts in this subtly,
smectite + calcite-altered diorite have been selectively
altered to sericite. These incomplete pseudomorphs,
termed core and zonal texture, are particularly common
in zoned feldspar, amphibole and mica crystals where
the cores, or one or more zones in zoned minerals,
are altered (Barker, 1990). In plagioclase crystals, like
those pictured here, the calcic zones are typically altered
to calcite or sericite (Sales and Meyer, 1948). Plane
polarised light.
Sample 152958, Pliocene-Pleistocene Luise Volcano, Lihir
Island, New Ireland Province, Ladolam epithermal Au
mine, Papua New Guinea.
F. Core and zonal texture in thin section
In this example of core and zonal texture, the core zones
of plagioclase phenocrysts have been altered to sericite.
The groundmass of this plagioclase + clinopyroxene-
phyric basalt was subtly and pervasively smectite +
calcite-altered. Plane polarised light.
Sample 152830, Pliocene-Pleistocene Luise Volcano, Lihir
Island, New Ireland Province, Ladolam epithermal Au
mine, Papua New Guinea.
G. Overgrowth texture in thin section
A discontinuous K-feldspar overgrowth encloses a
hematite-altered plagioclase phenocryst (P) in this
strongly and pervasively albite + quartz + sericite-altered
pumice breccia. K-feldspar nucleated on the plagioclase
phenocryst, spread outwards filling vesicles, and replaced
vesicle walls in the pumice clasts. Overgrowth textures
are mineral rims that may be composed of one or more
crystals of similar or different minerals. Cross polarised
light.
Sample 133814 from the Cambrian Hercules Pumice
Formation, Central Volcanic Complex, Mount Read
Volcanics, Hercules footwall, western Tasmania.
H. Altered nodules in pumice breccia
Blue-green celadonite nodules have overprinted pumice
clasts in this polymictic volcanic breccia. These nodules
are composed of fine-grained aggregates of celadonite
opal CT quartz, preserve uncompacted tube and round
vesicle pumice textures, and are surrounded by pervasive
smectite + mordenite + calcite alteration facies.
Sample FK2, Miocene Byobu-iwa Member, Tokiwa
Formation, South Fossa Magna, Green Tuff Belt, Fujikawa
River, Japan.
FIGURE 3.4 | Examples of replacement textures in altered volcanic rocks, cont.
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS I 47
I. Carbonate spheroids
Large dolomite spheroids are enclosed in the strongly
chlorite + quartz + dolomite-altered matrix of this
formerly plagioclase-phyric andesite. Nodules and
spheroids are spherical domains of alteration, which
may comprise radiating aggregates of fibrous crystals,
fine internally concentric structures, or mosaics of
anhedral grains, with or without cores (Allen, 1997; Hill
andOrth, 1995).
Sample 135756, Cambrian Que-Hellyer Volcanics,
Mount Charter Group, western volcano-sedimentary
sequences, Mount Read Volcanics, Hellyer footwall, western
Tasmania.
J. Carbonate spheroids in thin section
In thin section, the dolomite spheroids display
concentric zones and a coarse, radiating, fibrous texture.
This compositional zonation in the spheroids probably
indicates multiple stages of carbonate alteration (cf. Hill
and Orth, 1995). Plane polarised light.
Sample 135756, Cambrian Que-Hellyer Volcanics,
Mount Charter Group, western volcano-sedimentary
sequences, Mount Read Volcanics, Hellyer footwall, western
Tasmania.
K. Carbonate spheroids
Carbonate spheroids are concentrated in individual beds
in this strongly chlorite + carbonate + pyrite-altered
laminated volcaniclastic sandstone. The larger spheroids,
which are up to 2 mm in diameter, have coalesced.
Sample 138601, Archaean Mb5 Golden Grove Formation,
Luke Creek Group, Murchison Volcanics, Golden Grove,
Western Australia.
L. Carbonate spheroids in thin section
Thin section examination shows these carbonate
spheroids are supported in a fine-grained quartz +
sericite + carbonate matrix. Plane polarised light.
Sample 138601, Archaean Mb5 Golden Grove Formation,
Luke Creek Group, Murchison Volcanics, Golden Grove,
Western Australia.
F IG U R E 3. 4 | Examples of replacement textures in altered volcanic rocks, cont.
4 8 | CHAPTER 3
A. Incomplete infill in fractures
Incomplete filling of fractures in this altered diorite left
sub-planar vugs. The pyrite fill has botryoidal surfaces,
representing rounded shapes of either spherulitic
radiating aggregates of fibrous crystals or fine-concentric
internal structures (cf. Jensen and Bateman, 1981). This
massive plagioclase-phyric diorite has been pervasively
K-feldspar + pyrite (>quartz + illite) altered, and
dissolution of primary mafic minerals produced a fine,
spongy, porous texture.
Sample 152959, Pliocene-Pleistocene Luise Volcano, Lihir
Lsland, New Ireland Province, Ladolam epithermal Au
mine, Papua New Guinea.
B. Massive infill
Pale green epidote altered halos surround massive
chlorite-filled amygdales (A) in this basalt sample.
Sample 144753, Ordovician, Junee-Narromine Volcanic
Belt, Boundary Prospect, Lake Cowal, New South Wales.
C. Layered infill stringer vein
This banded vein consists of successive layers, from the
vein wall to centre, of quartz, quartz and intergrown
chalcopyrite and pyrite, and dolomite. A thin dolomite
vein has overprinted the stringer vein at an oblique
angle. These veins are hosted in strongly and pervasively
sericite + chlorite + albite + pyrite-altered andesite.
Cambrian Que-Hellyer Volcanics, western volcano-
sedimentary sequences, Mount Read Volcanics, Hellyer
footwall, western Tasmania.
D. Layered infill in amygdales
Amygdales (A) in this basalt clast, from a basalt-
mudstone peperite, contain concentric layers of quartz
and calcite, which have grown inwards from the vesicle
walls. The basalt groundmass has been pervasively
sericite + chlorite + calcite altered. The clast grainsize
decreases towards the clast rim, to the left of the field of
view in this photograph.
Sample 76836, Cambrian Que-Hellyer Volcanics, western
volcano-sedimentary sequences, Mount Read Volcanics,
Hellyer, western Tasmania.
FIGURE 3.5 | Examples of infill textures in altered volcanic rocks.
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 4 9
E. Layered infill texture in vesicles
Vesicles (V) in this plagioclase-phyric pumice clast have
been filled with roughly concentric layers of tan-coloured
mordenite, dark smectite and clear clinoptilolite. The
zeolites occur in clusters or aggregates of fine, radiating
fibres. The originally glassy vesicle walls (W) have been
replaced by mordenite + K-feldspar smectite. Plane
polarised light.
Sample OH8-387 m, Miocene Onnagawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
F. Layered infill texture in amygdales
Amygdales in this subtly altered perlitic rhyolite have
been filled with bands of fine-grained montmorillonite
and unknown radiating fibrous minerals. Pale polarised
light.
Sample J6-737 m, Miocene Nishikurosawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
G. Layered infill in amygdales
The amygdales in this palagonite-altered trachytic basalt
clast from a crystal- and lithic-rich pumice breccia are
filled with layers of montmorillonite and fibrous zeolites.
Plane polarised light.
Sample OH8-387 m, Miocene Onnagawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
H. Comb texture
This example of comb texture shows layers of prominent
sparry quartz + amethyst carbonate crystals projecting
inwards from the vein or cavity wall.
Sample T5> Cretaceous, andesite, Fresnillo epithermal
district, Mexico.
FIGURE 3.5 | Examples of infill textures in altered volcanic rocks, cont.
50 | CHAPTER 3
A. Dissolution vugs
This hand specimen of polymictic breccia has a spongy
or vuggy porous texture due to the dissolution of primary
mafic igneous minerals and glass. It has been intensely
and pervasively adularia + illite + pyrite altered with illite
replacing plagioclase crystals, secondary K-feldspar in
the altered matrix, and disseminated pyrite.
Sample 152726, Pliocene-Pleistocene Luise Volcano, Lihir
Island, New Ireland Province, Ladolam epithermal Au
mine, Papua New Guinea.
B. Dissolution vugs in thin section
In thin section, irregularly shaped, empty, corrosion
or dissolution vugs (V) are conspicuous in the matrix
and clasts. Some vugs cut across clast margins. Plane
polarised light.
Sample 152726, Pliocene-Pleistocene Luise Volcano, Lihir
Island, New Lreland Province, Ladolam epithermal Au
mine, Papua New Guinea.
C. Filled dissolution vug in thin section
Corrosion vugs, created by the dissolution of volcanic
glass or pre-existing minerals, are commonly filled
by subsequent mineral precipitation from solution.
Successive layers of montmorillonite and zeolite have
filled an irregular vug (V) in this thin section. The vug
occurs in the matrix and in a basalt clast, crossing the
clast-matrix contact. Plane polarised light.
Sample OH8-387 m, Miocene Onnagawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
D. Vuggy quartz
The prominent features in this quartz-rich sample are the
corrosion vugs, which were generated by the dissolution
of pumice clasts and crystals from this pumice and lithic
tuff.
Miocene rhyodacitic pumice and lithic tuff, Pierina Au-Ag
deposit, Peru.
FIGURE 3.6 | Examples of dissolution textures in altered volcanic rocks.
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS I 51
E. Kaolinite + dickite-altered andesite
Large (up to 4 mm), blocky feldspar phenocrysts have
been kaolinite altered in this sample of massive, coherent
andesite.
Miocene andesitic lava, Pierina Au-Ag deposit, Peru.
F. Vuggy quartz
In this sample of vuggy quartz, which is equivalent to
the previous kaolinite + dickite-altered andesite, the
feldspar phenocrysts have been dissolved resulting in
blocky vugs. The groundmass is composed dominantly
of quartz.
Miocene andesitic lava, Pierina Au-Ag deposit, Peru.
G. Stylolite in thin section
Stylolites (S
2
) in this rhyolitic pumice breccia have
concentrated fine-grained opaques and sericite
chlorite. The stylolites define the compaction foliation
and are crenulated by the dominant regional cleavage
{S
2
) defined by alignment of sericite in the subtly albite
+ quartz + sericite-altered matrix.
Sample 147422, Cambrian Kershaw Pumice Formation,
Central Volcanics Complex, Mount Read Volcanics,
Rosebery, western Tasmania.
E. Solution seams in thin section
These analcime-filled solution seams occur in a smectite-
rich fiamme, extending from the damme terminations
into the shard- and crystal-rich matrix of a crystal-rich
pumice breccia. They are interpreted to have formed
by dissolution and precipitation under the influence of
lithostatic load during diagenesis. Plane polarised light.
Sample FK7, Miocene Wadaira Tuff Member, Tokiwa
Formation, South Fossa Magna, Green Tuff Belt, Wadaira,
Japan.
FIGURE 3.6 | Examples of dissolution t ext ures in alt ered volcanic rocks, cont.
5 2 | CHAPTER 3
Vuggy silica (quartz) alteration facies is characterised by
fine-grained, microcrystalline quartz and abundant open vugs
or pores, which may be partly infilled (e.g. Fig. 3.6D, E and
F). It is common in high-sulfidation epithermal systems and
results from the extensive leaching of all phases, except SiO
2
and TiO
2
, from volcanic rocks by hot acid solutions (White
and Hedenquist, 1990).
Stylolites
Stylolites are common in altered volcaniclastic rocks (Allen,
1990; Allen and Cas, 1990; Marsaglia and Tazaki, 1992;
Gifkins and Allen, 2001). They are surfaces of dissolution
associated with strain (pressure solution). They are roughly
planar surfaces that exhibit mutual column and socket inter-
digitation and may branch. Stylolites result from mechanical
compaction and removal of elements by diffusion and
precipitation (Merino et al., 1983). They indicate volume
loss and may form parallel or sub-parallel to bedding during
burial, or at high angles to bedding during folding. Stylolites
often contain a residue of insoluble material and minerals
precipitated from solution. RecrystaUisation, dissolution,
grain growth, grain orientation, pressure twinning, fracturing
and residual accumulation of minerals along stylolites are
common (Amstutz and Park, 1967).
Irregular, anastomosing, bedding-parallel stylolites have
been recognised in originally glassy volcanic facies, especially
pumice breccias, in the Mount Read Volcanics (Allen, 1990;
Allen and Cas, 1990; Gifkins and Allen, 2001). These are
seams that concentrate fine-grained opaques and sericite at
the margins of originally glassy clasts, along tube vesicle walls
in pumice clasts and in the matrix (Fig. 3.6D: Gifkins, 2001).
These stylolites are interpreted as diagenetic compaction and
dissolution fabrics that formed by the dissolution of soluble
components, particularly glass, and by the precipitation of
clays and Fe-oxides as a result of pressure during burial (Allen,
1990; Allen and Cas, 1990; Gifkins, 2001).
Solution seams
Solution seams are non-sutured, discontinuous mineral-
filled seams that may form during diagenesis as a result of
stress-related dissolution of soluble components and re-
precipitation (Merino et al., 1983). Analcime-filled solution
seams in pumice-rich rocks from the Green Tuff Belt (Japan)
are anastomosing and roughly parallel to bedding. They occur
in the fine-grained matrix and within fiamme and partly
compacted pumice clasts (Fig. 3.6E: Gifkins et al., in press).
Static recrystaUisation textures
RecrystaUisation is the transformation of a mineral or glass
to a new grainsize, morphology or orientation of the same
mineral species or minerals of the same composition (i.e.
neomorphism, Folk, 1965). Pre-existing minerals recrystallise
to a new grainsize in an attempt to assume a more stable form
by minimising the ratio of the surface area to the volume
during changed physical conditions (Yardley, 1989).
RecrystaUisation textures are produced by changes in
the size, shape and arrangement of minerals in a rock. With
increasing temperature, recrystaUisation generally involves the
change from fine to coarse grainsize (aggrading), except for
static recrystaUisation to hornfels and dynamic recrystaUisation
where large, strained grains are replaced by a mosaic of tiny,
unstrained crystals (Folk, 1965). Minerals may be directed
or randomly orientated (non-directed: Spry, 1976). Directed
textures occur where recrystaUisation is accompanied by stress
(dynamic recrystaUisation). Non-directed textures occur where
the pressure is equal in all directions (static recrystaUisation).
Common macroscopic and microscopic recrystaUisation
textures include mineral overgrowths, porphyroblasts,
poikoblasts, and hornfelsic, granoblastic, granophyric and
decussate textures (Fig. 3.7).
Dynamic recrystaUisation textures
Directed fabrics or textures are common to regional
metamorphic rocks where recrystaUisation is accompanied
by stress (dynamic recrystaUisation). These textures include
subgrains (granoblastic, porphyroblastic and poikiloblastic
textures), foliations, layering and lineations.
In common usage the term foliation is non-genetic and
describes any planar, spaced or pervasive fabric in the rock.
Foliations may form during diagenesis, metamorphism
or tectonic deformation. Planar foliations are due to the
preferred orientation of minerals, particularly micas, aligned
perpendicular to the maximum compression direction
(Yardley, 1989). Planar foliations are subdivided on the basis
of grainsize and overall appearance of the altered rock. These
include slaty cleavage, schistosity and gnessic layers.
In fault zones or zones of intense ductile shear, two
characteristic textures occur: cataclastic and mylonitic
textures. Cataclase refers to a fine-grained fault gouge breccia
with an unfoliated matrix (Sibson, 1977). Ideally, cataclasis is
mechanical fragmentation without recrystaUisation, however
this rarely occurs in nature (cf. Sibson, 1977). Mylonite is a
term used for strongly foliated fine-grained rocks in which
the grainsize has been reduced by recrystaUisation (Bell and
Etheridge, 1973).
Deformation textures
Deformation textures are stress activated and develop in
response to overburden pressures or regional tectonic stress.
Because volcanic deposits typically have high initial porosities
they are easily modified by mechanical compaction during
burial, tectonic deformation or, in the case of pyroclastic
deposits, welding (Peterson, 1979; Allen, 1988; Branney
and Sparks, 1990). Textural modification associated with
compaction is essentially a result of increased pressure causing
the re-arrangement and deformation of grains and reduction
of intergranular pore space (Deelman, 1975).
Deformation textures result from the rotation, brittle
fracturing, flattening and distortion of existing grains or
fabrics, especially clasts that were previously altered to soft
minerals (McBride, 1978; Galloway, 1979; Craig and
Vaughan, 1981; Branney and Sparks, 1990).
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 5 3
A. Porphyroblasts
The large, strongly altered, cordierite porphyroblasts give
this rhyolite a distinctive coarsely spotty texture. This
texture inspired the terms dalmatianite, which was applied
by early workers in the Noranda Camp, and the spotted
fades, which was applied by Riverin and Hodgson (1980),
for the cordierite-altered Amulet rhyolite and Millenbach
andesite. The porphyroblasts are enclosed in a groundmass
of chloritised biotite + sericite + quartz. Porphyroblasts are
metamorphic crystals that are surrounded by a much finer
grained matrix of other minerals (Spry, 1976). These large
minerals have formed at the expense of the matrix and are
the metamorphic equivalent of phenocrysts.
Archaean Amulet Rhyolite Formation, Noranda, Abitibi
greenstone belt, Amulet Upper A deposit, Canada.
B. Porphyroblasts in gneiss
This biotite + garnet gneiss is characterised by spotty
5 mm diameter garnet porphyroblasts in a medium- to
fine-grained quartz + feldspar + biotite groundmass. The
garnet porphyroblasts commonly have a biotite rim. A
gneiss is a rock with coarsely differentiated layering
denned by the segregation of minerals of different
composition (typically dark and light minerals) in
medium- to coarse-grained, granular rocks. Layering
forms parallel to the tectonic foliation and in this case
deviates around the garnet porphyroblasts. The precursor
is interpreted to have been a felsic volcaniclastic rock.
Sample 154061, Proterozoic Potosi gneiss, Harp prospect,
Broken Hill Block, New South Wales.
C. Porphyroblasts in thin section
This sample of garnet hornfels, from the contact zone
between rhyolite and a diorite intrusion, contains
euhedral garnet porphyroblasts in the biotite + muscovite
+ quartz groundmass. Porphyroblasts, like these, with
well-developed crystal shapes are idioblastic or euhedral,
whereas those with poorly developed crystal shape are
xenoblastic or anhedral (Yardley, 1989). Plane polarised
light.
Sample 140868, Cambro-Ordovician Mount Windsor
Formation, east Thalanga, Mount Charter, Queensland.
D. Poikiloblast in thin section
This thin section of amphibolite displays an amphibole
poikiloblast with quartz and biotite inclusion trails.
Poikiloblasts are porphyroblasts that contain numerous
inclusions that may or may not show a preferred
orientation (Barker, 1990). Poikiloblastic texture is
analogous to poikilitic or micropoikilitic texture. Usually
the inclusions are minerals that also occur in the matrix
(Yardley, 1989). In this sample, the biotite inclusion
trails display snowball rotation indicating syntectonic
growth of the amphibole. Plane polarised light.
Sample 3215, Proterozoic Corella Formation, Mary
Kathleen Group, Malbon River, northwest Queensland.
FIGURE 3.7 | Examples of static recrystallisation textures in altered volcanic rocks.
5 4 | CHAPTER 3
Common deformation textures include intergranular
textures and fabrics such as foliations, lineations, and augen-
structure, and intragranular textures such as strained, bent,
kinked, flattened, twinned and broken grains (crystals or
clasts), as well as irregular grain contacts (e.g. Fig. 3.8A,
B and C: Deelman, 1976; Spry, 1976). Deformation can
modify pre-existing textures such as volcanic, hydration and
devitrification textures (e.g. Fig. 3.8D to H). Fiamme and
eutaxitic deformation textures are unique to volcanic facies
(e.g. Fig. 3.9: Ross and Smith, 1960; Allen and Cas, 1990;
McPhie et al., 1993; Gifkins et al., in press). Fiamme and
eutaxitic textures are characteristic of, but are not restricted
to, welded ignimbrites (e.g. Fig. 3.9A and B: Ross and Smith,
1960; Smith, I960), welded pyroclastic fall deposits (e.g.
Sparks and Wright, 1979), welded autobreccia (e.g. Sparks
et al., 1993) and pyroclastic deposits that have undergone
secondary welding as a result of contact with hot lava or
intrusions (e.g. Ross and Smith, I960; Christiansen and
Lipman, 1966; Schmincke, 1967; McPhie and Hunns, 1995).
Similar fiamme and eutaxitic texture also occur in non-welded
altered pumice-rich rocks (e.g. Fig. 3.9C and D: Fiske, 1969;
Allen, 1988; Branney and Sparks, 1990; Gifkins et al., in
press) and felsic lavas (e.g. Pichler, 1981; Allen, 1988).
The terms fiamme and eutaxitic texture are used herein
to describe the rock texture and not to imply any particular
origin. Fiamme are flame-like, glassy or devitrified lenses,
which define a pre-tectonic foliation (cf. McPhie et al., 1993).
Fiamme may have a wide variety of sizes (0.5 mm to 1 m),
length to height ratios (up to 40:1), shapes (e.g. flame-like,
bow tie, irregular branching and blocky) and internal textures
(aphyric, porphyritic, vesicular or stylolitic) (Gifkins et al., in
press). Eutaxitic texture is the pre-tectonic foliation defined
by the parallel alignment of fiamme (cf. Fritsch and Reiss,
1868; Ross and Smith, I960; Smith, I960). Eutaxitic texture
typically imparts a blotchy or streaky appearance to the rock
due to the colour contrast between the darker fiamme and
paler matrix (e.g. Fig. 3.9A and C).
3.2 I PSEUDOTEXTURES
The incomplete destruction of primary textures and the
combined effects of a number of different overprinting
alteration styles (polyphase alteration) can result in significant
textural modification and the development of false textures or
pseudotextures (De Rosen-Spence et al., 1980; Allen, 1988;
McPhie et al., 1993). Pseudotextures are alteration textures
that modify or obscure primary volcanic textures and often
lead to incorrect interpretation of the primary volcanic facies.
Allen (1988) described examples of altered silicic lavas and
autobreccias from Benambra, New South Wales, that have the
remarkably deceptive appearance of welded and non-welded
pyroclastic facies and thinly bedded tuffaceous rocks.
Pseudotextures can be subdivided into pseudoclastic
textures (pseudobreccia, pseudogranular, false thin-bedded
volcaniclastic) or false pyroclastic textures (false shards,
false pyroclastic or eutaxitic: Fig. 3.10). However, strong
pervasive alteration can also produce false massive textures
that resemble either massive volcaniclastic or coherent facies
(Allen, 1988; McPhie et al., 1993; Doyle and Huston, 1999;
Doyle, 2001). Polyphase and patchy alteration of monomictic
volcaniclastic facies can also result in false clast-supported and
false polymictic textures.
Pseudoclastic textures
The most common pseudoclastic textures are pseudobreccia
and false pyroclastic texture (also referred to as false eutaxitic
texture). Other pseudoclastic textures include false thin-
bedded volcaniclastic and pseudogranular textures.
Pseudobreccias have the appearance of breccias, but form
as a result of alteration of coherent facies (Carozzi, 1960; Allen,
1988). In outcrop they resemble coarse-grained, monomictic
or polymictic, clast- to matrix-supported breccias comprising
angular to sub-rounded clasts in a fine-grained matrix (Fig.
3.10A, BandC).
False pyroclastic textures occur in both coherent facies
and in situ hyaloclastite. In outcrop and hand specimen they
may have a eutaxitic texture and contain abundant fiamme
(e.g. Fig. 3.10D). In thin section they appear to contain
splintery and arcuate fragments, which may closely resemble
pyroclastic glass shards (false shards: Fig. 3.10E).
Both pseudobreccia and false pyroclastic texture develop
as a result of two-phase alteration of fractured (perlitic or
quench fractured) coherent or autoclastic facies and domain-
controlled alteration of nodular devitrification textures in
coherent facies (e.g. Fig. 3.11: Allen, 1988).
Networks of intersecting quench and/or perlitic fractures
may control polyphase alteration in the fractured glassy parts
of coherent lavas and intrusions because they are permeable
pathways for fluid flow. Initially glass immediately adjacent to
the perlitic fractures is altered, then, as the fractures are filled,
replacement fronts migrate away from the perlitic fractures
towards the core. This may either obscure the continuity of
the perlitic fractures or, if alteration is incomplete, enhance
the perlitic fractures. False shard textures develop either due
to the preservation of less altered, relatively siliceous slivers
between two or more fractures, or as altered segments of
the fractures themselves (e.g. Fig. 3.10E: Allen, 1988). The
shape of false shards is a function of the shape of the fracture
network. For example, cuspate false shards are produced from
classical perlite, whereas those resembling flattened or welded
shards result from banded perlite (e.g. Fig. 3.2D). False clasts
develop where altered perlitic glass is partly overprinted by
a subsequent alteration phase, thereby preserving isolated
relicts of the earlier phase. Alternatively, the earlier phase
may be incomplete, leaving isolated kernels of glass that are
subsequently altered to a different mineral assemblage.
Pseudobreccia may also result from domainal or selective
alteration of nodular devitrification textures: spherulites
and lithophysae (e.g. Fig. 3.10F and G). Spherulites and
lithophysae are typically recrystallised to quartzo-feldspathic
compositions, whereas the interstitial originally glassy
domains are altered to phyllosilicate-rich assemblages (Allen,
1988). Consequently, the originally glassy and crystalline
domains differ in alteration mineralogy and colour, and the
spherulites appear as rounded siliceous clasts in a fine-grained
phyllosilicate-rich matrix.
Pseudogranular or sandy textures resemble well-sorted
sandstones. These result from the recrystallisation of densely
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 5 5
A. Augen schist
This sample of quartz-augen schist comprises large
lenticular quartz-rich domains (Q) enclosed in a strongly
foliated, sericite + quartz chlorite-altered matrix. Augen
texture is common in deformed, strongly porphyritic
coherent and crystal-bearing volcaniclastic rocks. This
augen schist probably resulted from the superposition of
a strong regional cleavage on an altered pumice breccia.
The cleavage anastomoses around competent silicified
pumice clasts.
Sample 040617, Cambrian western volcano-sedimentary
sequences, Mount Read Volcanics, Rosebery hanging wall,
western Tasmania.
B. Broken crystals in andesite
This deformed andesite contains broken plagioclase
phenocrysts in a strongly foliated, sericite + chlorite
+ magnetite epidote-altered groundmass. Broken or
fractured grains may result from mechanical pressure
during tectonic deformation (McBride, 1978). Typically,
feldspar crystals have been fractured along their
cleavage planes, whereas quartz crystals have developed
conchoidal fractures (Taylor, 1950; Sippel, 1968). Cross
polarised light.
Sample 144387, Ordovician Lake Cowal Volcanic
Complex, Junee-Narromine Volcanic Belt, Lake Cowal,
Gateway Prospect, New South Wales.
C. Deformed grains
In this amphibolite grade volcaniclastic siltstone, the
quartz grains are deformed polycrystalline grains with
undulose extinction, and elongated parallel to the
regional cleavage. Cross polarised light.
Sample GA9, Early Proterozoic Supra crustal succession,
Bergslagen mining district, Garpenberg, Sweden.
D. Deformed clasts and pillows
Pillow fragments and clasts in this basaltic hyaloclastite
were deformed and stretched parallel to the regional
foliation. The clast shapes are irregular and difficult to
recognise as pillow or hyaloclastite fragments. Despite
this, many clasts preserve an internal zonation.
Amphibolite, Proterozoic Corella Formation, Mary
Kathleen Group, Malbon River, northwest Queensland.
FIGURE 3.8 | Examples of deformation textures, deformed clasts and pre-existing textures in altered volcanic rocks.
5 6 | CHAPTER 3
E. Deformed clasts
Lens-shaped siliceous clasts (Q in this volcaniclastic
breccia have been rotated and stretched into the strong
tectonic cleavage. The fine-grained matrix has been
foliated and chlorite + sericite + quartz altered.
Sample 133520, Cambro-Ordovician Trooper Creek
Formation, Seventy Mile Range Group, Mount Windsor
Subprovince, central Thalanga, Queensland.
F. Folded pumice clast
This sample of rhyolite-, pumice- and crystal-rich breccia
contains a folded tube pumice clast with an axial planar
cleavage (Sj) defined by aligned sericite. The pumice
clast has been albite + quartz + sericite altered. Plane
polarised light.
Sample KB304B, Siluro-Devonian Ural Volcanics, Ural
Ridges area, New South Wales.
G. Deformed relict perlite
Relict perlitic fractures in this jigsaw-fit andesitic
breccia are elongate and flattened, especially adjacent
to competent phenocrysts. The groundmass has been
sericite + chlorite + calcite + albite altered and the perlitic
fractures chlorite filled. Plane polarised light.
Sample 76902, Cambrian Que-Hellyer Volcanics, western
volcano-sedimentary sequences, Mount Read Volcanics,
Hellyer, western Tasmania.
H. Deformed grains
In this sample of volcaniclastic sandstone, strongly'
deformed feldspar grains and clasts have been rotated
parallel to the strong cleavage. Plane polarised light.
Sample 133520, Cambro-Ordovician Trooper Creek
Formation, Seventy Mile Range Group, Mount Windsor
Subprovince, central Thalanga, Queensland.
FIGURE 3.8 | Examples of deformation textures, deformed clasts and pre-existing textures in altered volcanic rocks, cont.
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 5 7
A. Fiamme and eutaxitic texture in welded
ignimbrite
Dark flame-like obsidian lenses or fiamme (F) are aligned
in this sample of subaerial welded rhyolitic ignimbrite.
Fiamme are commonly interpreted as flattened pumice
clasts. The fiamme in this sample are interpreted to result
from the plastic deformation, flattening and sintering
together of hot glassy pumice clasts during welding
(cf. Smith, I960). The bedding-parallel alignment of
flattened, elongate fiamme and glass shards defines the
eutaxitic texture.
Sample OW7, Pleistocene Owahoroa ignimbrite, Whitianga
Group, Coromandel Volcanic Zone, Owharoa Falls, New
Zealand.
B. Fiamme in thin section
In thin section, the former pumice clasts, fiamme (F),
lack uncompacted vesicles, have feathery terminations
and are enclosed in domains of cuspate and platy shards
(5), and quartz, feldspar and biotite crystal fragments.
Although some shards have preserved bubble-wall
shapes, others were plastically deformed and compacted,
especially adjacent to crystals. Plane polarised light.
Sample OW11, Pleistocene Owahoroa ignimbrite,
Whitianga Group, Coromandel Volcanic Zone, Owharoa
Falls, New Zealand.
C. Fiamme and eutaxitic texture in non-welded
pumice breccia
Dark, plagioclase-phyric, wispy, chlorite-rich fiamme are
enclosed in pale domains of quartz + chlorite + pyrite-
altered pumice clasts, shards and crystal fragments in
this non-welded rhyolitic pumice breccia. The bedding-
parallel alignment of fiamme defines the eutaxitic texture.
Alteration and compaction of pumice clasts during
diagenesis formed these apparent welding textures.
Sample 133809, Cambrian Hercules Pumice Formation,
Central Volcanic Complex, Mount Read Volcanics, Hercules
footwall, western Fasmania.
D. Fiamme in thin section
In thin section, the chlorite fiamme (F) have feathery
terminations and lack internal textures other than sparse
plagioclase phenocrysts and hematite-rich stylolites. The
pale quartz + chlorite + pyrite-altered domains contain
uncompacted tube pumice clasts (P). Plagioclase crystals
are dusted with hematite and sericite. Plane polarised
light.
Sample 133811, Cambrian Hercules Pumice Formation,
Central Volcanic Complex, Mount Read Volcanics, Hercules
footwall, western Fasmania.
FIGURE 3.9 | Examples of fiamme and eutaxitic texture.
5 8 I CHAPTER 3
A. Pseudobreccia in perlitic rhyolite
Sericite + chlorite-altered perlitic fractures and pink
albite-altered perlitic cores result in the pseudoclastic
texture in this plagioclase-phyric coherent rhyolite.
Sample BBP248-504.7 m from the Cambrian Central
Volcanic Complex, Mount Read Voleanics, Boco, western
Tasmania.
B. Pseudobreccia in macroperlitic dacite
Polyphase alteration of macroperlite in this plagioclase-
phyric dacite has resulted in dark chlorite + epidote-rich
domains along and adjacent to the perlitic fractures.
This overprinted and enclosed earlier, pale grey albite
+ sericite-altered perlitic cores. The chlorite + epidote-
rich domains resemble the matrix in a matrix-supported
breccia. Locally, the arcuate perlitic fractures are well
defined.
Cambrian Sterling Valley Volcanics, Mount Read Volcanics,
Sterling Saddle, western Tasmania.
C. False clasts
In this coherent andesite, plagioclase phenocrysts have
been extensively replaced by epidote, pyroxenes by
chlorite and the groundmass domains by green epidote
+ chlorite and orange albite. The domainal distribution
of the alteration facies gives the andesite a patchy
pseudoclastic texture. The false clasts have both sharp
and diffuse margins, which are transgressed locally by
altered plagioclase phenocrysts.
Sample 145147, Ordovician Forest Reefs Volcanics, Molong
Volcanic Belt, Cooramilla, New South Wales.
D. False pyroclastic texture
The most conspicuous feature of this sample is the wispy
dark green chlorite-rich lenses that resemble fiamme.
However, these lenses are aligned in the tectonic cleavage
and occur in an evenly porphyritic rhyolite. The lenses
are interpreted to result from domainal chlorite and
sericite + quartz + biotite alteration of the groundmass
in a coherent quartz + plagioclase-phyric rhyolite.
Sample 140727, Cambro-Ordovician Mount Windsor
Formation, MountWindsorSubprovince, centralThalanga,
Queensland.
E. False shards in thin section
Polyphase chlorite + biotite and K-feldspar alteration
of perlitic fractures, and their subsequent deformation
have resulted in irregular arcuate and platy shapes, which
resemble shards (arrows) in this quartz + plagioclase
+ pyroxene-phyric coherent rhyolite. Plane polarised
light.
Sample KB499, Siluro-Devonian Coan rhyolite, Mount
Hope Volcanics, Mount Hope area, New South Wales.
FfGURE 3.10 | Examples of pseudotextures in altered volcanic rocks.
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 5 9
F. Pseudoclastic texture in devitrified rhyolite
This rhyolite contains silicified nodules that are composed
of coalesced spherulites in a fine-grained sericite-altered
groundmass. These nodular devitrification textures give
a clastic appearance to the hand specimen and outcrop.
Late Devonian Bunga Beds, Boyd Volcanic Complex,
Bengunnu Point, New South Wales.
G. False clasts in thin section
The clastic texture in this rhyolite comes from the
uneven distribution of strongly chlorite + hematite-
altered spherulites in a strongly foliated, chlorite +
sericite + feldspar + hematite-altered groundmass. The
foliation has wrapped around the altered spherulites,
which have preserved fibrous textures and quartz cores.
Plane polarised light.
Sample KB536D, Siluro-Devonian Mount Hope Volcanics,
Boolahbone tank, Mount Hope area, New South Wales.
I. Pseudogranular texture in thin section
Recrystallised micropoikilitic textures in the groundmass
of this aphyric rhyolite resemble sand-sized, rounded
grains in a well-sorted quartzo-feldspathic sandstone.
The margins of the micropoikilitic domains are marked
by concentrations of sericite, which enhance the granular
texture. Plane polarised light.
Sample 147448, Cambrian Kershaw Pumice Formation,
Central Volcanics Complex, Mount Read Volcanics,
western Tasmania.
H. Pseudogranular texture
This altered dacite has a fine granular texture in hand
specimen and resembles massive sandstone. However, in
thin section it has a densely microspherulitic groundmass
in which the recrystallised quartz + albite spherulites
are separated by cuspate sericite-rich domains. The
fine-grained, densely packed spherulites give the hand
specimen its sandy texture.
Sample MR25, Cambrian Kershaw Pumice Formation,
Central Volcanics Complex, Mount Read Volcanics, Mount
Read summit, western Tasmania.
FIGURE 3.10 | Examples of pseudotextures in altered volcanic rocks, cont.
6 0 | CHAPTER 3
J. False thin-bedded volcaniclastic texture
Flow banding in this microspherulitic plagioclase-phyric
dacite is defined by alternating pale albite + quartz and
darker albite + sericite + quartz layers. The planar,
repetitive, thin flow banding resembles thin bedding in
clastic facies such as tuffaceous siltstones.
Sample 76772, Cambrian Que-Hellyer Volcanics, western
volcano-sedimentary sequences, Mount Read Volcanics,
Hellyer, western Tasmania.
K. False thin-bedded volcaniclastic texture in thin
section
In thin section, this finely flow-banded rhyolite
resembles a thin-bedded volcaniclastic facies with
fractured plagioclase crystals. However, axiolitic and
bow-tie shaped spherulitic textures are locally preserved
in the groundmass. Cross polarised light.
Sample KB132A, Siluro-Devonian Ural Volcanics, Ural
area, New South Wales.
L. False volcaniclastic texture in thin section
Fractured and broken plagioclase crystals, and re-
crystallised spherulites in the groundmass of this flow-
banded rhyolite contribute to its pseudoclastic texture.
Cross polarised light.
Sample 133837 Cambrian Mount Black Formation,
Central Volcanic Complex, Mount Read Volcanics, Mount
Read summit, western Tasmania.
M. False polymictic, matrix-supported texture
The dark chlorite-rimmed clasts in this plagioclase-
phyric basaltic breccia appear to be supported in a
compositionally different, pale calcite + chlorite-altered
matrix. However, the matrix comprises jigsaw-fit, blocky
and splintery clasts of perlitic basalt that are identical
to the darker clasts. The chlorite-rimmed clasts appear
subrounded, because chlorite alteration of the clast
margins and adjacent matrix has obscured the blocky
and splintery shapes.
Sample 76833, Cambrian Que-Hellyer Volcanics, western
volcano-sedimentary sequences, Mount Read Volcanics,
Hellyer, western Tasmania.
FIGURE 3.10 | Examples of pseudotextures in altered volcanic rocks, cont.
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS I 6 1
N. False polymictic texture
Overprinting domainal albite + hematite and epidote
alteration facies in this plagiocla.se + hornblende-phyric
dacite gives the sample a heterogenous appearance. The
abundance of pink, albitised plagioclase phenocrysts in the
red and green domains is equivalent, although they appear
more abundant in the epidote-altered domains due to the
colour contrast between the phenocrysts and epidote-
altered groundmass. Colour differences between the two
alteration facies and more prominent phenocrysts in the
epidote-altered domains obscure the massive, coherent
texture and uniform composition of this sample.
Sample 147557, Cambrian Mount Black Formation,
Central Volcanic Complex, Mount Read Volcanics, Pieman
Road, western Tasmania.
O. False matrix-supported texture
Clasts in this monomictic dacite breccia display
jigsaw-fit texture. However, feldspar + quartz + sericite
alteration facies has replaced the groundmass adjacent
to the quench fractures between clasts, enhancing the
matrix, and imparting an apparent matrix-supported
texture. The feldspar + quartz + sericite-altered matrix
has been more resistant to weathering than the sericite +
chlorite-altered clasts and forms ridges on the outcrop.
The larger clasts are perlitic, plagioclase + hornblende
porphyritic with planar and curviplanar margins typical
of clasts produced by quench fragmentation.
Cambrian, Mount Black Formation, CentralVolcanic Complex,
Mount Read Volcanics, Tullah lakeside, western Tasmania.
P. False matrix-supported texture
In this in situ andesitic hyaloclastite, blocky chlorite-
altered plagioclase + pyroxene-phyric clasts appear to
be supported in a pyrite + quartz + sericite-rich matrix.
However, thin section inspection reveals relict clasts
with jigsaw-fit textures preserved in the false matrix
domains.
Sample 144710, Ordovician Lake Cowal Volcanics, Junee-
Narromine Volcanic Belt, Boundary Prospect, New South
Wales.
Q. False coherent texture
This albite + quartz + sericite- and chlorite + epidote-
altered pumice and rhyolite breccia resembles a coherent,
feldspar porphyritic facies. The albite-altered plagioclase
crystals, pseudophenocrysts, are evenly distributed in
a fine-grained, sericite-rich false groundmass. On the
left side of the photograph, pervasive, massive albite +
quartz alteration facies obscures the plagioclase crystals
in a pseudo-aphyric texture.
Sample 147402, Cambrian Kershaw Pumice Formation,
CentralVolcanic Complex, Mount Read Volcanics, Rosebery,
western Tasmania.
FIGURE 3.10 | Examples of pseudotextures in altered volcanic rocks, cont.
6 2 | CHAPTER 3
FIGURE 3.11 | Sketches summarising the relationship between primary volcanic, high-temperature devitrification and hydration textures, and pseudotextures.
(A) Pseudotextures in classical perlite (after Allen, 1988). False shards may be produced either (i) as siliceous segments between phyllosilicate-altered perlitic
fractures or (ii) phyllosilicate-altered sections of the perlitic fractures, (iv) Alternatively, pervasive phyllosilicate alteration can obscure the perlitic fractures resulting in a
massive texture.
(B) Pseudotextures in flow-banded facies comprising alternating crystalline (spherulitic) and glassy flow bands (modified from Doyle, 2001). (i) Glassy domains are
altered to phyllosilicate-rich assemblages and spherulitic bands to quartzo-feidspathic assemblages. Consequently the originally glassy and crystalline domains
differ in alteration colour and mineralogy. The spherulites superficially appear as rounded siliceous dasts in a fine-grained phyllosilicate-rich matrix, resulting in a
pseudobreccia texture, (v) In parallel flow-banded facies this can result in a false thin-bedded voicaniciastic texture. The crystalline nature of the quartzo-feldspathic-
altered bands enhances the granular appearance of the false beds.
(C) Pseudotextures from in situ quench fragmented porphyritic glass (modified from Doyle, 2001). Alteration progresses as fronts away from the fractures and into
the non-fractured glass, (i) Where phyllosilicate alteration is incomplete the remaining kernels of glass are subsequently altered to quartzo-feidspathic assemblages.
The further the alteration progresses away from the fractures, the more matrix-supported the resulting pseudobreccia texture, (ii) Pseudoclastic texture also develops
as a result of the complete replacement by the first alteration phase (phyllosilicate alteration) and only partial replacement by the second alteration phase (quartzo-
feidspathic alteration), thereby preserving isolated relics of the earlier phase, (iii) False polymictic texture develops in quench fragmented porphyritic glass as
polyphase alteration results in colour variations between the false dasts and matrix. Varying intensities of alteration enhance the polymictic appearance. Phenocrysts
are more prominent in dark phyllosilicate domains than in paler siliceous domains, resulting in an apparent variation in crystal content consistent with a polymictic
rock, (iv) Pseudomassive texture develops from pervasive phyllosilicate alteration.
(D) Pseudotextures from porphyritic autobreccia or hyaloclastite (modified from Doyle, 2001). (i) and (ii) alteration commences along fractures and in the matrix of
autobreccia but gradually progresses into the dasts, greatly enhancing the clastic, matrix-supported appearance. The result is commonly false matrix-supported
texture, (iii) Polyphase alteration results in pseudoclasts and pseudomatrix that comprise different alteration mineral assemblages and colours, and therefore appear
to have different primary compositions (false polymictic texture). Phenocrysts are more prominent in dark phyllosilicate-altered domains, (iv) Pseudomassive texture
occurs where alteration has been extensive and pervasive.
(E) Pseudotextures from porphyritic pumice breccia (modified after Gifkins, 2001). False eutaxitic texture results from phyllosilicate alteration of originally glassy
pumice dasts or shards within the pumice breccia. The phyllosilicate-altered pumice dasts and shards are flattened by compaction (i) or tectonic deformation
(ii) resulting in a texture that resembles eutaxitic texture in welded ignimbrites. (iii) False polymictic texture as a result of colour variations and apparent variations
in phenocryst content due to polyphase alteration, (iv) False coherent textures result from the complete and pervasive phyllosilicate alteration and subsequent
compaction of originally glassy shards and pumice dasts producing a massive textured rock in which the original dasts are indistinguishable.
spherulitic or micropoikilitic groundmass of coherent felsic
fades (e.g. Fig. 3. 10HandI).
False thin-bedded volcaniclastic texture resembles thinly
bedded and tectonically folded volcaniclastic rocks, and can
occur in altered, planar flow-banded and flow-folded lavas
and intrusions (e.g. Fig. 3.10J and K). The false thin-bedded
volcaniclastic texture is due to the planar and uniform character
of flow banding and the apparent textural and compositional
differences between flow bands (Allen, 1988). This apparent
compositional difference between flow bands results from
domainal alteration: originally glassy bands were altered to
dark phyllosilicate assemblages and contain well-preserved
phenocrysts, whereas the originally cryptocrystalline bands
were altered to pale quartzo-feldspathic assemblages and
contain altered polycrystalline phenocrysts that are hardly
recognisable. Phenocrysts may also be extensively fractured
and dismembered during deformation, thereby contributing
to the pseudoclastic and finer grained appearance (e.g. Fig.
3.10L). Discrepancies in colour and the apparent relative
proportions of phenocrysts enhance the compositional
contrast between adjacent bands.
False polymictic texture
Domainal phyllosilicate alteration may impart a heterogenous
appearance due to colour variation and to variable preservation
of phenocrysts. Pseudobreccias and pseudoclastic facies
may appear polymictic due to patchy or mottled alteration
distribution resulting in colour variations and variable
preservation of phenocrysts. This is particularly common in
pseudobreccias because the pseudoclasts and pseudomatrix
contain different alteration minerals, and hence colours:
therefore they appear to have different primary compositions
(e.g. Fig. 3.10M and N). In addition, phenocrysts (especially
feldspar or quartz) are more prominent in dark phyllosilicate
domains than in paler siliceous domains, resulting in an
apparent variation in crystal content between different
alteration domains consistent with different compositions
(e.g. Fig. 3.10N).
False matrix-supported texture
Alteration along fractures in coherent facies and in the matrix
of autoclastic facies (hyaloclastite and autobreccia) results in
colour variation between the clasts and fracture selvages or
matrix, and greatly enhances the clastic appearance. Where
alteration has progressed into the clasts from the clast margins,
the clasts appear to be isolated within a matrix (e.g. Fig. 3.10O
and P). The result is an apparently matrix-supported breccia
that is difficult to recognise as hyaloclastite or autobreccia
(Allen, 1988).
False coherent textures
False coherent textures, such as pseudomassive texture,
are less common than pseudoclastic textures. They result
from the complete and pervasive alteration of an originally
volcaniclastic facies. For example, pervasive phyllosilicate
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 6 3
alteration of pumice breccia produces a massive textured
rock in which the original clasts are indistinguishable (Fig.
3.10Q).
3.3 I ALTERATION DISTRIBUTION
Alteration distribution refers to the mappable extent of the
alteration facies and its relationship to host rocks, structures,
mineralised rock and other alteration mineral assemblages or
zones.
Although alteration distribution, zonation and textural
relationships are easily observed in thin section and hand
specimen they are more difficult to assess on a macroscopic
scale. This is due to the typically sparse exposure in ancient
volcanic successions, structural complications, commonly
patchy mode of occurrence of alteration facies in volcanic
rocks, and the considerable amount of detailed work required
to determine the distribution and zonation. Closely spaced
drill holes may adequately delineate the alteration facies
distribution and zonation at prospect scales. However,
alteration zones may be superimposed and the original zonation
patterns modified. It is important to recognise disequilibrium
alteration mineral assemblages and overprinting textures
when determining zonation patterns, so as to account for any
superposition caused by subsequent alteration styles.
Is the alteration facies limited in extent or widespread?
Typically the macroscopic alteration distribution is described
as either regional or local in extent. Regional alteration styles
are widespread, affecting extensive (hundreds of metres to
tens of kilometres) expanses of rock. Local alteration styles
are limited in extent and can refer to alteration on a scale
of centimetres, such as wall-rock alteration associated with
a fault or vein, to hundreds of metres, such as the extensive
hydrothermal alteration associated with the Rosebery VHMS
deposit (western Tasmania) that extends up to 100 m
stratigraphically beneath the ore lenses, 10-20 m into the
hanging wall and 500 m along strike (Large et al., 2001b).
Is there a relationship between the distribution of the
alteration facies and stratigraphy?
Alteration facies can be distributed within individual volcanic
units or formations or cut across stratigraphic contacts.
Alteration facies that are confined to individual units or
formations and are either concordant or discordant within the
unit are described as stratabound or semiconformable (Guilbert
and Park, 1986; Large, 1992). In contrast, alteration pipes and
alteration plumes refer to the distribution pattern or shape
of alteration systems or facies that transgress stratigraphic
contacts (pipes, e.g. Sangster, 1972; Gemmell and Large,
1992; Doyle and Huston, 1999, and plumes, e.g. Jack, 1989;
Gemmell and Fulton, 2001).
6 4 | CHAPTER 3
Is the alteration fades associated with other alteration
fades, structures, mineralised rock or particular units?
Locally distributed alteration fades are commonly associated
with structures, such as veins or fractures, ore bodies or
particular volcanic units: forming halos, envelopes or
selvages. They may also occur around isolated clasts within
clastic facies. In zoned alteration systems or halos, they are
commonly spatially and temporally associated with other
alteration facies.
3.4 I ALTERATION ZONATION PATTERNS
Alteration zonation is a regular pattern in the spatial distribution
of mineral species, mineral assemblages, major or trace
elements or textures, and reflects mineralogical and chemical
changes that relate to fluid-rock ratios and temperature
gradients (e.g. Meyer and Hemley, 1967). In general, the halo
of altered rock is divided mesoscopically into altered zones.
These are imprecisely defined as the areal extent of different
alteration mineral assemblages or facies. Boundaries between
hydrothermally altered zones associated with VHMS deposits
are commonly sharp, and reflect changes in the abundance of
particular alteration mineral species such as chlorite, sericite
or quartz. However, in some cases alteration facies overlap,
resulting in diffuse boundaries between altered zones (e.g.
Titley, 1982; Doyle and Huston, 1999).
In reality, altered zones are identified and named using
index minerals - the dominant alteration minerals - and zone
boundaries - isograds - are drawn at the first appearance of
an index mineral characteristic of the next zone. Thus the
clinoptilolite + mordenite zone is placed at the first appearance
of clinoptilolite or mordenite (e.g. Iijima, 1978). In some cases
an altered zone may contain more than one alteration facies.
For example, at Thalanga the tremolite + chlorite carbonate
zone includes intense pervasive stratabound chlorite + tremolite
(no carbonate), intense pervasive stratabound chlorite +
calcite, and intense stratabound tremolite + dolomite + calcite
(no chlorite) facies. The expression of zoning may be limited
by exposure and modified by structural and compositional
homogeneity, faulting and/or intrusions.
There are three scales of zonation that are generally applied
to the relative distribution of metals (e.g. Kutina et al., 1965)
TABLE 3.3 I Scales of alteration zonation.
Regional zonation
District zonation
Local zonation
Sample-scale
zonation
Alteration facies or minerals occur in zones
throughout a region
Altered zones are associated with a cluster
of ore bodies, fractures or intrusions
Altered zones are associated with an
individual ore body, fracture or intrusion
Altered zones are associated with small-
scale features such as clasts or minerals in
the rock
and that can also be applied to the distribution of alteration
facies or minerals (Table 3.3): (1) regional zoning, (2) district
zoning, and (3) local or ore body zoning. In addition, altered
zones may occur within or surrounding individual clasts
in clastic facies. Thus the dimensions of altered zones may
vary from a few centimetres to several tens of kilometres
(cf. Bohlke et al., 1980; Galley, 1993). These variations in
dimension are a function of the size of the alteration system
(regional versus local systems), and changes in physical and
chemical conditions such as the porosity and composition of
the host rock, fluid-rock ratio and composition of the fluid
(Rose and Burt, 1979).
Different alteration processes result in different zonation
patterns. Regional alteration processes such as diagenesis
and metamorphism, produce thick, flat-lying, vertically
zoned systems. Local hydro thermal systems and contact
metamorphism result in footwall and hanging wall altered
zones or concentrically zoned altered halos or contact aureoles.
Zonation related to faults or fractures is generally parallel to
the structure, commonly cross cutting stratigraphic contacts.
Regional diagenetic zones
Diagenesis develops in response to increasing temperature
with depth during burial; as a result it forms a sequence
of flat-lying altered zones (Fig. 3.12: Iijima, 1974, 1978;
Fisher and Schmincke, 1984; Utada, 1991). These zones
are characterised by mineral assemblages, which reflect the
reaction of glass, primary minerals and diagenetic minerals
with interstitial pore water and seawater at a particular
temperature anywhere between 0 and 250C (Hay, 1978;
Iijima, 1978; Utada, 1991).
Sequences of diagenetic zones are typically between 500 m
and 6 km thick (Hay, 1978) and individual altered zones vary
from a few metres to several kilometres in thickness (e.g. Hay,
1978; Iijima, 1978; Vierecketal., 1982; Passagliaetal., 1995).
The thickness of diagenetically altered zones is dependent on
the geothermal gradient, rate of burial, and the porosity and
permeability of the volcanic succession.
The Miocene Hokuroku Basin, part of the Green Tuff
Belt in northern Honshu, Japan, is an excellent example of
diagenetic zonation (Fig. 3.12). The Hokuroku Basin contains
a 3 to 6 km thick submarine volcanic succession dominated
by rhyolitic to dacitic and minor basaltic coherent and clastic
volcanic facies (Horikoshi, 1969; Iijima, 1978; Tanimura et
al., 1983; Urabe, 1987; Utada, 1991). Diagenetic alteration
in the Hokuroku Basin has produced a series of flat-lying
diagenetic zones, which grade vertically from fresh glass at
the top, to smectites, zeolites and albite at depth (Fig. 3.12:
Iijima, 1978).
Regional metamorphic zones
Regional metamorphic zones develop due to regional
changes in temperature and pressure. Unless metamorphism
is related solely to orogenic deformation, progressive burial
results in dehydration and increasing metamorphic grade
with depth. This produces a vertical sequence of flat-lying,
regional metamorphic zones (Fig. 3.13B). It is generally
FIGURE 3.12 | Schematic cross-section of vertically developed diagenetic
zones in the central Hokuroku Basin, Japan (after lijima, 1974; Hay, 1978; lijima,
1978).
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 6 5
assumed that higher-grade metamorphic rocks formerly
had mineral assemblages typical of lower-grade zones and
were altered progressively as metamorphism proceeded.
However, variations in geothermal gradient, variable rates
of sedimentation and folding and faulting can modify the
progression of metamorphism and the resulting metamorphic
zones, in particular, folding and faulting may affect the
position of mineralogically determined isograds by promoting
reactions or causing local increases in temperature or heat flow
(Coombs et al., 1959).
Coombs (1954) first recognised vertical metamorphic
zoning of low-temperature, high-pressure metamorphic
mineral assemblages in a 10 km section of the uplifted
Permian to Jurassic Wakatipu Metamorphic Belt, in the
New Zealand geosyncline (Fig. 3.13A). The Wakatipu
Metamorphic Belt comprises submarine emplaced rhyolitic
to basaltic volcaniclastic and volcanogenic sedimentary rocks,
which were progressively metamorphosed (Coombs, 1954).
There is a continuous gradational sequence of regional
metamorphic zones characterised by zeolite, prehnite +
pumpellyite, pumpellyite + actinolite and greenschist facies (a
biotite zone and chlorite zone) at depth (Fig. 3.15B: Coombs,
1954; Coombs etal., 1959; Landis and Coombs, 1967). These
zones have an even thickness over a wide area (>10 km by
300 km) suggesting that the geothermal gradient, estimated
at l4-25C/km, was consistent throughout the Wakatipu
Metamorphic Belt (Landis and Coombs, 1967).
FIGURE 3.13 | Alteration zonation patterns associated with regional
metamorphism. (A) Regional metamorphic zones in the Wakatipu metamorphic
belt, South Island, New Zealand, (after Coombs et al.,1959, and Landis
and Coombs,196 7). (B) Schematic cross-section reconstruction of regional
metamorphic zones and isotherms during peak metamorphism of the Wakatipu
metamorphic belt, New Zealand (after Landis and Coombs, 196 7).
6 6 | CHAPTER 3
Regional, deep, semi-conformable altered zones
Deep, semi-conformable altered zones are interpreted to be
the products of hydrothermai alteration within regional,
subseafloor hydrothermai systems (Galley, 1993). These
regional hydrothermai systems involve the large-scale
convection of modified seawater driven by the emplacement
of synvolcanic intrusions into the subsurface (Spooner and
Fyfe, 1973; Norton, 1984; Cathles, 1993; Galley, 1993).
The upper contacts of sub-surface intrusions are typically
sub-parallel to the volcanic strata and hence the overlying
isotherms are also semi-conformable (Galley, 1993). The
result is a series of vertically stacked, sub-horizontal altered
zones, which are characterised by mineral assemblages that
reflect reactions between host volcanic facies and modified
seawater at temperatures transitional with diagenesis and
greenschist facies metamorphism (Galley, 1993; Skirrow
and Franklin, 1994). Sequences of deep, semi-conformable
altered zones may be up to 20 km wide and between 1 and
2 km thick (Gibson et al., 1983; Galley, 1993; Skirrow and
Franklin, 1994).
One of the first comprehensive descriptions of deep, semi-
conformable altered zones is by Gibson et al. (1983). These
authors documented a series of vertically stacked altered zones
in the Noranda district of the Archaean Abitibi belt, Canada
(Fig. 3.14). The Amulet Rhyolite Formation comprises
coherent and clastic basalts and andesites with minor
rhyolitic lava domes. The volcanic rocks are variably altered
and the distribution of the alteration facies can be related
to stratigraphic depth and primary host rock composition.
Within the andesites there are four regionally extensive
alteration facies, from top to bottom: chlorite + actinolite +
albite + epidote + quartz; silicified; mottled epidote + quartz;
and chlorite alteration facies (Fig. 3.14B). The upper chlorite
+ actinolite + albite + epidote + quartz alteration facies is
the result of low-grade regional metamorphism, whereas the
other three alteration facies are interpreted to be the products
of regional deep, semi-conformable alteration (Gibson et al.,
1983).
Local contact metamorphic or hydrothermally
altered halos
Halos or contact aureoles associated with granitoids, thick
synvolcanic sills, cryptodomes and dykes reflect changes
in the composition and temperature of the magmatic or
hydrothermai fluid away from the intrusion, and its interaction
with local pore water (Rose and Burt, 1979; Einsele et al.,
1980; Yardley, 1989). The result is a progressive sequence
of roughly concentric altered zones or metamorphic zones
FIGURE 3.14 | Diagram showing alteration zonation patterns associated with regional hydrothermai alteration.
(A) A schematic cross-section through a vertically stacked sequence of deep, semi-conformable altered zones in the Amulet
Rhyolite Formation, Noranda district, Abitibi belt, Canada (after Gibson, 1989, in Galley, 1993). (B) A reconstructed section
of deep, semi-conformable altered zones in the Amulet Rhyolite Formation at Turcotte Lake (after Gibson et al., 1983).
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS I 6 7
that correspond with decreasing temperature away from the
intrusion contact (e.g. Fig. 3.15). For example, submarine
basaltic lavas and breccias of the Upper Triassic Karmutsen
Subgroup, Vancouver Island, Canada, have undergone low-
pressure, high-temperature contact metamorphism related to
the shallow emplacement of the Jurassic Coast Range Batholith
(Carson, 1973; Kuniyoshi and Liou, 1976). Intrusion of the
Coast Range Batholith resulted in two locally developed
contact metamorphic zones: hornblende + plagioclase zone
and epidote + actinolite zone, which are superimposed on
a prehnite + pumpellyite facies regional metamorphic zone
(Fig. 3.15A: Kuniyoshi and Liou, 1976). The hornblende
+ plagioclase zone is approximately 2600 m wide and the
epidote + actinolite zone less than 900 m wide (Kuniyoshi
and Liou, 1976).
Local hydrothermally altered halos around ore
deposits
Local halos of hydrothermally altered rock around
mineral deposits result from the interaction between the
hydrothermal fluid/s responsible for mineralisation and the
surrounding country rock. In the case of VHMS deposits,
hydrothermal fluid temperatures range up to 35OC (Large,
1977, 1992; Urabe et al., 1983; Gemmell and Large, 1992).
The migration of hydrothermal fluids and mixing with cold
seawater at the margins of the hydrothermal system produces
three-dimensional altered zones composed of successively
lower temperature mineral assemblages away from the site of
mineralisation (Date et al., 1983; Urabe et al., 1983; Large,
1992; Schardt et al., 2001). The shape, dimensions and
distribution of fluid circulation (and thus the altered zones)
are usually closely related to initial patterns of permeability,
porosity and compositional contrasts in the volcanic succession
(e.g. Yamagishi and Dimroth, 1985; Large, 1992; Doyle,
2001). Accordingly, the altered halos exhibit a wide variety
of shapes and zonation around VHMS deposits, typically
with intense proximal footwall and weak hanging wall zones
(Large, 1992). In contrast, altered halos around porphyry
deposits may have a more concentric zonation, comprising
a potassic core zone, an outer propylitic zone and, in some
examples, minor phyllic zones (Gustafson and Hunt, 1975).
Later hydrothermal alteration and mineralisation zones may
overprint these initial zones, such as the widespread phyllic
and argillic zones at the El Salvador deposit, Chile and Batu
Hijau in Indonesia (Gustafson and Hunt, 1975; McMillan
and Panteleyev, 1998; Garwin, 2002).
Hydrothermally altered zones associated with VHMS
deposits can extend laterally up to 500 m and persist at least
500 m stratigraphically (Iijima, 1974; Utada et al., 1974;
Eastoe et al., 1987; Gemmell and Large, 1992; Gemmell
and Fulton, 2001). Footwall zones occur either as diffuse
stratiform zones (e.g. Rosebery, Fig. 3.16C: Green et al., 1981;
Large et al., 2001b) or well-defined and zoned alteration
pipes (e.g. Hellyer, Fig. 3.16A: Millenbach, Fig. 3.16D:
Franklin et al., 1981; Morton and Franklin, 1987; McArthur,
1989; Gemmell and Large, 1992). Hanging wall alteration is
normally of lower intensity, and developed above the thickest
part of the ore body either as a diffuse stratabound zone (e.g.
Woodlawn and Scuddles) or an alteration plume (e.g. Hellyer,
FIGURE 3.15 | Diagram showing alteration zonation patterns associated with
contact metamorphism. (A) Contact metamorphic zones in the northeastern
Vancouver Island, Canada (from Kuniyoshi and Liou, 1976 ). (B) Schematic
cross-section of contact-metamorphic zonation.
Fig. 3.16A: Large and Both, 1980; Jack, 1989; Large, 1992;
Gemmell and Fulton, 2001).
In the Hokuroku district, local hydrothermally altered
halos associated with the Kuroko deposits comprise up to
four altered zones in a roughly concentric distribution around
the ore deposits (Fig. 3.16B: Urabe et al., 1983; Utada et al.,
1988; Utada, 1991). From core to margin the altered zones
are: sericite + quartz pyrite zone; chlorite + sericite + quartz
zone; mixed-layer clay (illite-montmorillonite) zone; kaolinite
or smectite-montmorillonite zone (Date et al., 1983; Urabe
et al., 1983). These altered zones vary in thickness from
approximately 16 m to 100 m, transgress stratigraphic
boundaries, and grade into and interfinger with regional
diagenetic zones such as the clinoptilolite + mordenite zone
in Figure 3.16B (Utada, 1970, 1991; Iijima, 1974).
In contrast, the mineral assemblages in altered zones
associated with Australian VHMS deposits reflect higher-
grade regional metamorphic overprints. Some have proximal
chlorite zones and halos of quartz + sericite, sericite and
carbonate zones (Fig. 3.16A and C: Allen, 1988; Large, 1992;
Doyle, 2001).
Vein and fracture altered halos
Altered halos around fractures or veins are typically limited
in width (from millimetres to tens of metres) occurring
6 8 | CHAPTER 3
FIGURE 3.16 | Schematic cross-sections showing hydrothermal alteration zonation
patterns associated with VHMS deposits. (A) Hydrothermally altered zones developed in
the footwall alteration pipe and hanging wall alteration plume to the Hellyer VHMS deposit
(after Gemmell and Fulton, 2001). (B) Hydrothermally altered halo developed around
the Uwamuki group deposits, Kosaka VHMS mine, Japan (after Urabe et al., 1983).
(C) Hydrothermally altered halo developed around the K-lens VHMS ore lens at Rosebery
(after Large et al., 2001b). (D) Hydrothermally altered zones developed in the footwall
alteration pipe and hanging wall alteration plume to an ore lens in the Millenbach VHMS
deposit, Canada (after Riverin and Hodgson, 1980).
either as narrow selvages around individual veins or as wide
pervasive fracture-controlled altered zones. Decreasing fluid
temperatures and changes in fluid chemistry (especially pH)
away from fractures or faults results in alteration zonation
parallel to the fracture or vein surface.
3.5 | OVERPRINTING RELATIONSHIPS
AND TIMING OF ALTERATION
The aim in determining overprinting relationships is to
establish a time sequence for mineral growth, which is referred
to as a paragenetic sequence (Kutina et al., 1965). Normally
paragenesis refers to the growth of minerals from oldest to
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 6 9
youngest. The paragenetic sequence is typically depicted
using a horizontal bar chart (e.g. Fig. 3.17) or a space-time
diagram, such the schematic diagram by Wilson et al. (2003)
that shows alteration and vein stages relative to the intrusive
history of the Ridgeway Complex at the Ridgeway Au-Cu
porphyry deposit in New South Wales.
Unfortunately, interpreting timing relationships between
different alteration facies can be difficult and confusing
as there are few unambiguous overprinting textures. The
superposition of many different alteration phases on the
same rock can obscure or complicate alteration textures. In
addition, while one mineral is being deposited under certain
conditions at one place, other minerals are forming elsewhere
under different conditions. Thus the deposition of one
mineral may overlap with another in both time and space.
F IG U R E 3 . 1 7 | I nterpretation of the relative timing of alteration mineral assemblages in the northern Central V olcanic Complex, western
Tasmania (modified after Gifkins and Allen, 2 001 ). S1 is the bedding-parallel, stylolitic foliation interpreted as a diagenetic compaction
and dissolution fabric (Allen, 1 990). S2 is the regional tectonic cleavage related to Devonian folding. Assemblages in brackets refer to
pre-metamorphic equivalents. For example, early regional sericite is probably a metamorphosed equivalent of early clays. The feldspar
+ quartz + sericite alteration mineral assemblage may include the growth of early zeolites and the replacement of zeolites and glass by
K-feldspar and albite during diagenesis.
70 | CHAPTER 3
Method
The relative timing of alteration mineral assemblages is
determined by documenting overprinting relationships
among the alteration mineral assemblages, mineral deposits,
successive volcanic units or intrusions, the compaction
foliation and regional tectonic cleavages. This requires
detailed examination of textural and mineralogical features
that are visible in the field, hand specimen (including drill
core) and thin section. Although thin sections can be useful in
displaying the inter-relationships between different minerals,
they may be too small to show larger cross-cutting features
(e.g. veins) and textural features in coarse-grained rocks.
Paragenetic determination relies on a representative suite of
samples that includes all rock types and alteration facies in
the alteration system. Because overprinting relationships are
not always straightforward the inspection of a large number
of samples is required for a systematic understanding of the
overprinting textures.
Overprinting textures
Overprinting texture refers to any texture or geometry that
can be used to infer that one mineral or group of minerals
has formed later than another mineral. Table 3.4 describes
the three main types of overprinting textures. There is a wide
range of overprinting textures, some of which are described
in Sections 3.1 and 3.2; however, the reader is referred
to Ramdohr (1980), Craig and Vaughan (1981), Ineson
(1989) and Taylor (1996) for more detailed discussions of
overprinting textures and their significance.
The possible relationships among alteration mineral
assemblages, mineralised rocks, successive volcanic units or
intrusions, the compaction foliation and regional tectonic
cleavages and their implication for the timing of alteration
are described in Table 3.5.
Recent work (Allen, 1997; Gifkins, 2001; Gifkins and
Allen, 2001) in pumice breccias around the Rosebery VHMS
deposit has unravelled a complex sequence of hydrothermal
and non-hydrothermal alteration mineral assemblages.
This work was based on overprinting relationships between
alteration mineral assemblages, compaction foliation (S
:
) and
tectonic cleavage (S
2
: Fig. 3.17). Alteration mineral assemblages
that infill primary porosity, preserve delicate uncompacted
TABLE 3.4 | Types of overprinting textures.
Mineral superposition Where one mineral or a group of minerals
can be seen to have nucleated upon a
pre-existing mineral or grain; this applies
to replacement, infill and recrystallisation
textures
Where fractures fragment pre-existing
minerals or grains, provide space for
subsequent infill and focus wall-rock
alteration
Foliation superposition Where foliations or lineations rotate, modify,
distort or fracture pre-existing minerals or
components
Fracture superposition
pumice textures and/or are overprinted by Sj and S
2
are
interpreted to be early, prior to compaction and complete
lithification. These include thin films of sericite, chlorite +
sericite + hematite and calcite (or their pre-metamorphic
equivalents, smectites, calcite and palagonite) that coated all
original surfaces including vesicle walls, plagioclase crystals,
shards and fractures and pre-date the infilling of vesicles by
subsequent minerals. Feldspar + quartz + sericite, chlorite +
sericite, chlorite + sericite + hematite, chlorite, sericite and
quartz + sericite alteration facies have filled and preserved
vesicles, and replaced glass, indicating that they (or their pre-
metamorphic equivalents such as zeolites and clays) formed
prior to compaction and deformation.
Alteration features that define the bedding-parallel
compaction foliation, such as chlorite + sericite + hematite
and chlorite + sericite fiamme, are crenulated by the regional
cleavage (S
2
) and are interpreted to be pre- to syn-S,.
Alteration features that overprint the early alteration
mineral assemblages and S
p
but which are strongly foliated
by S
2
, are broadly syntectonic or post-Sj pre-S
2
. A chlorite
+ epidote alteration assemblage has overprinted the early
sericite films and chlorite + sericite fiamme but is strongly
foliated by S
2
and interpreted to be post-Sj and pre- to syn-S
2
.
Chlorite + calcite + magnetite has replaced and post-dated the
chlorite + sericite fiamme and chlorite + sericite + hematite
stylolites that define S
v
and which are commonly transposed
into the S
2
cleavage and associated with syn-S
2
chlorite veins.
Chlorite pyrite is associated with shear zones parallel to S
2
,
and undeformed post-S
2
brittle fractures and faults. Sericite
+ calcite + actinolite epidote recrystallised earlier alteration
mineral assemblages and are aligned parallel to S
2
.
Alteration mineral assemblages that are unaffected or only
very weakly affected by the S
2
foliation are interpreted to be
post-deformation. These alteration mineral assemblages are
related to chlorite + quartz + calcite vein infill and associated
vein wall-rock replacement and altered halos associated with
Devonian granites, such as quartz calcite tourmaline +
magnetite veins.
Early Mn-carbonate, chlorite, sericite and quartz + sericite
alteration facies are spatially associated with ore at Rosebery
(Fig. 3.16C), whereas the other alteration facies are regionally
extensive, spatially associated with synvolcanic or Devonian
intrusions, veins, faults or shear zones. The overprinting
relationships, combined with spatial associations and
distributions of the alteration facies, suggest that diagenetic
alteration began shortly after eruption in the Cambrian and
continued until the transition to regional metamorphism.
Hydrothermal alteration associated with the VHMS deposits
at Rosebery and Hercules commenced prior to burial
compaction, but was synchronous with diagenesis and the
intrusion of synvolcanic sills. Peak regional metamorphism
was synchronous with Devonian deformation and is
overprinted by contact metamorphic assemblages associated
with the emplacement of Devonian granites.
The interpretation of textures to determine a paragenetic
sequence can result in misleading results unless the subsequent
effects of metamorphism and deformation upon the textures
and minerals are appreciated. For example, in the Que River
VHMS deposit galena commonly occurs in the S
2
cleavage.
As a result, conventional interpretation may conclude that
it formed syn-S
2
(Devonian). This interpretation would
suggest that galena post-dates the other sulfides (sphalerite
and pyrite), which are overprinted by S
2
. However, geological
evidence (Large et al., 1989) combined with lead isotope data
(Gulson and Porritt, 1987) indicates that galena deposition
occurred on the seafloor at the same time as the other sulfides
in the Cambrian. In this case, the textural evidence constrains
COMMON ALTERATION TEXTURES AND ZONATION PATTERNS | 7 1
the timing of recrystallisation of the galena to syn-S
2
, whereas
the geological and isotopic evidence suggests syn-depositional
formation of galena. Misinterpretations of this type are
common for soft and easily recrystallised minerals such
sericite, chlorite, clay minerals, galena and chalcopyrite.
as
TABLE 3.51 Overprinting relationships in altered volcanic rocks and their implications for the relative timing of alteration (modified after
Allen and Large, 1996 )
Alteration facies to primary volcanic texture
Alteration facies truncated by clast margins
Clasts with different alteration facies in the same rock
Alteration facies infills primary porosity
Alteration facies cross cuts primary porosity or clast margins
Rock contains relict high-temperature devitrification textures
Alteration facies to successive volcanic units or intrusions
Alteration facies truncated by younger less-altered rocks
Alteration facies overprints intercalated intrusions
Alteration facies overprints younger rocks
Between alteration facies and diagenetic compaction textures
Alteration facies protects primary texture from compaction
Alteration facies defines the compaction foliation (e.g. fiamme)
Alteration facies overprints the compaction foliation
Hydrothermal assemblages and diagenetic facies
Hydrothermal facies overprinted by early diagenetic facies
Hydrothermal facies overprinted by late diagenetic facies
Hydrothermal facies overprints late diagenetic facies
Alteration facies to early bedding-parallel stylolitic dissolution foliation (S1)
Alteration facies overprinted by stylolitic foliation
Alteration facies overprints stylolitic foliation
Alteration facies to tectonic foliations and lineations (>S2)
Alteration textures deformed and cut by tectonic foliation
Alteration textures less deformed than primary volcanic textures
Undeformed alteration textures
Alteration facies to metamorphic assemblages and textures
Alteration facies overprinted by metamorphic assemblages
Alteration facies overprints metamorphic assemblages
Megascopic alteration facies distribution
Regional distribution
Stratabound in formerly permeable facies
Localised in fractured domains of coherent volcanic rocks
Restricted to faults and shear zones
Overprinting relationships between all alteration assemblages
Pre-fragmentation
Pre-frag mentation
Pre-lithification
Post-lithification
Post-devitrification
Older syn-volcanic/intrusion
Syn- to post-intrusion and overlying units
Syn- to post-younger rocks
Pre-compaction
Pre- to syn-compaction
Syn- to post-compaction
Pre- to early-diagenetic
Syn-diagenetic
Post-diagenetic
Pre- S1
Post-S1
Pre-cleavage
Syn-cleavage
Post-cleavage
Pre-metamorphic
Post-metamorphic
Diagenetic or metamorphic
Pre- to syn-lithification
Post-lithification and post-fracturing
Syn- to post-faulting
Provides a paragenetic sequence for
different alteration assemblages
72
I 73
4 | GEOCHEMISTRY OF ALTERED ROCKS
This chapter reviews three principal types of geochemical
data that are used to characterise, quantify and interpret the
processes of alteration: (1) whole-rock lithogeochemistry
including major, trace and immobile trace elements; (2)
mineral chemistry or composition of individual mineral phases
in alteration mineral assemblages; and (3) stable isotope ratios
of whole rocks and specific alteration minerals.
Geochemistry is the study of the abundance and
distribution of chemical elements in the earth (Mason, 1966).
It has particular applications for the interpretation of altered
rocks. Geochemical data can aid recognition of the type and
composition of the precursor rocks before alteration. The
chemical compositions of metasomatised or altered rocks
and mineral phases reflect compositional changes during
alteration and provide clues about the alteration processes.
On a utilitarian level, geochemical data are routinely used
in mineral exploration to identify broad chemical halos
and gradients or vectors toward mineralised zones, and
to discriminate between prospective and non-prospective
targets.
4.1 | LITHOGEOCHEMISTRY
Literally, lithogeochemistry is the determination and study of
the abundances and distribution of elements in stones. In this
chapter, we understand it to mean the whole-rock chemistry
of the 10 or 12 major elements in rocks (usually expressed
as oxides) and various groups of trace elements, such as rare
earth elements (REE) and immobile elements. Galley's (1995)
review of lithogeochemical applications in massive-sulfide
exploration also included mineral and isotope chemistry but
those methods are considered here in separate Sections (4.2
and 4.3).
There are three main applications of lithogeochemistry in
mineral exploration (Eilu et al., 1997):
(1) identification or discrimination of prospective and non-
prospective areas and lithological units
(2) recognition of large alteration and geochemical halos to
increase the size of exploration targets
(3) definition of exploration vectors based on compositional
gradients around ore deposits.
Several aspects of lithogeochemistry that attempt to 'see
through' the effects of alteration to reveal the compositions
of the protoliths, identify the process or type of alteration
and quantify the chemical changes associated with alteration
have particular relevance to studies of altered rocks. Chemo-
stratigraphy uses immobile elements to identify lithologic
correlations, magmatic affinities and geotectonic settings
in otherwise unrecognisable altered rocks. Mass-transfer
techniques, also based on immobile elements, are used to infer
the compositions of unaltered precursors and quantitatively
estimate the major element changes that occurred. Rare-
earth-element geochemistry can facilitate recognition of rock
alteration processes (e.g. Eu anomalies in exhalites). Alteration
indices may assist in the discrimination of alteration styles or
facies, and the quantification of alteration intensity (Section
2.4).
If composition data are available, lithogeochemical
methods certainly contribute to the interpretation and
evaluation of altered rocks. However, they are not quick
solutions to all problems and may, if used in isolation, lead
to false conclusions. In every case it is important to consider
other geological information: field relationships, distribution
or zonation of alteration mineral assemblages, macro- and
micro-scale textures, mineralogy et cetera.
Before describing the lithogeochemical methods, we
provide explanations of some common lithogeochemical
terms (Table 4.1), brief descriptions of geochemical sampling
and analytical techniques, and draw particular attention to
the phenomenon of closure in composition data.
Sampling and analytical methods
How do we acquire lithogeochemical data? As in many other
fields, the quality of interpretation rests on the quality of
the data. Rocks are particularly difficult to analyse because
of their wide compositional range and the chemical diversity
of the elements of interest. An inappropriate analytical tech-
nique can lead to an expensive list of useless numbers and
to the discredit of the unwary geologist who attempts to use
them. This section offers advice on effective methods (they
are never inexpensive) and some pitfalls to be avoided.
74 | CHAPTER-
TABLE 4.1 | Explanation of lithogeochemical terms.
Composition data
Mass
Units of mass change
Absolute mass change (A
a
)
Relative mass change (A
r
)
Net mass change
Proportion
Data expressed as part of a whole (e.g. in weight percent, wt%, or parts per million, ppm).
A quantity of material, whole rock or its components, measured in weight units (g).
Usually expressed in grams per hundred grams (g/100 g) to avoid confusion with composition data in weight
percent, although they are effectively the same.
Mass changes that are proportional to the whole initial mass of the rock. Usually expressed in grams per hundred
grams (g/100 g). It may refer to individual components or the sum of all components (net mass change).
The absolute mass change in a component expressed as a percent proportion of the initial composition of that
component in a rock. Relative mass changes distort the perception of chemical processes. For example, the
addition of a small quantity (say 0.5 g/100 g) of MnO initially present at 0.5% would produce a 100% relative gain
in MnO. In comparison, addition of the same quantity of SiO
2
(0.5 g/100 g) to the same rock initially containing
50% SiO
2
would represent only 1% relative gain. This gives the false impression that MnO was a vastly
more important element of mass change than SiO
2
. Absolute mass changes are preferable because they are
quantitative rather than proportional, and accurately reflect the quantities of materials added to or subtracted from
the system.
The sum of positive and negative rock component mass changes relative to the initial mass of the rock. Usually
expressed in grams per hundred grams (g/100 g).
The amount of a component expressed as a proportion of the whole rock. Usually given in units of weight
percent (wt%) for major elements and oxides, and in parts per million (ppm) for trace elements. It is analogous to
" concentration" in chemical solutions.
Sampling strategy
The number and size of samples to be used in a lithogeochemical
investigation depends on many factors. These include the
available budget, degree of exposure, geological complexity,
compositional homogeneity of rock units or altered zones,
the elements of interest and the volume of rock that can be
physically removed from the outcrop or drill-core shed. There
can be no universally applicable strategy but generally, more
is best. Obviously, a good understanding of the geological
context is fundamental to any subsequent lithogeochemical
interpretation. Smash and grab sampling may lead to in-
adequate and often meaningless interpretations.
Field methods should avoid sampling rocks potentially
modified by weathering or other superficial secondary
processes, unless these processes are under investigation. The
sampler must also recognise potential overprinted alteration
facies. For example, a sample containing a dense network of
metamorphic quartz veins is of little value if the purpose is
to study synvolcanic pervasive alteration phases; it would be
better to select a sample without veins. Handling, transport
and subsequent storage of lithogeochemical samples should
not allow contamination or cross contamination. When
sampling drill cores or cuttings stored at a mine site, be aware
that mines and concentrators are dusty places, and expect to
see the signature of the ore in your trace-element data if the
samples are not clean.
In all but very coarse-grained rocks, samples of around
1 kg should be adequate (Potts, 1987). In fine-grained or
glassy volcanic rocks, samples in the range 200 to 500 g are
generally sufficient for major element analyses. However, for
low-level trace elements (<1 ppm) the 'nugget effect' may be
significant if elements of interest are concentrated in sparsely
disseminated grains (e.g. zircon). Potts (1987) provided illus-
trative tables of the sample weights required for analytical
precision, at various concentrations and grainsizes. These
represent the quantities of sample that actually undergo analy-
sis and should not be confused with the amount of pulverised
sample from which smaller portions are taken for analysis.
Sample preparation
All whole-rock analytical techniques (except neutron activa-
tion analysis) require samples to be crushed and ground to
fine powder for direct analysis, acid digestion or fusion. There
are opportunities here for sample contaminations from wear
on the grinding machinery, cross contamination between
samples in a batch, and contamination from previous batches
of samples put through the mill. Reputable laboratories will
use routine procedures to minimise those problems. However,
if low-level trace elements are important, it may help to inform
the laboratory personnel and request extra care.
Jenner (1998) cautioned against the use of new diamond
saw blades and recycled coolant for trimming slabs. He
suggested that if sawn rock slabs must be analysed, they should
be washed in acetone and distilled water. If the rock samples
include overprinting phases, such as veins and weathered
rinds, these are best cut out before crushing. The practice
of crushing to -10 mm and attempting to pick out the
rock chips without extraneous materials does not eliminate
contamination.
Jaw crushers are infamous contaminators. Crevices
between their moving parts harbour dust that causes cross
contamination, and the wear of plates directly contributes steel
particles. Magnetic separation of steel particles is inadvisable
because it may lead to selective removal of magnetic minerals.
A better alternative is to crush rock samples down to ^5 mm
fragments in a hydraulic press fitted with tungsten carbide
plates. The crushed rock, including any fine dust, can then be
coned and quartered down to about 80 or 100 g for pulverising
to a final grainsize of less than 75 (Am in a swing mill.
Materials used in swing mills include tungsten carbide,
chrome steel, alumina ceramic, and agate. All of these are
potential contaminators of some elements. Chrome steel is
the least expensive, but seriously compromises Cr and low-
level Fe analyses. Tungsten carbide is popular and durable but
will contribute Co and W to the sample, and possibly also Ti,
Ta and Nb (Potts, 1987). Agate mills are fragile, expensive and
slow but are the least likely to contaminate samples destined
for low-level trace element determination.
Precision, accuracy and reference materials
Precision is a measure of the repeatability of data within an
analytical session, and reproducibility in different analytical
sessions over longer intervals. Analysing several duplicate
samples within batches and subsequent batches provides an
estimate of the precision of a technique. Precision is reported
as the co-efficient of variation (CV), which is equivalent to
relative standard deviation (RSD) expressed as a percentage
of the mean.
where C
r
' is the true or accepted proportion of element i in
the reference material, and C
a
]
is the apparent or analysed
proportion of element i in the reference material.
Reputable analytical laboratories will routinely run refer-
ence materials to calibrate their instruments and maintain an
acceptable level of accuracy. It is helpful if the client, upon
submittal, gives the analyst an indication of the rock types,
mineral assemblages and compositional range of the samples,
to enable an appropriate selection of reference materials.
However, laboratories do not routinely report their analyses
of reference materials for comparison with accepted values.
Even if they do, the potential for fudging the results provides
the user with a less than satisfactory assurance of accuracy.
The solution is for the user to acquire some appropriate
geochemical reference materials and include a sample or two,
preferably disguised, in each analytical batch. This can involve
considerable expense. Potts (1987) suggested that a minimum
of 10 reference materials should be submitted with each batch
for a full assessment of accuracy.
In many cases, the user may not be greatly concerned
with the 'correct' values provided that the data are relatively
consistent or precise within and between batches. Submitting
GEOCHEMISTRY OF ALTERERD ROCKS | 75
representative duplicate samples allows the user to monitor
long-term precision and calibration drift. It permits greater
confidence in the analytical data, which justifies the additional
expense. Keeping the reference materials and duplicates
anonymous can be problematic, even if the sample numbers
are disguised. A few small packets of powdered or crushed rock
stand out prominently in a batch of drill-core samples and
analysts have no difficulty in recognising them as standards.
They may analyse them with special care, develop their own
comparative data over successive batches, and be tempted to
adjust any outliers. However, the recognition that the client
is prepared to carry out independent quality control generally
has a positive effect on analytical practice and discourages
complacency at the laboratory.
Limit of detection
Limit of detection of an element is commonly understood to be
'the lowest concentration that can be confidently measured by
a particular method on an average sample' (e.g. Anonymous,
1997). However, as pointed out in detail by Potts (1987), the
levels of confidence are frequently not stated. That obscures
the reality that the quoted detection limits are often below
the levels at which reliably quantitative measurements are
possible.
Potts (1987) proposed three new terms for better definition
of the much abused 'detection limit':
Lower limit of detection (LLD) for a signal level of three
standard deviations higher than the mean background
(mean + 3s). This is the lowest level at which an element
can be recognised but not quantitatively estimated.
Limit of determination (LOD) representing a level six
standard deviations above the background signal (mean
+ 6s). It is the lowest level at which the signal can be
quantitatively measured for a confident analysis.
Limit of quantitation (LOQ) is set at 10 standard deviations
higher than background (mean + 10s) to provide extra
confidence in the analysis in cases where there are legal,
commercial or statutory implications placed on the
interpretation of detection limits.
According to Jenner (1998) the limit of detection
commonly quoted by analysts is equivalent to Pott's (1987)
lower limit of detection (LLD; mean + 3s). Jenner (1998)
gave the example that if LLD for an element is 0.3 ppm,
then data reported between 0.3 and 1 ppm may or may not
be significantly different, but the element is recognisably
present. Data above 1 ppm may be considered quantitative,
or in other words, data at 1.1 and 1.5 ppm can be confidently
regarded as different.
The consequence is that one cannot place much reliance
on data reported at close to the lower limit of detection. A
safe rule of thumb is to treat with circumspection any data of
less than one order of magnitude above the quoted detection
limit. Therefore, select an analytical method that will provide
quantitative data at an order of magnitude lower than the
threshold of interest.
where s> is the standard deviation of element i in the samples
analysed, and %
J
is the sample mean.
Accuracy is a measure of how close analytical data lie to the
'true' values. It may be evaluated by including in each analytical
batch some international geochemical reference materials, or
in-house standards for which the true values are relatively
well determined. It is important that such reference materials
have compositions near the compositional range of the rocks
being analysed (i.e. a standard rhyolite is inappropriate when
analysing a batch of basalts). For each element accuracy is
expressed as a positive or negative percentage difference,
relative to the true or accepted values of the standard.
7 6 | CHAPTER 4
Analytical techniques
The most popular methods for analysis of whole-rock samples
are X-ray fluorescence spectrometry (XRF), inductively coupled
plasma atomic emission spectrometry (ICP-AES) and inductively
coupled plasma mass spectrometry (ICP-MS). These methods
offer good precision for a large number of major and trace elements
over wide concentration ranges. XRF remains the preferred
method for major elements (Robinson, 2001). However, some
commercial laboratories have recendy converted to ICP-AES,
presumably for its rapid throughput and ability to measure most
major and trace elements. ICP-MS is used only for trace element
determinations. Neutron activation analysis (NAA) provides
low limits of detection for REE, some platinum-group elements
and some high-field-strength elements but it requires access to
a nuclear reactor, produces radioactive waste and has slow turn-
around of weeks or months.
The geochemical laboratory at the University of
Tasmania's School of Earth Sciences uses a combination of
XRF on flux-fused and powdered samples (for major elements
and moderate-abundance trace elements, respectively) and
solution or laser-ablation ICP-MS for low-abundance trace
elements (Robinson, 2001).
To obtain the appropriate quality of data, it is important
to involve the analysts in the selection of analytical methods.
Inform them of the approximate compositional range in the
samples and the geochemical objectives of the analyses.
X-ray fluorescence (XRF) can be used to analyse up to
about 60 elements with atomic numbers greater than 10
(Na upwards) at concentrations from 100% down to a few
parts per million (Rollinson, 1993). Detection limits for
trace elements in the range 0.5 to 2 ppm are achievable, and
precision for major elements approaches less than 1% RSD
(Robinson, 2001). However, routine procedures used in
commercial laboratories generally result in higher detection
limits of 2-10 ppm.
Optimum detection limits and precision for trace
elements, at concentrations below about 0.2%, are obtained
by analysis of 610 g undiluted rock powder pressed into a
pill. The rock powder must be ground to a grainsize of less
than 75 [im to ensure homogeneity in the sample.
Major element concentrations are determined on glass
discs made by fusing a small amount of powdered rock
diluted with lithium metaborate and tetraborate fluxes. The
fusion produces a homogenous glass, which enables analysis
of the light major elements and minimises X-ray absorption
and enhancement matrix effects. The composition of the
rock influences the type and dilution factor of the flux to
be used, especially in sulfide-, base-metal- and carbonate-
bearing samples. Accordingly, at the geochemical laboratory
in the University of Tasmania's School of Earth Sciences,
approximate proportions of S, Fe, Ca, Ba, Cu, Pb and Zn
are first determined on the powder pills, along with the trace
elements, to enable appropriate selection of fusion fluxes.
In sulfide-bearing samples, fusion should be a two-stage
process with LiNO
3
in the flux to oxidise sulfur. Initially, the
carefully weighed sample-flux mixture is heated and held for
10 minutes at 700C to ensure that the sulfur is retained as
sulfate, and not evolved. It is then heated to 1000C for a
further 10 minutes to complete the fusion and the melt is cast
into a 32-mm diameter glass disc.
In the past decade, inductively coupled plasma (ICP)
analysis has revolutionised geochemical analysis, particularly
for trace elements. There are two separate methods of analysis
known as ICP-AES and ICP-MS. Both use inductively
coupled high-temperature argon plasma to generate atomic
and ionic emissions in the sample. In ICP-AES (atomic
emission spectrometry) the spectrum of atomic emissions is
measured by an array of photomultipliers. This method is
sometimes called ICP-OES (optical emission spectrometry).
ICP-MS uses a mass spectrometer to measure ionic particles
in plasma-sample gases. ICP-AES provides low limits of
detection typically 210 ppm for trace elements and 10
100 ppm for major elements. ICP-MS enables determinations
of heavier trace elements at extreme detection limits, up to
four orders of magnitude lower than ICP-AES (e.g. REE and
HFSE detection limits in the range of 0.1 to 2 ppb).
ICP methods are able to measure most elements at low
detection limits with high precision using linear calibration
over eight orders of magnitude (Robinson, 2001). Up to 50
elements can be analysed simultaneously in a few minutes on
samples of less than 100 mg.
There are a number of disadvantages to ICP related to
the requirement that the rock sample must be dissolved in
dilute solution before introduction to the plasma. Ensuring
complete dissolution of rocks, including refractory phases
such as zircon and REE minerals, and maintaining them in
solution without contamination, is a difficult task (Yu et al.,
2001). Samples are typically digested by strong acid cocktails
in sealed teflon vials for one or two days, dried by evaporation
and then the residue is re-dissolved in dilute nitric acid ready
for analysis. In some cases, samples are fused with lithium-
borate fluxes or sodium peroxide before acid dissolution. It
is advisable to test for complete solution by comparing the
ICP-MS data with XRF determinations of some relatively
immobile elements, such as Zr, Nb, Y and REE (Robinson,
2001). The likely loss of some volatile elements during fusion
and evaporation may render the method unsuitable for Hg,
Tl, Sb, Se and As.
H
2
O and CO
2
A combination of XRF and ICP-AES or ICP-MS can provide
accurate analyses of most major elements and a surfeit of
trace elements. However, they cannot determine H
2
O and
CO
2
as XRF is limited to elements of atomic numbers greater
than 10, and ICP, which measures the samples in solution,
obviously cannot determine water. Loss on ignition (LOI) is a
poor substitute for H
2
O and CO
2
analyses. For geochemical
analyses of altered rocks that contain significant hydrous
minerals or carbonates, it is preferable to separately determine
H
2
O and CO
2
.
Hydrogen and total carbon in rock samples can be analysed
by the 'C-H-N elemental analyser', an instrument designed for
routine determinations of C, H and N in organic compounds
(Potts, 1987). It is a relatively straightforward and inexpensive
technique requiring only about 25 mg of finely ground rock
powder. On the, generally safe, assumption that altered
volcanic rocks do not contain significant organic substances,
the total H and C determinations can be recalculated to H
2
O
and CO
2
for inclusion in major element composition data.
Interference
Some instrumental analytical methods are susceptible to
inaccuracies caused by peak overlap, particularly in mineralised
rocks with exotic-metal contents. For instance, in XRF analyses
on powdered samples, high concentrations of Ba and Pb
interfere with determination of Ti and Zr, respectively. These
problems may be minimised by instrument configuration or,
in extreme cases, another method of analysis. Therefore, it
is important that the analyst is advised which elements are
critically important for the geochemical interpretation and
the approximate compositional ranges of the rocks submitted,
so that the appropriate analytical methods and procedures are
applied. For example, if the purpose of analyses is to infer
the volcanic precursors of Pb + Zn + Ba-bearing mineralised
rocks by immobile element chemostratigraphy, it would
help the analyst to know that accurate Ti and Zr analyses are
priorities, and that the mineral assemblage includes Pb- and
Zn-sulfides, and barite.
Reporting data
Major elements in rocks are conventionally reported in weight
percent (wt%), mostly as oxides and in order of decreasing
cation valency: SiO
2
, TiO
2
, A1
2
O
3
, FeO (or Fe
2
O
3
total),
MnO, MgO, CaO, Na
2
O, K
2
O, P
2
O
5
, H
2
O
+
, CO
2
, S and
Total. The percentages are relative to the dry sample; analysed
after driving off moisture (H
2
O~) by heating for a few hours
at 105C. H
2
O
+
in the analysis represents structural water in
crystals or glass. All other elements are typically regarded as
trace elements although some may be present at greater than
the conventional 0. 1% cut off. Trace elements are usually
reported in alphabetic order in parts per million (ppm),
which is equivalent to the SI unit |ig/g preferred by analysts.
It is often useful to re-order the trace elements into groups
according to their geochemical characteristics (e.g. immobile
trace elements, rare earths, etc.) to facilitate plotting. It is
important to note that geochemical analyses are usually
reported to one decimal place more than can be quoted with
confidence (Potts, 1987).
Loss on ignition
Hydrogen and carbon, mainly as H
2
O and carbonate, are
significant components of some silicate rocks, particularly
altered rocks. However, since XRF or ICP methods cannot
determine these light elements, it has become common to
report loss on ignition (LOI) as a proxy. LOI is determined
by igniting a weighed sample at 1000-1050C for at least
12 hours and then weighing the residue. Expressed in weight
percent (as for the major elements), it is often assumed that
LOI represents the combined proportions of H
2
O
+
and CO
2
in the rock. However, for several reasons pointed out by Potts
(1987), the determination of LOI has little geochemical
value:
The maximum temperature may not be high enough to
dehydrate some minerals, such as talc, topaz, cordierite
and epidote.
Oxidation of ferrous iron (e.g. in silicates, magnetite and
GEOCHEMISTRY OF ALTERERD ROCKS | 7 7
sulfides) can produce a weight gain, partly offsetting the
LOI. Even if FeO and Fe are separately determined, it
is not reasonable to assume that they will be completely
oxidised in the LOI process.
LOI may include a contribution from volatilisation of
alkali metals, sulfur oxides and fluorine; the loss may be
only partial and not predictable. For example, in samples
containing both sulfide and carbonate, some SO
2
may
react with CaO and remain in the ignited sample.
The variables affecting LOI and the wide range of
compositions in altered and mineralised rocks may lead
to unpredictable results that are not readily interpretable.
Consequently, LOI is not a reliable substitute for H
2
O
+
and
CO
2
analyses, and it may significantly underestimate the
evolved volatiles.
Totals
The sum of the elements in a major-element analysis is
frequently taken as an indication of analytical reliability.
Considering the shortcomings of LOI determinations, it is
unreasonable to expect that their inclusion in a major element
analyses will provide totals close to 100%. Nor should one
expect that the error values on determinations of individual
elements should cancel each other to produce totals of 100%.
Further ambiguity is due to the usual practice of analysing
and reporting total Fe as Fe
2
O
3
, irrespective of its actual
oxidation state.
Sums of XRF major element oxides, sulfur and LOI data
will exceed 100%. This is because the sulfur is measured
twice: in the fused disc XRF and in the LOI determinations.
However, simple subtraction of XRF determined sulfur does
not solve the problem because the sulfur may not be entirely
evolved in the LOI process. Reasonable totals are obtained if
H
2
O
+
and CO
2
are separately determined by another method
and summed with XRF major element oxides and sulfur.
Low totals do not necessarily indicate erroneous analyses;
it may be that some significant elements were not determined.
For example, boron constitutes about 10% of tourmalines and
some micas contain significant proportions of Li, Fl, Rb and
Ce (Deer et al., 1966), and mineralised rocks may contain
significant proportions of base metals.
Potts (1987) cautioned that, although it is nice to find
totals near 100%, it is not a satisfactory test of the quality
of analytical data. He considered that the only acceptable
way of checking accuracy of modern instrumental analytical
techniques, such as XRF and ICP, was to analyse appropriate
geochemical reference materials. By applying this quality
control practice to all of the methods used to produce a set of
major-element data, the user can have reasonable faith in the
relative accuracy of the individual elements analysed. Likewise,
confidence in the accuracy of individual determinations
permits acceptance of odd totals as accumulated errors, or
indications of incomplete analyses.
Recalculating to volatile free
It is common practice, particularly in petrological literature,
to recalculate major-element analyses to 100% 'anhydrous'
7 8 | CHAPTER 4
or 'volatile free'. The recalculation involves multiplying the
proportion of each major-element oxide in an analysis, except
LOI, by a factor derived from the formula 100/2 Q, where 2
C| is the weight percent sum of all major oxides (not including
LOI) in the analysis.
Thus the 10 major oxides are commonly adjusted to sum
to exactly 100%. LOI is reported, but not included in the
total (e.g. Crawford et al., 1992; Stolz, 1995). The object
of recalculation is to remove apparent variations in the
proportions of major oxides that may be due to differences
in LOI values. This may be a valid approach for studies of
petrogenesis and fractionation in unaltered or weakly altered
rocks, where alteration was limited to hydration or minor
calcite vesicle fillings (Crawford et al., 1992). In these cases,
additions of H
2
O
+
and CO
2
may be the only significant
metasomatic changes, although CaO may be added if it is
not otherwise derived from the decalcification of plagioclase.
It is not unreasonable to 'subtract' the estimated additions
due to hydration and carbonation to determine equivalent
anhydrous magmatic compositions.
In more altered rocks, however, many of the other major
elements, particularly Si, Fe, Mg, Ca, Na and S, are also likely
to be involved in significant mass changes. Thus recalculating
to volatile free would distort the pattern of mass changes,
artificially increasing the changes of other elements relative to
the constituents of LOI. In cases where the 10 major oxides
(SiO
2
to P2O
5
) sum to low totals, volatile free recalculation
followed by immobile-element-based mass change estimates
could result in actual small net mass losses of some elements
appearing as mass gains. Similarly, it may upwardly distort
the estimates of net mass change, with implications for
interpretation of volume changes.
Barrett and MacLean (1994a) recommended using volatile
free recalculated major element data for mass change estimates.
They neglected to mention whether the same recalculation
factor should also, for consistency, be applied to all the trace
elements. Failure to similarly adjust the trace elements could
lead to small inconsistencies in immobile element ratios
(e.g. Ti/Zr, which is usually calculated from major element
TiO
2
data). It could also positively skew subsequent mass
change estimates based on immobile elements. The volatile
free recalculation has a proportionally greater effect on data
for altered rocks because they generally have higher LOI
than their unaltered precursors. Upward recalculation of all
trace elements, as well as majors, can avoid those problems.
However, comprehensive recalculation is unnecessary as both
chemostratigraphic methods and mass change estimates are
based on ratios of elements and hence are unaffected by
recalculation.
The method used in some studies (e.g. Gemmell and
Large, 1992), of recalculating the 10 major oxides but not
other major elements, such as sulfur, is also inconsistent. It
has the effect of under-estimating mass changes in sulfur
relative to the other oxides.
In dealing with altered rocks, it is preferable to obtain as
near as practical 'complete' major element analyses. The major
element suite should include sulfur, H
2
O
+
and CO
2
and any
other elements likely to exceed 0.1% (e.g. base metals). Ensure
adequate quality controls and then trust in the accuracy of the
data. They do not require fudging or normalisation at the risk
of introducing new errors and misinterpretations. Otherwise,
where analyses of altered rocks are limited to the usual 10
major oxides and LOI, it is best to treat LOI as a (somewhat
fuzzy) component in its own right, and not to recalculate to
artificial 100% totals. This particularly applies to data that are
to be used in mass transfer calculations.
Closure in composition data
Closure, also known as the constant sum effect, affects all
analytical data expressed as proportions of a whole; that is,
as composition data (or as 'concentrations' in Stanley and
Madeisky, 1996). The total of all elements in a rock analysis
must sum to 100%, or one million ppm, and so forth,
according to the units of measurement. Chemical mass
transfers that change the total mass of a system by adding
or removing some elements (e.g. hydrothermal alteration)
will affect the proportions of all elements, even those not
involved in the mass transfers (Fig. 4.1). Consequently, the
apparent differences in composition data between unaltered
and altered rocks do not accurately reflect the real material
changes, except in systems where there has been no net mass
changes
Closure particularly obscures mass changes in the major
elements such as Si, Al and Fe, which exist in high proportions
in primary unaltered rocks. For example, intense silicification
of felsic volcanic rocks may not be apparent in composition
data (e.g. footwall alteration zone of the Thalanga deposit,
Herrmann and Hill, 2001). Closure is less of a problem for
elements of low initial proportions. For example, in felsic
volcanic rocks, Ca and Na concentrations average only a few
weight percent and their depletion associated with plagioclase-
destructive alteration is usually evident in the composition
data.
Trace elements, by definition, are present at low
proportions in background rocks. They tend to provide
high-contrast anomalies in mineralised and altered rocks,
commonly orders of magnitude greater than background
levels, and are therefore practically unaffected by closure.
FIGURE 4.1 | Schematic illustration of the effect of closure in composition
data (after Eilu et al., 1997). Alteration leading to a net mass gain in the system
results in a lower proportion (or concentration) of element i, although the actual
mass of element i is not changed (AM, = 0). M
p
= initial mass and M
A
= altered
mass of total system.
GEOCHEMISTRY OF ALTERERD ROCKS I 7 9
Rollinson (1993) presented a detailed discussion of the
problem of closure in geochemistry and suggested some
solutions. It is not a trivial matter in the study of altered rocks.
The techniques for estimating alteration-related mass transfers
overcome the problem of closure by quantifying the amounts
of elements or oxides gained by or lost from altered rocks.
Chemostratigraphy
Major element compositions are routinely used to classify
volcanic rocks in terms of petrogenesis and tectonic setting
(e.g. Pearce and Cann, 1973). However, the same method is
not applicable to altered rocks because many of the major
elements, especially Si, Fe, Mg, Ca, Na and K, are relatively
mobile during alteration. Consequently, compositional
changes related to alteration may considerably outweigh
their primary variations. Fortunately, several elements are
chemically immobile during most types of alteration and these
can be reliably used to classify and correlate altered volcanic
rocks. In this context, immobile means elements that are
neither added to, nor taken from, the rock during alteration.
Immobile elements may be involved in phase changes and
perhaps be mobile at millimetre scale, but their mass in the
altered rock remains unchanged. Although the proportions
(concentrations) of immobile elements may change, due to
net mass changes in the size of the system, their inter-element
ratios remain the same.
Immobile elements
The high-field-strength elements Ti, Zr, Nb and Y are
relatively immobile during hydrothermal, diagenetic and
weathering alteration, and during regional metamorphism up
to mid-amphibolite facies. Ratios of these immobile elements
are the basis of tectono-magmatic discrimination diagrams
developed in the 1970s (e.g. Pearce and Cann, 1973; Floyd
and Winchester, 1978).
Many studies of VHMS deposits have shown that Al, Ti,
Zr, Nb, Y, heavy REE (Lu, Yb), Hf, Ta and Th, and in some
cases P, Sc, V and Cr, remain essentially immobile during
alteration. Their immobility has been documented even in the
most intense hydrothermally altered zones directly beneath
deposits (e.g. MacLean and Kranidiotis, 1987; Skirrow and
Franklin, 1994; Barrett and MacLean, 1994a). Barrett and
MacLean (1994a) recognised some mobility of the light REE
in proximal, intense chlorite altered zones beneath some
deposits and suggested that they may be useful as exploration
vectors. Y and Nb show considerable scatter in some datasets
and this may be partly attributable to primary variations
(Ewart, 1979) or slight chemical mobility in some systems.
Analytical precision may also be a factor, particularly for Nb,
which typically occurs at low concentrations not much above
XRF detection limits.
In practice, Ti and Zr are the most reliably immobile
elements. They can be inexpensively and accurately analysed
by XRF on pressed powder pellets and they exist at easily
detectable levels in most volcanic rocks, unlike the heavy
REE, Sc, Nb, Ta, Hf and Th, which generally exist at less
than 20 ppm.
Incompatible elements
Incompatible elements are those that tend to be excluded from
the lattices of minerals crystallising from magmas and are
instead partitioned into the melt phase. Hence, incompatible
elements exist at highest proportions in the most evolved
felsic rocks.
The high-field-strength elements (HFSE) Zr, Y and Nb
are generally incompatible, except in some calc-alkaline suites.
They have similar low magmatic liquid-solid distribution co-
efficients and so tend to retain similar inter-element ratios
throughout a single magmatic fractionation series, and on x-y
bivariate plots form linear trends that project from the zero
origin. Subsequent alteration involving net mass gain or loss
can change the proportion of incompatible elements in the
whole rock but their primary ratios are preserved. As a result
alteration trends coincide with the primary fractionation
trends.
The gradients of these trends vary according to magmatic
affinity (MacLean and Barrett, 1993; Barrett and MacLean,
1994a). Samples from different magmatic suites thus produce
separate linear trends of magmatic enrichment, which regress
to the origins on incompatible-incompatible immobile
element plots (e.g. Fig. 4.2).
Incompatible-incompatible element ratios and bivariate
plots have chemostratigraphic applications even in
hydro thermally altered samples in which the major elements
may not be reliable discriminants. Incompatible-incompatible
element ratios are used to identify magmatic affinities,
favourable volcanic suites and terranes.
Compatible elements
Compatible elements have high magmatic distribution co-
efficients (>1) and are preferentially taken up by mineral
FIGURE 4.2 | Schematic Y-Zr (incompatible-incompatible) plot used for
determination of magmatic affinities in altered volcanic rocks (modified after
MacLean and Barrett, 1993).
8 0 | CHAPTER 4
phases crystallising from magma. Consequently, compatible
elements are depleted from the melt phase (this is opposite
to the enrichment in incompatible elements). Thus, the
relative proportion of compatible and incompatible elements
in residual melts changes as fractionation proceeds. Batches
of magma that are successively tapped off and emplaced as
eruptive or intrusive units will have successively smaller
compatible-incompatible element ratios.
Aluminium, Ti, Cr, Sc and V are generally compatible
during crystallisation and immobile during alteration
(Barrett and MacLean, 1994a). Bivariate plots of immobile
compatible-incompatible element data for least-altered
samples from a particular magmatic affinity should show
smooth fractionation trends, generally with negative slopes
(Fig. 4.3).
Subsequent net mass gains and losses imposed by alteration
will produce, on compatible-incompatible immobile element
plots, separate alteration lines for each chemically distinct
rock unit (Fig. 4.3). This property is particularly useful for
the discrimination and correlation of initially homogenous
but subsequently altered volcanic units that may be otherwise
unidentifiable.
Chemostratigraphic fingerprinting is commonly done
with Ti/Zr ratios and TiO
2
Zr bivariate plots. However,
in tholeiitic suites Ti enrichment parallels the typical Fe-
enrichment trend at the mafic end of the fractionation
series (MacLean and Barrett, 1993). In other words, Ti is
initially incompatible in tholeiitic fractionation series up
to about the composition of basaltic-andesite. On a TiO
2

Zr plot, the mafic end of the tholeiitic series has a positive


fractionation trend, essentially similar to the alteration trends
superimposed by subsequent metasomatic net mass changes
(Fig. 4.3). Therefore, TiO
2
-Zr is not a reliable discriminant
of altered mafic tholeiites. A reasonable substitute is A1
2
O
3

Zr, which has a near linear, slightly negative trend in tholeiites


(MacLean and Barrett, 1993). Titanium becomes compatible
in tholeiitic magmas more evolved than basaltic-andesite and
the TiO
2
-Zr fractionation curve then has a negative slope.
Two examples of chemostratigraphic discrimination and
correlation in the Mount Read Volcanics are presented in
Figures 4.4 and 4.5.
FIGURE 4.3 | Schematic TiO2-Zr (compatible-incompatible) plot showing the
negative curvilinear fractionation trend typical of co-genetic calc-alkaline volcanic
suites (after Barrett and MacLean, 1994a). Mafic tholeiites may show a positive
trend up to the composition of basaltic-andesite, because of TiO2 incompatibility
in the early stage of magmatic differentiation.
It is worth emphasising that immobile element ratios must
be used with discretion in chemostratigraphic discrimination
and correlation. This method is most effective in rocks
with primary compositional homogeneity, such as coherent
lavas and sills, and possibly in some massive syneruptive
volcaniclastic units such as pumice breccias. Processes
of magma generation, crystallisation and fractionation
determine the immobile element ratios of magmas. Therefore,
immobile element ratios are likely to be uniform within single
coherent eruptive or intrusive emplacement units. However,
volcaniclastic debris may be subject to unhomogenising
processes. Both lateral and vertical compositional variations
may occur in volcaniclastic units as a result of:
mechanical sorting of components of different densities,
such as clasts, pumice, scoria, crystals (e.g. Ti-oxides and
zircon) and glass shards, during eruption and transport
winnowing of glass shards from turbidity currents
mixing of debris from a variety of volcanic sources in
variable proportions
or
incorporation of extraneous clasts into the base of
volcaniclastic mass flows.
Gifkins (2001) showed that thick, graded, rhyolitic
pumice breccia units in the Central Volcanic Complex,
western Tasmania, have Ti/Zr ratios that vary from -5 near
the crystal- and lithic-rich bases, to -9 in the fine-grained,
shard-rich tops of the units. This is consistent with an
increased concentration of zircon crystals towards the base of
the units but may also reflect the abundance of felsic clasts
in the basal portions. In contrast, graded rhyolitic volcanic
breccia units in the Rosebery hanging wall, western Tasmania,
display the opposite trend. Large et al. (2001b) suggested that
decreasing Ti/Zr ratios towards the top of these units were
due to physical fractionation of Zr-poor crystal and lithic
components from Zr-bearing originally glassy pumice and
shards during emplacement.
Testing immobility
It is preferable that element immobility in any system is
established, rather than assumed, before proceeding with
chemostratigraphic interpretations of altered rocks. This is
also an important preparatory step for methods of estimating
mass transfers of mobile elements.
The simplest test is to plot potentially immobile elements
on x-y bivariate diagrams with the origins at zero. If possible,
the tests should use data from unaltered and variably altered
samples of a single, originally compositionally uniform volcanic
unit, such as a coherent lava or sill. If the selected elements
are immobile, the data points for a single-precursor system
should align in a highly correlated linear trend, which projects
to the origin of the plot and through the data points of the
least-altered samples. These linear trends, or alteration lines,
are due to net mass gains and losses of the mobile elements
in the altered rock samples (Fig. 4.6). Typically there is some
data scatter due to analytical errors and slight inhomogeneities
in the primary rock. However, if both elements are immobile,
calculated linear correlation co-efficients (r) for alteration
lines should exceed -0.85 (Barrett and MacLean, 1994a).
In contrast, elements that were mobile during alteration are
GEOCHEMISTRY OF ALTERERD ROCKS | 81
0 10 20 30 40 50 0 25 50 75 100
Ti/Zr Alt erat i on Index
100(K2O+MgO)
(K 2 0+Mg0+Ca0+N a2 0)
FIGURE 4.4 | Graphic log and down-hole Ti/Zr and Al plots of drill hole NC4, near Lake Newton, western Tasmania. The Ti/Zr
data clearly delineates units of different primary volcanic compositions despite effects of strong hydrothermal alteration in rocks
intersected in the middle and lower part of the hole. The high Ti/Zr ratios at -2 00-2 30 m led to recognition of an altered mafic
volcanic breccia unit, which had previously been interpreted as a zone of chlorite-altered felsic volcaniclastic rocks. Another
altered mafic unit occurs below 530 m. Quartz and feldspar crystal-rich sandstones in the upper 100 m have a range of Ti/Z
ratios, which suggests that they do not have a unique provenance.
readily identifiable by their erratic distribution or near total
removal (Finlow-Bates and Stumpfl, 1981).
Mass transfer techniques
Mass transfer techniques aim to quantify the amounts of
individual elements added to and subtracted from the rock
during alteration in order to overcome the distortions of
closure that are inherent in composition data.
As noted by Barrett and MacLean (1994b), significant
mass change anomalies may not be apparent in untreated
compositional data, due to closure. The results of mass
transfer calculations are usually easy to relate to mineral
assemblages and may reveal clues about the composition,
source and temperature of hydrothermal fluids (e.g. Barrett
and MacLean, 1994b). Mass change data have been used
to infer hydrothermal water-rock ratios (e.g. MacLean and
Hoy, 1991). They are also used in the interpretation of
whole-rock 6
18
O and REE data (e.g. MacLean and Barrett,
1993). Thus, they may enable discrimination of favourable
alteration systems and altered zones within systems. When
plotted spatially mass transfer data can be used as quantitative
exploration vectors (Section 8.2).
82 I CHAPTER 4
FIGURE 4.5 | Chemostratigraphic correlation diagram of the volcano-sedimentary succession that hosts the Rosebery massive sulfide deposit, western Tasmania.
A thin unit of dacitic pumice breccia, between the massive sulfide lens and the feldspar + quartz + biotite porphyry sill intersected in hole 120R, is texturally
indistinguishable from the footwall rhyolitic pumice breccias (Ti/Zr = 7-9) but has a distinctive Ti/Zr ratio of between 12 and 14.
FIGURE 4.6 | TiO2-Zr plot of data from a tholeiitic volcanic
suite (after MacLean and Barrett, 1993). The data points for
least-altered basalt, andesite and rhyolite samples define
the magmatic 'fractionation curve'. Two linear arrays of data
represent variably altered samples of originally homogenous
units of andesite and rhyolite. These highly correlated
alteration lines intersect the least-altered data points on
the fractionation curve and project towards the origin. Data
points on the alteration lines below the fractionation curve
represent net mass gains, those above the curve represent
net mass losses. The positive slope at the basaltic end of the
fractionation curve is due to incompatibility of TiO2 in the early
stages of differentiation of tholeiitic magmas. This may lead
to confusion of fractionation and alteration trends in mafic
tholeiites.
There are several approaches to estimating mass transfers.
These include the mathematically complex graphical
method of Gresens (1967); subsequently simplified by Grant
(1986) and Huston (1993) to produce the isocon method;
the immobile element techniques of MacLean and Barrett
(1993), and the Pearce element ratio analysis of Stanley and
Madeisky (1996). These techniques all depend on recognition
of immobile elements. They also, with the exception of the
Pearce element ratio analysis, depend on the identification of
precursor-rock compositions.
The determination of unaltered precursor compositions is
often problematic in the altered, lithologically and structurally
complex rocks that host ore deposits. The safest approach
is to consider the geological context, established from field
relationships and rock textures, in combination with various
immobile element tests (e.g. MacLean and Barrett, 1993;
Stanley and Madeisky, 1996). This ensures that precursors are
appropriately matched to the altered rocks under investigation.
The isocon method does not include a procedure for selecting
precursors and it commonly produces erroneous results
because incorrect geological assumptions are applied. For this
reason, we prefer the more rigorous MacLean and Barrett
(1993) method, which is also simplest to calculate.
Another potential difficulty arises from primary compo-
sitional variations in volcanic rocks related to magmatic
fractionation or volcaniclastic mixing processes. Huston
(1993) recognised this as a deficiency of the isocon method.
He suggested examining standard deviations of data from
least-altered samples to screen their suitability, estimating
errors for the consequent mass changes and recognising that
small apparent mass changes may be artefacts of primary
inhomogeneities, not due to alteration. The Pearce element
ratio analysis method (Stanley and Madeisky, 1996) and the
multiple-precursor variant of the MacLean and Barrett (1993)
method attempt to overcome the limitations of primary
variability. Of these two, we prefer the MacLean and Barrett
approach, again for its relative ease of calculation.
GEOCHEMISTRY OF ALTERERD ROCKS | 8 3
The MacLean and Barrett multiple-precursor method
involves sorting samples into affinity groups, calculating
primary variability trends (fractionation curves) and using
immobile elements to synthesise primary compositions for
altered samples that may fall between compositions of the
available least-altered samples. This method relies on the
assumptions that co-genetic groupings can be recognised
and separated, and compositional variations are smooth
linear or curvilinear trends for every major element. This
is generally true for magmatic fractionation, and therefore
applicable to volcanic suites, but the variations may be erratic
if volcaniclastic, sedimentary or multiple alteration processes
were involved (Eilu et al., 1997). Thus, due caution must
be exercised when using this method in mixed provenance
volcano-sedimentary successions. Filtering sample sets
through geological field evidence and petrographic textural
observations are obvious fundamental precautions. Sample
sets from restricted prospect-scale areas are less likely to have
complex primary compositional variations than district-scale
sample sets. Peculiar aberrations or significant mass changes
in elements otherwise expected to be immobile, such as Al,
should arouse suspicion that primary fractionation trends
may have been incorrectly modelled.
Single precursor mass transfer technique
The MacLean and Barrett method of estimating mass transfers
in single-precursor systems incorporates testing of immobility
as a fundamental step. The analytical data for potentially
immobile elements, from an initially homogenous but
variably altered lithological unit, are plotted on x-y bivariate
diagrams and linear regressions are calculated. The existence
of highly correlated (r >0.85) trends that pass through the
origin enables selection of the optimal element to be used
as the immobile monitor in the mass change calculations.
This graphical process also highlights any outliers or samples
FIGURE 4.7 | Diagram representing the calculation of
mass transfers for an altered rock with an initial mass of
100 units (after MacLean and Barrett, 1993). The proportion
of immobile element in the altered rock decreases (is
diluted) because of net mass gains in mobile element(s).
The ratio of immobile elements Z/Z
a
represents the factor,
which multiplied by the proportion of immobile element in
the altered rock, would restore it to a mass unit, rather than
a proportion of the whole. The same factor applied to all
other elements produces the 'reconstructed composition' in
mass units. The mass transfers in each element are then
calculated by subtracting its primary composition from the
reconstructed composition. In this example, a large mass
gain in X combined with a small mass loss in Y produces a
large net mass gain.
8 4 | CHAPTER 4
from different precursors, which should be eliminated or
treated separately. It identifies any primary compositional
inhomogeneities in the rocks, which may require treatment
by the multiple-precursor method (MacLean and Barrett,
1993) or Pearce element ratio analysis (Stanley and Madeisky,
1996). These more complex methods are not explained here;
the reader is referred to the relevant references for details.
The single-precursor mass transfer method proceeds
by calculating the ratios of the proportions of an immobile
element in the altered and unaltered samples. Each of the
mobile element proportions is then multiplied by that ratio
to obtain a reconstructed composition (Fig. 4.7). The mass
change of each element is found by subtracting its percent
proportion in the precursor from that in the reconstructed
composition.
The steps below and the flow chart in Figure 4.8 outline
the procedure for estimating mass changes in single-precursor
systems. Figures 4.9, 4.10 and 4.11 present a worked example
based on compositions of Thalanga footwall rhyolites,
Queensland.
(1) Acquire and tabulate the lithogeochemical analyses (e.g.
Figure 4.9). The first part of this section (4.1) presents
some guidance on sampling and analytical techniques.
Barrett and MacLean (1994a) recommended using major
element composition data that are recalculated to 'volatile-
free' totals of 100%. However, it is unreasonable to expect
Comme nt s related to
Thalanga data in Figures 4.9,
4.10 and 4 . 1 1 .
Values below detection replaced
by arbitrary small values; e.g.
0.5 x detection limit or zero
(Figures 4.9 and 4.10).
Data should preferably be from
mappable single-precursor
units. This example includes 34
rhyolites from diverse volcanic
units. However, the remarkable
uniformity of im m obile- el em ent
ratios suggests they were
essentially co-magmatic and
can be treated as a single-
precursor system.
No obvious outliers. Yttrium
(Fig. 4.10) typically shows
considerable scatter, which may
be a result of primary variation.
AI?O3 andZr both have three of
four correlation co-efficients
>0.85 (Fig. 4.10). Either could
be used as the immobile
monitor component. Zirconium
was selected because of
marginally higher average
correlation (0.86 vs 0.83).
Appropriate $ symbols in the
formula enable it to be filled
across and down in the
spreadsheet and maintain cell
references to Z, Z
a
and C.
Large positive and negative net
mass changes are mainly due to
SiO2 gains and losses, or
addition of Fe and S in pyritic
samples (Fig. 4.11), Note that
because of closure, the SiO2
changes are not obvious in the
analytical data.
FIGURE 4.8 | Flow chart for mass-change calculations by the single-precursor method (after MacLean and Barrett, 1993).
elements of an analysis to total exactly 100%, and the
use of recalculated data may produce positive distortions
in subsequent mass transfer estimates. In dealing with
altered rocks, it is preferable to obtain accurate analyses of
all the major elements (including S, CO, and H
2
O
+
) and
include them in the mass transfer calculations. If S, CO
2
and H
2
O
+
data are not available, and losses on ignition
(LOI) are significant (>2%), then LOI could be included
as a separate, albeit loosely denned, component of mass
change.
(2) Test for immobility. Plot analytical data for potentially
immobile elements (e.g. Al, Ti, Zr, Nb, Y, etc.) on x-y
bivariate diagrams with the origins at zero (e.g. Fig. 4.10).
If possible, use data from variably altered and unaltered
samples of a single uniform rock unit.
(3) Inspect for outliers and, with geological considerations,
decide whether to cull them, or treat the data as a multiple-
precursor system. Readers are referred to MacLean and
Barrett (1993) for details of the multiple-precursor
method.
(4) The immobile element data for single-precursor systems
should plot on highly correlated linear trends (r >0.85)
that pass through the origin and through the data points
for least-altered samples. Evaluate the bivariate diagrams
GEOCHEMISTRY OF ALTERERD ROCKS | 85
and correlation factors to determine which element
consistently occurs in highly correlated trends and is most
suitable as an immobile monitor (i.e. was uniform in the
primary rock and least mobile during alteration).
(5) Select a composition for the precursor. This could be from
a single unaltered sample or an average of several unaltered
samples.
(6) Calculate the absolute mass change for each component
using the formula:
A
a
= [Z7 Z
a
. C
a
] - C
where A
a
is absolute mass change expressed in g/100 g
C
a
= wt% proportion of component in altered rock
C = wt% proportion of component in precursor
Z
a
= proportion of immobile element in altered rock
Z = proportion of immobile element in precursor.
The mass changes may be calculated from compositions
of individual altered samples or from average compositions
of sample groups representing certain mineral assemblages
or altered zones.
(7) For visual comparison, plot the absolute mass changes for
individual elements on a bar graph (e.g. Fig. 4.11).
1
2
3
4
5
6
7
8
9
1 0
1 1
12
13
1 4
15
16
17
18
1 9
2 0
21
2 2
2 3
2 4
2 5
2 6
2 7
2 8
2 9
30
31
32
33
34
35
36
37
38
39
4 0
4 1
4 2
4 3
4 4
4 5
4 6
4 7
4 8
4 9
A B C
Whole-rock major and trace element composition data
Sample no.
* 140802
* 140727
* 140808
* 140724
* 140902
TH394-142
TH144B-34
TH41A-575
TH005-256
TH238-236
TH038-191
TH085-125
TH148-159
TH085A-422
TH270-278
TH270-145
TH085-312
TH018-26 6
TH038-054
TH085-159
TH085A-348
TH238-194
TH270-220
TH085A-335
TH085-204
TH085-215
TH085A-241
TH085-188
TH270-313
TH06 1-086
TH41A-713
TH06 1-157
TH270-381
TH085A-384
Precursor
composition
Mass Changes
140727
140808
140724
140902
Alteration facies Si O,
east-altered footwall 76 .40
-noderate, foliated sericite + chlorite 70.90
strong, pervasive quartz + sericite + pyrite chlorite 75.70
ntense, pervasive quartz + pyrite 6 7.00
ntense, microcrystalline quartz + K-feldspar 84.20
east-altered footwall 77.39
east-altered footwall 77.25
76 .71
77.80
74.6 2
77.81
75.25
82.02
6 9.97
6 4.37
72.70
72.73
6 6 .82
75.37
75.29
6 6 .93
74.41
6 3.38
78.04
71.22
6 7.08
71.59
74.54
72.56
74.39
75.87
78.48
55.36
49.30
average of 140802, TH394-142 and TH144B-34 77.01
{Formula in cell C451 ={$Q$38/$O5"C5)- ' .
1 C$38] is ?lled across and down. ' units h
\ Placement of $ symbols is critical to \ DL 0.05
i maintain cell references for Zo, Za and Co. \
\
V _ siOj
moderate, foliated sericite + chlorite ' > -15
strong, pervasive quartz + sericite + pyrite chlorite 7
ntense, pervasive quartz + pyrite 43
ntense, microcrystalline quartz + K-fe!dspar 72
units
D
TiO2
0.11
0.10
0.07
0.05
0.06
0.09
0.10
0.07
0.07
0.09
0.07
0.07
0.08
0.11
0.12
0.09
0.08
0.08
0.08
0.09
0.10
0.09
0.05
0.06
0.05
0.07
0.07
0.06
0.10
0.11
0.10
0.08
0.10
0.10
0.10
0.01
Ti O2
0
0
0
0
g/100q
E
A IA
11.90
14.50
11.40
6 .6 0
7.40
11.85
11.90
10.18
8.83
13.40
10.54
11.13
8.84
13.06
17.56
12.22
12.33
11.26
10.70
12.79
14.57
12.18
7.99
9.48
8.28
10.44
11.03
8.40
12.55
13.75
12.17
11.6 0
13.90
13.47
11.88
0.05
A IA
1
1
0
1
9/100g
F
F e A
1.6 4
1.6 9
5.14
14.34
0.6 0
1.13
2.09
0.99
1.90
1.40
1.51
2.32
0.6 8
2.80
4.73
2.39
5.48
2.46
4.47
1.96
5.70
1.42
13.46
4.45
9.18
8.6 2
6 .91
6 .95
4.93
1.56
1.70
0.88
14.36
19.14
1.6 2
0.01
F e A
0
4
24
-1
g/100g
G
MnO
0.04
0.04
0.08
0.02
0.00
0.02
0.03
0.08
0.06
0.02
0.09
0.06
0.02
0.03
0.07
0.09
0.01
0.13
0.02
0.09
0.08
0.05
0.13
0.04
0.09
0.13
0.06
0.05
0.07
0.03
0.05
0.01
0.14
0.30
0.03
0/
0.01
MnO
0
0
0
0
q/100q
H
MgO
0.6 7
3.94
2.38
1.6 7
0.08
0.76
0.6 8
4.33
2.14
3.54
3.13
1.90
1.05
1.98
4.35
4.24
1.10
9.88
2.91
2.51
4.70
4.16
5.88
1.76
3.97
6 .42
2.53
2.77
3.88
1.19
1.12
0.6 1
4.90
7.79
0.70
0.01
MgO
3
2
2
-1
q/100q
I
CaO
1.42
0.12
0.00
0.05
0.11
0.46
0.90
1.75
3.08
0.12
0.25
3.03
0.24
0.28
0.01
0.27
0.00
0.01
0.00
0.08
0.00
0.07
0.00
0.02
0.00
0.01
0.08
0.00
0.00
0.34
0.08
0.12
0.11
0.19
0.93
0.01
CaO
-1
-1
-1
-1
q/100q
J
Na2O
2.27
1.05
0.21
0.06
0.34
4.79
2.77
1.57
0.84
1.82
0.90
0.59
0.53
0.38
0.24
0.22
0.17
0.12
0.12
0.10
0.09
0.08
0.00
0.16
0.15
0.12
0.11
0.09
0.03
0.95
0.42
0.37
0.18
0.08
3.28
0.05
Na.O
-2
-3
3
-3
q/100g
K
K2O
4.04
3.92
2.39
1.99
5.87
1.82
3.6 0
2.80
3.06
2.23
3.40
3.6 8
5.33
3.6 9
4.15
4.73
3.75
5.08
2.43
4.41
2.91
3.57
0.91
2.45
1.51
1.59
2.82
2.14
2.50
6 .33
6 .36
7.6 0
2.40
0.93
3.15
0.01
K2O
0
-1
0
7
q/100q
L
PA
0.02
0.01
0.01
0.01
0.01
0.01
0.02
0.02
0.05
0.01
0.01
0.02
0.02
0.05
0.02
0.01
0.01
0.02
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.01
0.02
0.01
0.01
0.01
0.02
0.02
0.02
0.02
0.02
0.01
PA
0
0
0
0
g/100g
M
S
0.01
0.00
0.6 5
10.6 1
0.48
0.00
0.24
0.01
1.33
0.00
0.06
1.39
0.30
2.03
1.08
0.84
3.47
0.53
1.6 1
0.81
0.86
0.44
7.6 4
1.57
4.97
2.89
3.59
4.36
0.44
0.00
0.6 6
0.13
5.92
5.12
0.08
0/
0.01
s
0
1
19
1
q/100q
N
COj
0.00
0.10
0.00
0.00
0.15
0.15
0.99
0.00
0.26
0.00
0.18
0.10
0/
0.1
co
2
0
0
0
0
q/100q
O
Total
98.52
96 .37
98.03
102.40
99.15
98.47
99.74
98.51
100.15
97.25
97.77
99.44
99.11
94.38
96 .70
97.80
99.13
96 .39
97.72
98.14
95.95
96 .74
99.45
98.04
99.43
97.38
98.81
99.37
97.07
98.6 6
98.55
100.08
97.39
96 .44
98.91
P
LOI
0.6 0
2.6 7
2.75
8.24
0.49
0.49
0.77
1.38
2.12
2.49
2.08
1.75
1.03
3.39
4.23
2.56
4.07
2.6 5
3.54
2.50
4.11
2.79
7.6 0
2.95
5.33
5.02
4.73
4.50
3.17
1.24
1.71
0.76
7.22
7.83
0.6 2
Q| R
Zr Nb
146 13
16 2 19
128 14
79 7
80 9
137 14
141 12
113 12
114 13
140 13
119 13
116 12
114 11
16 2 17
195 20
130 14
129 15
132 15
119 14
137 16
176 16
138 16
83 9
105 11
84 9
116 12
120 14
92 11
140 16
158 18
157 17
122 13
154 19
145 16
141 13
ppm ppm
1 1
S T
Y
40
54
36
44
20
34
33
30
57
30
41
42
28
49
46
36
35
45
27
48
46
39
19
30
17
29
37
29
39
52
47
37
63
44
35
ppm
1
^Formula in cell 045 is Tiled
1 down [ =sum(C45:N45)i
Net
-15
9
84
76
q/100q
FIGURE 4.9 | Example of layout and calculation of mass changes by the single-precursor method for 34 rhyolites from the Thalanga footwall, Queensland. Data from
Paulick (1999) and Paulick et al. (2001). Values below limit of detection replaced by zero. Sample numbers with * are featured in the Thalanga data sheets (Section 7.7).
8 6 | CHAPTER 4
FIGURE 4.10 | Bivariate plots of potentially immobile elements: AI2O3, TiO2, Zr, Nb and Y (Thalanga data from Figure 4.9). (A) TiO2 versus Zr, (B) AI2O3 versus Zr,
(C) Nb versus Zr, (D) Y versus Zr, (E) TiO2 versus Nb, (F) AI2O3 versus Nb, (G) Y versus Nb, (H) TiO2 versus Y, (I) AI2O3 versus Y, (J) TiO2 versus AI2O3, (K) Na2O
versus Zr, and (L) sulfur versus Zr. These plots facilitate recognition of compositional outliers, which should be excluded from single-precursor calculations, and
selection of the least-mobile element to serve as the immobile monitor in mass-change calculations. In this example, Zr and AI2O3 both show consistent highly
correlated trends that project to the zero origin. Note the considerable scatters, and hence poorer linear correlations, in plots involving Y and the major elements sulfur
and Na2O. These are consistent with primary compositional variations or significant chemical mobility, particularly for sulfur and Na2O.
FIGURE 4.11 | Bar graph showing estimated absolute mass changes of major elements in four samples representing major alteration fades in the Thalanga footwall,
Queensland (data from Figure 4.9).
GEOCHEMISTRY OF ALTERERD ROCKS I 87
Rare-earth-element geochemistry related to
alteration
In the past few decades, trace elements have become basic tools
or pathfinders in ore deposit exploration and in petrogenetic
interpretation. Immobile trace elements have valuable appli-
cations in studies of altered rocks. The rare earth elements
(REE) have some special properties during alteration, which
may be useful in interpretation and should be understood in
order to avoid false conclusions.
Mobility of light REE and effects of net mass change
Rare earth elements (with the exception of Eu) are generally
incompatible during igneous fractionation. Heavy REE (Lu
and Yb) are essentially immobile, whereas the light REE may
be variably mobile during alteration (MacLean and Barrett,
1993). Lanthanum is the most likely to be affected and the
mobility of the other REE decreases towards the heavy REE
(Barrett and MacLean, 1994a). These incremental changes in
the lighter REE modify the slopes of chondrite-normalised
REE patterns, which may confuse petrogenetic interpretation.
Therefore, immobility of REE needs to be established before
they are used to infer magmatic affinity. Plotting geochemical
data for each rare earth element against a reliably immobile
element, such as Zr, is a means of testing for scatter and
mobility. If the elements were immobile, REE and Zr data
FIGURE 4.12 | Modifications in REE patterns due to hydrothermal alteration
illustrated by REE profiles of variably altered rhyolites from the Ansil and
Delbridge deposits, Canada (after Barrett and MacLean, 1994b). (A) Partial
leaching of the mobile light REE in the Ansil footwall is reflected in profiles with
different slopes converging towards the least-mobile heavy REE. (B) Immobility
of all REE is evident in the sub-parallel profiles for rhyolites from the Delbridge
footwall. However, net mass changes in mobile major elements have caused
changes in the proportions of the REE, producing vertical shifts in the profiles.
from altered single-precursor systems or co-genetic volcanic
suites should produce highly correlated linear trends on
bivariate plots.
In systems where REE were chemically mobile, both
positive and negative shifts in light REE concentrations have
been recorded. Significant mobility of REE appear to occur
in proximal altered zones associated with VHMS deposits
(MacLean and Barrett, 1993). The greater mobility of the
light REE may produce a fan shaped array in chondrite-
normalised plots with the REE profiles converging towards
the heavy, least-mobile REE (e.g. Fig. 4.12A)
Even in cases of REE immobility, net mass changes
associated with alteration may produce significant vertical
shifts in chondrite-normalised REE patterns. The slopes of
the REE patterns are retained, but the y-axis magnitudes are
modified (downward by net mass gain and upward by net
mass loss; Fig. 4.12B).
Europium anomalies in seafloor sediments
Recent studies of sediments and hydrothermal precipitates in
modern and ancient massive sulfide environments have found
them relatively enriched in light REE, particularly in Eu
(Barrett et al , 1990; Peter and Goodfellow, 1996; Shikazono,
1999). The explanation is that Eu exists in a divalent state in
felsic magmas and hence is compatible in feldspars, unlike the
other trivalent REE, which remain incompatible (Rollinson,
1993). The divalent Eu
+2
in feldspars may be liberated by
subseafloor hydrothermal alteration to sericite or chlorite,
whereas the incompatible REE that are concentrated in
alteration resistant phases, remain relatively immobile. The
liberated Eu
+2
is transported by reduced acidic hydrothermal
fluids and may ultimately be precipitated by oxidation at
the seafloor. Therefore, in felsic volcanic successions, altered
VHMS-footwall zones tend to be depleted in Eu. In contrast,
positive Eu anomalies exist in seafloor sediments and jaspers,
and probably indicate proximity to hydrothermal vents
(Barrett et al., 1990). Both phenomena have significance for
massive sulfide exploration in modern and ancient submarine
volcanic environments. The recognition of positive Eu
anomalies in stratiform jasper lenses recently contributed
to the discovery of a small satellite massive sulfide deposit
at Thalanga in the Mount Windsor Volcanics (Miller et al.,
2001).
4.2 I MINERAL CHEMISTRY
Principles
Minerals, by definition, are natural homogenous solids of
definite chemical composition and definite atomic structure
(Dana, 1957). However, many minerals do not have simply
defined chemical formulas. Their compositions may lie
between limits defined by two or more end-member formulas,
effectively forming solid solutions. Mineral crystal structures
can accommodate various impurities where atoms and ions
8 8 | CHAPTER 4
of suitable size and charge can substitute for others in the
lattice, occupy interstices in the lattice or be omitted from a
proportion of lattice sites.
The considerable variety of linked tetrahedral crystal
structures in silicate minerals permits a wide range of chemical
substitutions and interstitial solid solutions. A frequently
cited example is the olivine series in which Mg
2+
and Fe
2+
ions, having similar charge and size, substitute for each
other between the end member compositions of forsterite
(Mg
2
SiO
4
) and fayalite (Fe
2
SiO
4
).
The sheet-like structures of phyllosilicates allow a great
range of ionic substitutions and interstitial contaminants.
For example, muscovite, with the ideal formula of
K
2
Al
4
[Si
6
Al
2
O
20
] (OH)
4
, commonly contains the isomorphous
substitutions of Na, Rb, Cs, Ca and Ba for K; Mg, Fe
2+
, Fe
3+
,
Mn, Li, Cr, Ti, and V for octahedral Al, and F for OH and
tetrahedral cation proportions from Si
6
Al
2
to Si
7
Al] (Deer
et al., 1966). Layered clays (particularly smectites) also
accommodate many cationic substitutions and exchanges as
well as physical mixtures and inter-layered structures of more
than one clay mineral (e.g. smectite-illite).
The causes of such variations are both physical and
chemical. Temperature, pressure, fluid pH, JO
2
, and cation
solubility may all affect the stability and composition of
the minerals formed. Chemical and temperature gradients
in hydrothermal alteration systems commonly produce
spatial zonation of alteration mineral assemblages. Similarly,
they may produce gradational and zonal variations in some
alteration mineral compositions. For example, the Mn
content of metasomatic pyroxenes associated with Zn skarns
generally increases systematically along the fluid pathway, and
can be used to identify proximal and distal skarns, and altered
zones (Meinert, 1993).
The standard methods for determining mineral com-
positions and crystal structure are electron microprobe and
X-ray diffraction analyses. Although Galley (1995) noted that
these methods are more readily available than previously, the
impracticality of mineral chemistry as an exploration tool, and
the complexity and cost of these laboratory techniques means
that they are rarely used other than in academic research.
The development of field portable short wavelength infra red
(SWIR) spectrometers and spectral interpretation software
during the last decade, has allowed mineral chemistry to be
practically integrated into exploration programs for a variety
of deposit types (Thompson et al., 1999). SWIR spectral
analysis can determine compositional and crystal-structural
variations in white micas, smectites, clays, chlorites, biotites
and carbonates (Pontual et al., 1997). These minerals are
prominent in altered volcanic rocks and are also prone to
significant compositional variations. The technique reliably
estimates variations in white mica composition but appears
to be less effective at analysing chlorites in typical mixed
phyllosilicate assemblages (Herrmann et al., 2001). Carbonates
have relatively weak SWIR absorptions, which tend to be
obscured in mixed assemblages containing phyllosilicates and
are thus less amenable to spectral analysis.
Although there are many possible applications in mineral
deposit exploration, we do not know of any cases where
systematic investigations of mineral chemistry have led to
an ore discovery. As pointed out by Simmons and Browne
(2000), the extent to which patterns of mineral distribution
and chemical variations can be used in exploration depends
largely on whether they are related to a single phase of
hydrothermal activity that produced equilibrium mineral
assemblages.
Applications
The three main applications of mineral chemistry in alteration
studies are:
(1) interpretation of the processes and physicochemical
conditions of alteration
(2) discrimination or identification of metasomatic alteration
and mineralisation styles from mineral compositions
(3) determination of spatial variations and exploration vectors
to ore.
Interpretation of processes
The compositional and crystal-structure variations in some
minerals are diagnostic of particular alteration processes
because of physicochemical influences on mineral stability and
composition. Thus, mineral chemistry may be used to infer
the geological environment in which an alteration mineral
assemblage formed. However, these kinds of compositional
variations may not always be universally applicable; they may
require orientation testing to determine their usefulness in
different districts or sites.
For example, Dill et al. (1997) found that the kaolin-
alunite deposits in felsic volcanic rocks of western Peru could
be classified into hypogene (hydrothermal) and supergene
(weathering) types on the basis of the chemical variations
in kaolinite. Hydrothermal kaolinite tended to be rich in
Ba, Sr and sulfur, whereas weathering-related kaolin clays
concentrated Cr, Ti, Nb and REE. This approach has direct
applications to mineral exploration because hypogene kaolinite
alteration in Peru is associated with high-sulfidation epithermal
Au-Ag deposits. Similarly, Yang et al. (1999) alluded to low-
and high-crystallinity forms of kaolinite in the Comstock
district of Nevada, which they respectively attributed to low-
temperature weathering and higher temperature hydrothermal
alteration processes. They suggested that spectral recognition
of distinctive kaolinite could be used in satellite-borne remote
sensing to detect prospective hydrothermal altered zones.
Discrimination of hydrothermal alteration styles
Mineral compositions can be used to identify or fingerprint
hydrothermal-alteration mineral assemblages, and possible
associations with mineralised rock. This application is useful
at early stages of mineral exploration to discriminate between
economically favourable and less favourable alteration and
mineralisation styles. For example, Zn-skarn assemblages are
commonly dominated by pyroxene with varying amounts
of garnet, amphibole, bustamite, chlorite and carbonate,
which may all be Mn enriched. Manganese-rich pyroxene
(johannsenite) is virtually diagnostic of distal Zn skarns
(Meinert, 1983) and could be used as an index mineral in
exploration for this type of deposit.
Another example that is more relevant to submarine
volcanic successions, is the recognition of a class of pyritic-
alteration systems in the Mount Read Volcanics that have
some characteristics of high-sulfidation epithermal deposits.
These include Western Tharsis and Lyell-Comstock, which
contain sub-economic Cu + Au resources (Huston and
Kamprad, 2000; Corbett, 2001), and Basin Lake and Chester,
which appear to be barren (Boda, 1991; Green and Taheri,
1992; Williams, 2000; Williams and Davidson, 2004). White
micas in the central zones of these systems have distinctive,
non-phengitic sodic compositions (Herrmann et al., 2001).
The recognition of sodic white mica, along with low 6
34
S
values in pyrite and the presence of pyrophyllite, enables the
discrimination of this type of alteration system from those
associated with economic Zn-rich polymetallic VHMS
deposits (i.e. Rosebery and Hellyer). The altered footwall
zones of Zn-rich polymetallic VHMS deposits contain
normal potassic to slightly phengitic white micas. The fact
that the compositional variations in white mica can be simply
determined by SWIR spectral analysis makes this a practical
method for selecting and ranking exploration targets.
Mineral chemistry exploration vectors
Many documented studies have shown spatial variations in
mineral chemistry in altered zones around VHMS deposits.
Chlorite has received the most attention, particularly in the
last two decades, and there are few deposits for which no
data are available. There has also been significant interest in
carbonate and, to a lesser extent, white mica compositions.
Chlorite compositions are typically Mg-rich in the
proximal altered zones of VHMS deposits. They commonly
show systematic distal trends to more Fe-rich compositions.
These trends are typically recognisable over several hundred
metres, both laterally and stratigraphically into the footwall,
away from the ore. Some examples include the Seneca and
Corbet deposits in Canada (Urabe et al., 1983), the Arctic
deposit in Alaska (Schmidt, 1988), and the Thalanga deposit
in north Queensland (Paulick et al., 2001). However, there
are many cases where the opposite trend exists and Fe-rich
chlorites occur in proximal altered zones. The Aznacollar and
Masa Valverde deposits are two examples in the Iberian pyrite
belt (Sanchez-Espana et al., 2000). At the Home deposit,
Canada (MacLean and Hoy, 1991), chlorites in proximal
chlorite-rich zones are more Fe-rich than in the enclosing
sericite + chlorite zone (Fig. 4.13). Similarly, at Mattagami
Lake (Abitibi belt, Canada) there is a general trend of Fe
enrichment in chlorites upwards towards the ore position and
outwards from the core of the altered footwall zone (Costa
et al., 1983). In northern Turkey, the dacite-hosted deposits
of the eastern Black Sea province have altered footwall zones
of Mg chlorite and sericite (Cagatay, 1993). In contrast, the
western Black Sea ophiolite-hosted pyritic Cu deposits of
the Kure district are associated with Fe-rich chlorites and
trends of Fe enrichment toward ore. Some deposits exhibit
inconsistent patterns of variations in chlorite composition.
McLeod (1987) found that Mg chlorites around the Mount
Chalmers deposit (Queensland) have a stratigraphic upwards
trend of Fe enrichment in the footwall and a sharp reversal to
Mg enrichment in the mineralised zone. Two recent regional-
GEOCHEMISTRY OF ALTERERD ROCKS | 89
scale studies by Hannington et al. (2003a, 2003b) showed
contrasting compositional trends in chlorites associated
with VHMS deposits in the Noranda district, Canada, and
Kristineberg deposits of the Skellefte district, Sweden. In
the Noranda district, the moderately Fe-rich compositions
of chlorites (Fe/Fe+Mg 0.4-0.9) associated with sulfide
deposits and surrounding district-scale hydrothermally
altered zones, contribute to discrimination of prospective and
non-prospective volcanic centres. However, chlorites in the
Kristineberg district are distinctly Mg rich (Fe/Fe+Mg 0.05-
0.5) and show little variation between proximal and distal
alteration facies. These studies also found the chlorites
associated with sulfide deposits contained significant Mn (up
to 1% MnO) and Zn (up to 0.5% ZnO) suggesting that these
could be used as proximity indicators in exploration.
McLeod and Stan ton (1984) investigated several eastern
Australian VHMS deposits and showed that chlorites in
sphalerite-rich ores are relatively Mg rich compared to
those in chalcopyrite-rich ores. Furthermore, the Mg/Fe
ratios of chlorites are related to Mg/Fe ratios of co-existing
phyllosilicates and the Fe content of co-existing sphalerite.
Therefore, zonal compositional variations in chlorite may
be reflected in other alteration mineral species, such as
white mica, which may be more easily measured by SWIR.
Importantly, McLeod and Stan ton (1984) concluded that the
compositions of chlorites and other phyllosilicates had not
been significantly modified by subsequent greenschist facies
metamorphism.
Variations in chlorite composition have also been used,
with some success, as empirical and thermodynamically
calculated geothermometers to estimate temperature gradients
in hydrothermal systems above 200C (e.g. Cathelineau
and Nieva, 1985; Walshe, 1986). They are sensitive to
re-equilibration and therefore not reliable indicators of
hydrothermal temperatures in subsequently metamorphosed
terrains (Green and Taheri, 1992).
FIGURE 4.13 | Al
iv
-Mg-Fe cation plot showing trend to Fe-rich chlorite with
proximity to the Cu-Au VHMS deposit at the Home Mine, Quebec, Canada (after
MacLean and Hoy, 1991). Where Ab = albite, Ep = epidote, Mt = magnetite and
Ser = sericite.
90 | CHAPTER 4
White micas, commonly referred to as sericite, are
nearly ubiquitous in massive sulfide-related hydrothermal
alteration systems and they can vary considerably from the
ideal muscovite formula of K
2
Al
4
[Si
6
Al
2
O
20
](OH)
4
(Deer et
al., 1966). The term phengite refers to white micas in which
Fe, Mg and some other cations substitute for Al in octahedral
sites and the charge balances are maintained by increased Si/Al
ratios in tetrahedral sites. Phengitic micas form solid solutions
between the end members of muscovite and celadonite:
K
2
(Mg,Fe2
+
)
2
(Al,Fe
3+
)
2
Si
8
O
20
](OH)
4
. Barium-rich phengitic
micas, in which Ba substitutes for K in inter-layer sites, also
exist in some sediment-hosted sulfide and VHMS deposits
(e.g. Schmidt, 1988; Jiang et al., 1996; Leistel et al., 1998).
At low to moderate temperatures, there may be limited Na
substitution for K, with Na/Na+K ratios up to about 0.2.
The sodic muscovites generally have low phengite contents.
White micas formed at temperature below 300C may have a
significant proportion of vacancies in inter-layer sites normally
occupied by K, as well as phengite-like Fe-Mg substitution
in octahedral sites. These are commonly termed illites; they
form complex solid solutions between three end-members:
muscovite, celadonite and pyrophyllite. Yang's (1998) review
provides a more detailed description of variations in white
mica compositions.
Although white mica composition has been examined
in a number of massive sulfide-related hydrothermal
alteration systems, few studies were systematic enough to
evaluate its usefulness as an exploration tool. White micas
in the proximal altered zones of the weakly metamorphosed
Hellyer deposit, western Tasmania, are more phengitic than
the normal muscovites in distal altered zones (Yang, 1998).
At the nearby but slightly more metamorphosed Que River
deposit, Offler and Whitford (1992) found considerable
small-scale compositional variations in mica, even within
single samples, due to a complex alteration history.
Although the metamorphic phases preserve hydrothermal
alteration compositional trends, no convincing vectors were
recognised, possibly because of structural complications.
Around the Arctic deposit, Alaska, white micas span almost
the entire compositional range between muscovite and
celadonite (Schmidt, 1988). Metamorphic micas outside the
hydrothermal altered zones are highly phengitic. Micas in a
variety of proximal alteration mineral assemblages are variably
phengitic; some contain up to 0.4 cations of Ba per formula
unit and the least-phengitic types are significantly sodic. It is
not clear whether these variations are systematic enough to be
used as broad exploration vectors.
Altered zones in the Iberian pyrite belt also have a confusing
variety of mica composition patterns. Plimer and de Carvalho
(1982) found that white micas in altered footwall zones
around the Salgadinho Cu deposit are phengitic, and appear
to show increase in Fe/Fe+Mg ratios towards the mineralised
zone. In contrast, in the Rio Tinto deposit the proximal altered
zones contain muscovite and the distal altered zones (up to
2500 m from the deposit) contain micas of more phengitic
composition (Leistel et al., 1998). The Masa Valverde deposit
is associated with Ba-rich muscovites and some ore bodies in
the Aljustrel district have extensive halos of sodic white mica
(Leistel et al., 1998; Carvalho and Barriga, 2000).
With the possible exception of Ba substitution, most of the
compositional variations in white micas are semi-quantifiable
by SWIR spectral analysis. This has been demonstrated
by several recent studies in the Mount Read Volcanics
(Herrmann et al., 2001). Figure 4.14 provides an example of
variations in wavelengths of Al-OH bond-related absorption
features in SWIR spectra of white mica in samples taken at
intervals from a single drill hole. These wavelength variations
are directly related to white mica compositional variations.
Some alteration systems, particularly those associated with
disseminated Cu-Au deposits and/or kaolinite pyrophyllite
assemblages, exhibit compositional gradients in white mica
compositions that are measurable over a few hundred metres.
The background white mica compositions are commonly
variably phengitic and tend to non-phengitic muscovite
or sodic-muscovite in proximal altered zones (e.g. Huston
and Kamprad, 2000; Herrmann et al., 2001). Figure 4.15
illustrates variations in wavelengths of Al-OH absorption
features, related to white mica composition, spatially around
the Western Tharsis deposit.
Carbonates are a third group of minerals that can
accommodate compositional variations and are common in
some VHMS altered zones. Documentation of carbonate
compositional trends and zonal distributions is fairly sparse.
However, it seems that massive-sulfide-related carbonates are
typically Fe-, Mg- or Mn-bearing phases, and background
diagenetic or metamorphic carbonates are commonly calcic.
Documented examples include the Hokuroku district in
Japan (Shikazono et al., 1998), the Rosebery deposit in
western Tasmania (Large et al., 2001b), and the South Bay
deposit in northwest Ontario (Urabe et al., 1983).
FIGURE 4.14 | Stack of selected SWIR hull quotient spectra of core samples
from a diamond-drill hole through the altered zone at the Chester deposit,
western Tasmania. Annotations on the left side are depths in metres down the
hole. The spectral features are almost entirely attributable to white mica in the
alteration mineral assemblages. Note the distinct variation in wavelengths of
the Al-OH absorption features at around 2200 nm. These indicate that the hole
intersected mineral assemblages containing normal potassic muscovite in the
upper part, sodic white mica from about 100 to 250 m and muscovite to slightly
phengitic white mica in the lower part.
GEOCHEMISTRY OF ALTERERD ROCKS | 91
FIGURE 4.15 | Cross-section of the Western Tharsis deposit (western
Tasmania) showing zonation of wavelengths of AI-OH absorption features in
SWIR spectra. The background of 2200-2210 nm, corresponding to slightly
phengitic white mica, decreases over a few hundred metres to 2194-2198 nm,
attributable to non-phengitic, slightly sodic white mica in the proximal altered
zone associated with disseminated pyrite and chalcopyrite.
Inevitably, there are exceptions, such as the deposits of
the northern Iberian pyrite belt, which have calcite, ankerite
and dolomite in proximal alteration mineral assemblages
(Sanchez-Espana et al., 2000). The Mattabi deposit in the
Sturgeon Lake area, Canada, is underlain by a funnel shaped
siderite-rich altered zone grading outwards to dolomite,
which is widespread on a district scale in the footwall and
hanging-wall volcanic rocks.
Hydrothermal carbonates in the Rosebery-Hercules area,
western Tasmania, are conspicuously Mn rich, (Khin Zaw and
Large, 1992; Large et al., 2001b). Large et al. (2001b) showed
that Mn-siderite and ankerite carbonates in the footwall of
the Rosebery deposit increase in Mn content towards ore (Fig.
4.16). Magnesium contents of carbonates in altered footwall
zones of the South Bay deposit, Canada, increase steadily
towards ore over distances of tens to hundreds of metres
(Urabeetal., 1983).
Chlorite, white mica and carbonate all have potential as
mineral exploration vectors, at least on a prospect or deposit
scale. However, the considerable diversity of compositional
trends in the published data indicate that exploration vectors
need to be empirically established on a district or deposit
specific basis, and are not universally applicable.
FIGURE 4,16 | Downhole plot of drill hole 120R illustrating the distribution of
Mn-rich carbonates (kutnahorite and manganosiderite-rhodochrosite) in proximity
to K-lens of the Rosebery Pb-Zn VHMS deposit, western Tasmania. Magnesium-
carbonates occur in the footwall and in a thin unit of altered pumice breccia
immediately above the ore lens; carbonates more than 50 m above ore in the
hanging wall sequence are Ca rich.
92 I CHAPTER 4
4.3 I STABLE ISOTOPES
TABLE 4.2 | Natural abundances of H, C, 0 and S isotopes, and standards in
common use (data from Rollinson, 1993).
Theoretical background
Isotope geochemistry is a diverse and rather specialised
science. This section aims to provide a bare outline of
aspects that have particular relevance to interpretation of
altered volcanic rocks. It includes only a brief introduction
to theoretical principles, necessary to grasp the applications.
We recommend that interested readers supplement this by
referring to other textbooks Rollinson (1993) provides an
excellent working basis.
Isotopes are distinct atomic forms of elements that
have the same number of protons but different numbers
of neutrons in their nuclei. Of the 92 naturally occurring
elements, 60 consist of more than one isotope and many
of them have two or more stable isotopes. That means that
they are non-radioactive, and do not change naturally, or
decay, into other radiogenic elements by emissions of sub-
atomic particles from their nuclei. Some natural radiogenic
isotopes have important geological uses in geochronology,
petrogenesis and metallogenesis, because their rates of decay
are constant and measurable. Stable isotopes also have many
geological applications, mainly based on their properties
of isotopic fractionation. The stable isotopes of the light
elements H, C, O and sulfur have received the most attention
from geochemists because they are naturally abundant in the
hydrosphere and in crustal rocks, not least in altered rocks.
Informal isotopic notation uses the chemical symbol of the
element preceded by the mass number of the isotope written
as a superscript. Thus
17
O denotes the oxygen isotope with
17 nucleons, comprising eight protons and nine neutrons. A
single isotope, which usually has equal numbers of protons
and neutrons, typically dominates the isotopic composition
of each element (Table 4.2). Therefore isotopic ratios are very
small numbers (e.g. for the average abundances of oxygen
isotopes,
18
O /
16
O = 0.002). To avoid direct comparison of
these unconvincingly small ratios, stable isotopic proportions
are expressed in parts per thousand (i.e. per mil, %o) relative
to a standard material (i.e. delta form). For example:
Stable isotopes undergo fractionation (or selective
partitioning into different phases) according to thermodynamic
properties that are related to their differing atomic weights
and consequent ionic bond strengths (Faure, 1986; Rollinson,
1993). Fractionation may occur by several physicochemical
processes of which the most geologically important are
isotopic exchange reactions between phases. The degree of
fractionation is controlled by physical and chemical factors,
which vary according to the elements and fractionation
processes involved. Thus, O-isotopic fractionation is largely
dependent on temperature, whereas S-isotopic fractionation
is influenced by temperature, pH, / O
2
, and the activities of
sulfur and other cations involved with sulfate.
Isotopic applications in alteration studies
Isotopic studies of alteration mineral assemblages associated
with mineralised zones may help to estimate alteration
temperatures and water-rock ratios, interpret fluid origins,
discriminate between alteration styles and identify altered
halos around ore deposits. However, it is worth repeating
Ohmoto's (1986) cautionary advice to integrate isotopic
studies with geologic, mineralogic and geochemical data. He
stated: 'there is more than one process, which may produce
the same isotopic characteristics (in an ore deposit) and
the same geological process may produce entirely different
isotopic characteristics in different conditions. Therefore,
isotopic data alone cannot provide a unique answer to any
geological problem, especially when the data are limited to
isotopes of one element.'
Diagenetic and hydro thermal alteration of volcanic rocks
invariably involves hydration reactions, between minerals
and water, and so the amount and isotopic composition
of water are important variables. Apart from geologic
variability, sample preparation, isotopic analytical methods
and calibration of fractionation factors also introduce
significant uncertainties. Experimentally, empirically and
thermodynamically determined isotopic fractionation factors
provide a confusing diversity of choice for use in isotopic
calculations. The Laboratoire de geochimie isotopique at
Universite Laval, Quebec, has a comprehensive compilation
of fractionation factors from many published sources and
is accessible at <www.ggl.ulaval.ca/personnel/beaudoin/labo>
(Beaudoin and Therrien, 1999).
Geothermometers
The temperature dependency of isotope fractionations
between mineral pairs forms the basis of isotope geo-
1
H 99.98 4 4 Std mean ocean water (SMOW), V ienna-
SMOW (V -SMOW) or PDB belemnite.
2
D 0.01 5 6
1 2
C 98 .8 9 PDB belemnite
1 3
C 1 .1 1
1 6
O 99.7 6 30 Std mean ocean water (SMOW), V ienna-
SMOW (V -SMOW) or PDB belemnite.
1 7
O 0.037 5
1 8
O 0.1 995
32
S 95 .02 Troilite in Canon Diablo meteorite (CDT)
33
S 0. 7 5
34
S 4 . 2 1
36
S 0.02
GEOCHEMISTRY OF ALTERERD ROCKS | 93
thermometry. Provided that the paired minerals formed in
equilibrium, that their fractionation factors are known and
are significantly different, their original isotopic compositions
have been retained and can be separately determined, then a
combination of equations can be solved for temperature of
formation.
This approach is useful with O isotopes, because O is
common to, and abundant in, silicates and other alteration
phases, such as carbonates and sulfates. It is also applicable
to S-isotopic compositions of minerals in complex sulfide
and ore assemblages. Hydrogen isotopes are not generally
reliable as geothermometers because they are readily modified
by subsequent fluid interactions. Furthermore, the mineral
fractionation factors are relatively insensitive to temperature
and are not well calibrated (Ohmoto, 1986).
It is usually difficult to physically separate fine-grained
minerals for isotopic analysis and to petrographically demon-
strate equilibrium between the analysed phases. However,
close agreement between several temperature estimates of two
or more pairs of minerals in a single assemblage (e.g. quartz +
magnetite, muscovite + chlorite and calcite + chlorite) would
inspire reasonable confidence in their isotopic equilibrium
and the calculated temperature.
In fluid-dominated hydrothermal systems, mineral-water
O-isotope fractionation factors can be used to estimate relative
temperatures. Although it is difficult to reliably measure
the isotopic composition of the water from fluid inclusions
(Nesbitt, 1996), an assumed value can provide approximate
or relative temperature estimates. This approach is used in
the determination of ocean palaeo-temperatures from 6
18
O
values of the carbonate shells of marine organisms (Rollinson,
1993) and also has applications in mineral exploration (e.g.
Miller et al., 2001).
Fluid origins
Natural waters have a broad range of H- and O-
isotopic compositions because of fractionation effects
in the hydrosphere, lithosphere and mantle (Fig. 4.17).
Consequently, it may be possible to infer the source or sources
of alteration fluids, and something about their evolution,
from their isotopic signature.
Fluid inclusions in hydrothermal minerals may permit
direct measurement, but commonly the fluid compositions are
calculated from isotopic compositions of alteration minerals
with known fractionation characteristics, that are assumed to
have been in equilibrium with the hydrothermal fluid. Isotopic
composition of a single hydrothermal mineral may constrain
the fluid composition if independent temperature estimates,
such as fluid inclusion data, are available. Otherwise, isotopic
compositions of mineral pairs in equilibrium can be used (as
outlined above) to deduce temperature, which can then be
applied in the mineral-water fractionation relationship to
estimate fluid-isotopic composition.
Water-rock ratios
Knowledge of water-rock ratios may help to determine
the processes of alteration and interpret the hydrology of
FIGURE 4.17 | 6 D- and S
18
O-isotopic compositions of natural waters (from
Taylor, 1979; Ohmoto, 1986 ). SMOW is standard mean ocean water with 6 D-
and6
18
O values of 0%o.
hydro thermal alteration systems. Water-rock ratios can be
estimated from whole-rock O-isotope data or from inferences
of mass transfers and solubilities (e.g. Ohmoto et al., 1983).
Unaltered mafic volcanic rocks have initial 6
18
O values
in the range 6 to 7.5%o, slightly higher than the mantle
value of 5.7%o, and unaltered felsic volcanic rocks have
values up to about 10%o, (Hoefs, 1973). The 6
18
O values of
hydrothermally altered volcanic rocks will differ from initial
values, depending on the temperature and mineral assemblage
(which affect fractionation), the initial isotopic composition
of the water and the quantity of water that reacted with a given
amount of rock. The water-rock ratio is usually expressed in
atomic proportions of oxygen.
Taylor (1979) presented the following equations expressing
these relationships in closed and open hydrothermal systems:
closed systems
where the superscripts and subscripts i, f, w and r, respectively
refer to initial, final, water and rock.
These equations can be plotted as curves of the type
illustrated in Figure 4.18, which relate 8
18
O / to water-rock
ratios. Thus, a measured final 5
18
O
r
can be used to estimate
the amount of water involved in hydrothermal or diagenetic
alteration, under assumed (or otherwise determined) values
for temperature, whole-rock fractionation factors and the
initial isotopic compositions of fluid and rock.
However, as discussed in some detail by Ohmoto (1986)
and noted by Green and Taheri (1992), natural geologic
systems are not likely to be simple isothermic, closed or open
systems. Rates of isotopic re-equilibration vary according to
temperature, and the isotopic compositions of both rock and
94 | CHAPTER 4
30
FIGURE 4.18 | Curves illustrating relationships between water-rock ratio
and final-rock 6
18
O at various equilibration temperatures under parameters of:
6
18
Owatef = 0 %o, S
18
Cy = 7 %0, fractionation factor Aw' = (2.6 8 x K W ) - 3.53
(plagioclase). Solid and dashed lines represent open and closed systems,
respectively. Note that re-equilibration with a small amount of water at low
temperature can produce a large increase in rock 6
18
O.
fluid change incrementally along the flow path. The final rock
5
I8
O reflects an integrated history of fluid-rock reaction and
offers only broad constraints on temperature and water-rock
ratio. Green and Taheri (1992) suggested that conditions
of diagenesis might approximate a closed system, whereas
submarine hydrothermal convection is more analogous to an
open system. Water-rock ratios calculated under assumptions
of either closed or open systems are likely to represent the
minimum values because of kinetic and incremental factors
affecting rates of re-equilibration. Natural open systems may
require water-rock ratios one or two orders of magnitude
greater to achieve equivalent shifts in the isotopic composition
of the rock (Ohmoto, 1986).
Nevertheless, consideration of water-rock ratios is
important in evaluation of whole-rock 6
18
O data. This will
be further explained in the following section on isotopic
exploration vectors.
Oxygen-isotope exploration vectors
Early isotopic studies (e.g. O'Neil and Silberman, 1974;
Taylor, 1974) discovered the link between terrestrial epi-
thermal Au-Ag deposits, meteoric-hydrothermal convection
and very broad halos of low 6
18
O in volcanic host rocks. These
extensive isotopic halos had obvious potential as semi-regional
exploration vectors and stimulated further investigations into
volcanic successions hosting other deposit types.
Among them was the landmark study by Green et al.
(1983) on whole-rock O-isotope geochemistry in the host
rocks to VHMS deposits in the Hokuroku district, Japan. They
found concentric zonation of whole-rock 6
18
O values around
the cluster of Fukuzawa ore bodies ranging from 6.7 + 1.3%o
FIGURE 4.19 | Cross-section illustrating the distribution of whole-rock 5
18
O values (black contours), and altered footwall zones around the Fukuzawa deposits,
Hokuroku district, Japan (modified after Green et al., 1983).
in the proximal sericite + chlorite zone and 11.1 + 2.5%o
in the surrounding 13 km-wide montmorillonite zone,
to 16.9 2.7%o in the outer zeolite zone (Fig. 4.19). The
6
18
O anomaly is significantly broader and less variable than
elemental geochemical halos; it extends up to 1 km laterally
beyond the Na
2
O depletion anomaly and at least 400 m into
the hanging wall above the mineralised zone. The wide extent
of the whole-rock 5
18
O anomaly is advantageous for regional
exploration. It has particular application in deformed terrains
where the original mineral assemblages of hydrothermally
altered zones has been obscured by subsequent metamorphism,
because the hydrothermal whole-rock 6
18
O patterns may still
be preserved. This is because regional metamorphism typically
involves low water-rock ratios.
The observed whole-rock 6
18
O values are consistent with
isotopic exchange between the host rocks and large amounts
of seawater (0%o, w/r >1) at different temperatures. Isotopic
modelling, using a plagioclase fractionation factor as an
average felsic volcanic rock value, showed that high 6
18
O
values in the diagenetic zeolite zone could be produced by
interaction with fluid of virtually any source (magmatic, sea
or meteoric) at low temperatures and relatively small water-
rock ratios. This may be due to the high fractionation factors
between silicates and water at temperatures below 100C.
However, low 8
18
O values of the proximal sericite + chlorite
zone are more consistent with conditions of equilibrium with
GEOCHEMISTRY OF ALTERERD ROCKS | 95
either meteoric waters at unrealistically low water-rock ratios
(-8%o, <0.2) or seawater at 200-300C and moderate to large
water-rock ratios. The submarine volcanic environment and
implications of hydrothermal mass transfers favour the latter
interpretation.
Cathles (1983) carried out detailed thermal, geochemical
and isotopic analysis of a hypothetical, but geologically
realistic, submarine intrusion-heated convective hydrothermal
system. His model produced O-isotopic results that were
consistent with 5
18
O data observed by Green et al. (1983)
in the Hokuroku district. It predicts that rocks in the shallow
substrate become isotopically heavier by reaction with
down-welling seawater at low temperatures. Lower isotopic
fractionation, due to increased temperatures at depths greater
than about 2 km below the seafloor, produces a zone of low
rock 5
18
O. As the convective system evolves, and depending
on permeability and rate of isotopic exchange, the deep-
low 8
18
O zone migrates up through the shallow-high 8
18
O
anomaly, to produce a low 5
18
O isotopic anomaly around the
vent site (Fig. 4.20).
Chapter 8 summarises some other deposit- and district-
scale isotopic studies, which illustrate the possible complexities
in submarine volcanic successions, but indicate significant
potential for whole-rock O-isotope geochemistry in targeting
mineral exploration: potential that has not been widely
applied outside academic studies.
FIGURE 4.20 | Modelled distribution of changes in whole-rock 6
18
O values due to hydrothermal alteration generated by the convection of fluid
around a subseafloor intrusion (after Cathles, 1983). The intrusion is 1 km wide and 3.25 km deep, and emplaced with its top 1.75 km below the
seafloor. Re-equilibration with down-welling, low-temperature seawater produces a shallow zone of higher 8
18
O. Increasing temperatures at depth
create a sub-horizontal zone of low 8
18
O, which propagates up to the seafloor resulting in the characteristic low 8
18
O surrounded by a halo of
positive anomalies. Note that the contours represent shifts from the initial rock 8
18
O values, not the actual rock 8
18
O values.
96
I 97
5 | SEAFLOOR- AND BURIAL-RELATED
ALTERATION
This chapter discusses the alteration processes and their
products (textures, minerals and zones) that occur immediately
after deposition and during burial of volcanic facies in
submarine environments. It encompasses the relatively low-
temperature processes of hydration, diagenesis and earJy
burial metamorphism.
Provided sufficient time, burial-related alteration ultimately
results in the lithification of clastic facies in the succession.
Oxidation, hydration, dissolution, dehydration, ion exchange,
and hydrolysis reactions result in the breakdown of volcanic
glass, precipitation of authigenic minerals in pore space, and
replacement of glass and magmatic minerals by new minerals.
Alteration mineral assemblages may change over time due
to changing physical and chemical conditions during burial,
and may progress to low-pressure, high-temperature regional
metamorphic assemblages at depth (Coombs et al., 1959).
The recognition and description of burial-related alteration
styles in submarine volcanic successions has implications
for exploration and ore genesis studies, because of dramatic
changes in porosity and permeability, which result from
cementation, compaction and dissolution during diagenesis.
These changes influence subsequent fluid pathways and the
sites of hydrothermal venting and mineralisation. Although
it has been frequently assumed that compositional changes
associated with diagenetic alteration are limited, they may
involve mass changes of up to 9% (Gifkins and Allen, 2001).
Diagenetic and burial metamorphic mineral assemblages and
the thickness of altered zones can also be used to determine a
basin's thermal history (e.g. Utada, 1991).
5.1 | ALTERATION RELATED TO SEA-
FLOOR PROCESSES AND BURIAL
Distinctive weathering and burial-related alteration processes
occur in submarine volcanic successions because of rapid
accumulation rates, and the presence of abundant glass and
seawater. Silicate glasses are more susceptible than minerals
to alteration, because they lack well-developed crystal
structures and thus will readily devitrify, dissolve or alter to
minerals. Glass fragments are especially prone to alteration
because of their reactivity and large surface area to volume
ratio. At elevated temperatures, volcanic glass readily alters
in the presence of alkaline fluids, but the rate of alteration is
reduced under dry conditions or in the presence of pure water
(Lofgren, 1970). For example, hydration and devitrification
rates of feJsic voicaniclastic Facies increase one to Eve orders
of magnitude in the presence of seawater (Lofgren, 1970,
1971b).
Another important aspect of burial-related alteration in
volcanic and igneous rocks is that anhydrous primary igneous
minerals that have crystallised at high temperatures (e.g.
olivine and pyroxene) become unstable and alter to hydrous
minerals at lower temperatures. The extent of these retrograde
reactions depends on the availability of water and the rock
permeability.
The effects of diagenesis and burial metamorphism on
thick, proximal volcanic successions are relatively poorly
understood and documented, and detailed studies are
almost exclusively limited to well-sorted, fine-grained felsic
voicaniclastic facies.
The Ocean Drilling Program (ODP) in fore-arc and
back-arc basins in the western Pacific region has provided
important information on the behaviour of volcanic
components during early low-temperature alteration and
lithification, and the factors controlling the intensity and
depth of diagenetic alteration in Miocene to Recent felsic to
intermediate sandstones (e.g. Hein and Scholl, 1978; Taylor
and Surdam, 1981; Klein and Lee, 1984; Hay and Guldman,
1987; Marsaglia and Tazaki, 1992; Tazaki and Fyfe, 1992;
Torres et al., 1995). Studies in mafic volcanic successions have
generally been limited to seafloor alteration (e.g. Bonatti,
1965; Hay and Iijima, 1968a; Honnorez, 1978; Zhou and
Fyfe, 1989).
Limited work in uplifted and eroded ancient submarine
successions provides data on diagenetic and burial
metamorphic minerals, textures and zones that formed at
depths greater than 1 km (e.g. in New Zealand, Coombs,
1954; Coombs et al., 1959; in Canada, Kuniyoshi and Liou,
1976; Starkey and Frost, 1990; in Australia, Smith, 1969;
Smith et al., 1982; Gifkins and Allen, 2001; Gifkins et al., in
press; and in Japan, Hay and Iijima, 1968a; Seki et al., 1969;
Iijima and Utada, 1972; Utada, 1991).
Active geothermal regions provide direct measurements of
temperatures, alteration mineral assemblages and pore water
98 | CHAPTER 5
chemistry at relatively shallow depths, less than 2 km (e.g.
Coombs et al., 1959; White and Sigvaldason, 1962; Viereck
et al., 1982). In addition, experimental work on the alteration
of natural and synthetic glasses by modified seawater provides
estimates of alteration mineralogy, temperature ranges for
mineral species, fluid-rock ratios, elemental variations in glass,
and variations in fluid chemistry over time. Basalt-seawater
experiments were performed by: Hajash (1975, 1977),
Keene et al. (1976), Seyfried and Bischoff (1977), Mottl and
Seyfried (1977), Seyfried et al. (1978), Hajash and Archer
(1980), Seyfried and Mottl (1982), and Ghiara et al. (1993).
Rhyolite-seawater experiments were conducted by: Ellis and
Mahon (1964), Sakai et al. (1978), Hajash and Chandler
(1981), Shiraki et al. (1987) and Shiraki and Iiyama (1990).
Physical conditions
Early studies assumed that burial-related alteration mineral
assemblages and zonation patterns in submarine volcanic
successions were controlled by pressure and temperature
conditions. However, it is now believed that the composition
and pressure of intergranular fluids and the composition of
the primary facies are more important (e.g. Miyashiro and
Shido, 1970; Surdam, 1973). Differences in the mineral
assemblage, intensity, stratigraphic position and sequence of
burial-related altered zones may be explained by variations in:
primary rock composition, pore-fluid composition, pore-fluid
pressure, geothermal gradient and hence temperature, burial
history and sediment accumulation rate, interaction time or
age, fluid-rock ratio, porosity and permeability, and tectonic
setting (Hay, 1966; Surdam, 1973; Furnes, 1975; Boles and
Coombs, 1977; Ratterman and Surdam, 1981; Lee and Klein,
1986; Marsaglia and Tazaki, 1992; Ghiara et al., 1993).
Temperatures reached during diagenesis and burial
metamorphism are directly related to the geothermal gradient
and in submarine settings these range from 0C at the seafloor
to 250C at a depth of 2-10 km (Alt and Honnorez, 1984;
Morrow and Mcllreath, 1990; Alt, 1995b; Torres et al.,
1995). In modern volcanic successions, measured geothermal
gradients average 40C/km, although some are as high as
200C/km (Palmasson et al., 1979; Viereck et al., 1982).
High geothermal gradients, associated with magmatism and
regions of lithospheric extension such as back-arc basins and
rifts, can enhance diagenetic reactions by increasing reaction
rates (Boles, 1977; Surdam and Boles, 1979; Torres et al.,
1995).
The geothermal gradient may have varied in different parts
of a geosyncline or basin; it was likely to have been lowest
where the sediment was thickest and where sedimentation
occurred most rapidly (Coombs et al., 1959). Taylor et al.
(1990) proposed that examples of minimal diagenesis in some
basins may be explained by rapid sediment accumulation
rates that did not allow sufficient time for diagenetic
reactions to occur at depth or for the development of pore-
fluid gradients. In addition, magmatism provides heat to
the geothermal system, locally increasing the geothermal
gradient and compressing isograds near volcanic centres or
large intrusions (e.g. Schiffman et al., 1984; Neuhoff et al.,
1997). Coeval volcanism, plutonism and rapid burial may
establish short-lived elevated geothermal gradients in many
submarine volcanic successions. The result is low-pressure,
high-temperature diagenesis and metamorphism, and the
suppression of some facies or zones (e.g. pumpellyite-
actinolite facies, Patuki ophiolite sequence, New Zealand,
Sivell, 1984).
Definitions
The term spilite refers to an altered basalt or dolerite,
commonly porphyritic and vesicular, in which Ca-plagioclase
has been albitised and is accompanied by chlorite, calcite,
epidote, prehnite or other low-temperature hydrous minerals
typical of greenschist facies (e.g. Cann, 1969; Jolly and Smith,
1972; Grapes, 1976). Spilites are interpreted to result from
seawater-basalt interaction during diagenesis on or near
the seafloor (Coombs, 1974; Turner, 1980). Similarly, the
term keratophyre, although originally restricted to lavas, has
been applied to all felsic rocks that contain albite or albite-
oligoclase, chlorite, epidote and calcite.
5.2 | HYDRATION
Hydration of glass is typically the first stage of alteration of
volcanic facies in submarine settings and occurs during low-
temperature (<50C) seafloor weathering and the early stages
of diagenesis. Hydrated glasses (e.g. perlite or palagonite) are
very susceptible to alteration (Lipman, 1965). Hydration
facilitates subsequent reactions as it increases the alkalinity
of the pore fluid, which assists glass dissolution, promotes
crystallisation, and may produce perlitic fractures, which
further increase porosity and permeability (Lofgren, 1970;
Friedman and Long, 1984; Noh and Boles, 1989; Casey and
Bunker, 1990).
Hydration involves the diffusion of water into solid glass;
typically accompanied by a volume change (e.g. reaction R5.1
from Noh and Boles, 1989). As water is rapidly absorbed on
to glass surfaces, hydration initially affects the outer surfaces
of glassy clasts, lavas or shallow intrusions, margins along
fractures in glassy facies, pillow margins, and densely welded
pyroclastic deposits. This is followed by the slow diffusion
of water into the glass as hydration proceeds inwards along
hydration fronts defined by strain birefringence, and changes
in glass colour and refractive indices (Ross and Smith, 1955;
Friedman et al., 1966; Lofgren, 1971a). The rate of diffusion
is dependent on composition and temperature and, hence,
the extent of alteration is dependent on the time that glass
has been in contact with water (O'Keefe, 1984). Most glasses
will not undergo hydration to great thicknesses unless parallel
reactions relax the glass structure allowing water penetration
(Casey and Bunker, 1990).
dacitic glass + nH
2
O -^ perlitic glass + Na
+
+ (OH)- (R5.1)
Hydration increases the H
2
O content of glass, reorganises
the glass structure and may form palagonite or silica gels.
Changes in the glass structure may include volume changes,
and the formation of evenly spaced tiny bubbles and perlitic
fractures. Boundaries between glass and hydrated glass are
typically sharp (Peacock, 1926; Lofgren, 1971a; Fisher and
Schmincke, 1984).
Palagonite
Palagonite is a dull, resinous, yellow-orange to brown wax-
like substance formed from hydrous altered sideromelane
(basaltic) glass (Fig. 5.1). It is a mineraloid mixture of relict
hydrated glass, nontronite, montmorillonite and other sheet
silicates (Hay and Iijima, 1968b; Honnorez, 1969; Jakobsson
and Moore, 1986). Eggleton and Keller (1982) described
palagonite as a transitional alteration phase between volcanic
SE AFLOOR-AN D BURIAL-RELATED ALTERATION | 9 9
glass and smectite; however, the end product may not always
be smectite.
There are two main varieties: gel-palagonite and fibro-
palagonite (Peacock, 1926). Gel-palagonite is isotropic, dark
brown and commonly banded, forming directly adjacent to
unaltered glass (Peacock, 1926; Zhou and Fyfe, 1989). Fibro-
palagonite is orange-yellow, transparent and birefringent
(Zhou and Fyfe, 1989).
Palagonite is widespread in submarine basaltic facies and
common around the edges of glassy grains in basaltic tuffs, in
pillow rinds, along fractures in glass, and in originally glassy
vesicle walls (Moore, 1966; Baragar et al., 1977; Friedman
and Long, 1984). Partly altered basaltic pillows typically
A. Gel-palagonite in pillow basalt
The sideromelane groundmass of this plagioclase +
augite-phyric basalt is altered to yellow-brown palagonite
adjacent to the vesicle (V). The gel-palagonite exhibits
banding parallel to the vesicle wall and perpendicular
contraction cracks. Plane polarised light.
Sample 153254, Miocene Waitakere Group, Muriwai,
Northland region, New Zealand.
B. Palagonite-altered basalt clast rind
The basalt clast in this polymictic conglomerate has a
thin palagonitised rind. The plagioclase-phyric clast is
concentrically zoned with an unaltered sideromelane
core (C), yellow-brown gel-palagonite altered zone (P)
and a brown fibro-palagonite rim (R). The conglomerate
matrix includes palagonitised basaltic shards and crystal
fragments. Plane polarised light.
Sample 131562, Tertiary Macquarie Plains volcanics,
Bushy Park, Tasmania.
C. Banded palagonite
The palagonite-altered rind on this basalt clast displays
fine concentric banding. Plane polarised light.
Sample 131562, Tertiary Macquarie Plains volcanics,
Bushy Park, Tasmania.
FIGURE 5.1 | Photomicrographs of palagonite.
1 0 0 | CHAPTER 5
have glassy cores successively surrounded by concentric zones
of gel-palagonite and fibro-palagonite (+ smectite), which
are enhanced by bands of fine Fe- and Ti-oxides (Fig. 5.2:
Dimroth and Lichtblau, 1979; Zhou and Fyfe, 1989).
Palagonites have variable compositions with 10-20 wt%
H
2
O (Brey and Schmincke, 1980; Eggleton and Keller, 1982;
Pichler et al., 1999). Compared with sideromelane, Fe
2+
is
oxidised, K
2
O, FeO, TiO
2
and Cl may be locally gained, and
Na
2
O, Al
2
O
3
, SiO
2
and CaO lost (Baragar et al., 1977, 1979;
Jakobsson and Moore, 1986; Zhou and Fyfe, 1989). However,
whole-rock compositions are not significantly changed, except
for H
2
O. Palagonitisation is typically accompanied by the
growth of authigenic minerals in open pore spaces (Fig. 5.2)
and these commonly account for the elements lost from the
glass (e.g. Baragar et al., 1979; Jakobsson and Moore, 1986).
Genesis of palagonite
Zhou and Fyfe (1989) and others have proposed a two-
stage solution-precipitation mechanism for palagonitisation
of sideromelane based on physical characteristics, chemical
changes and the presence of etch or dissolution pits at alteration
fronts. The first stage is Ti constant: glass is dissolved and
gel-palagonite formed. There is a dramatic reduction in the
glass volume due to the loss of greater than 60% of the SiO
2
,
A1
2
O
3
, MgO, CaO and Na
2
O. The second stage is volume
constant: gel-palagonite is replaced by fibro-palagonite, and
zeolites begin to fill adjacent fractures and vesicles. CaO
and Na
2
O are lost, and K
2
O and SiO
2
, Al
2
O
3
and MgO are
gained from solution. Titanium and Fe
3+
are localised into
nearby fracture-filling clay and oxide minerals.
The rate of palagonitisation is temperature dependent and
doubles with every 12C increase in temperature (Jakobsson
and Moore, 1986). Palagonitisation proceeds rapidly at
temperatures above 50C and up to 150C (Jakobsson,
1972, 1978). Jakobsson and Moore (1986) noted that
palagonitisation of glass varied from less than 40% at 60C,
through 90% at 100C and was complete at temperatures
above 120C. They also found that both gel- and fibro-
palagonite occurred below 87C, but only fibro-palagonite
occurred above this temperature.
The thickness of palagonite rinds is time and temperature
dependent. Palagonite rinds in pillow basalts systematically
increase in thickness with time and doubles for every 8C
temperature increase (Moore, 1966; Jakobsson and Moore,
1986).
Perlite
Perlite is a textural term referring to networks of fine fractures
or cracks that range from concentric arcuate fractures
enclosing cores of glass (classical perlite; e.g. Fig. 5.3A and B)
to long sub-parallel fractures linked by short cross fractures
(banded or ladder perlite) (Fig. 3.2C and D: Ross and Smith,
1955; Friedman et al., 1966; Allen, 1988). Perlitic fractures
are a common feature of glassy rock fragments, felsic lavas and
synvolcanic sills, and also occur in the glassy rinds of mafic to
intermediate lavas.
Felsic perlites typically contain 2-6.5 wt% H
2
O compared
with non-hydrated obsidian, which contains a few tenths of
one percent (Ross and Smith, 1955; Noh and Boles, 1989).
In addition to gains in H
2
O, perlites typically gain K
2
O, and
lose Na
2
O and to a lesser degree CaO and SiO
2
(Lipman
et al., 1969; Fisher and Schmincke, 1984; Noh and Boles,
1989). Iron is oxidised, volatile components Cl
2
and F
2
may
be lost, and 8O
18
isotope values modified by interaction
with external fluids (Lipman, 1965; Jezek and Noble, 1978;
Cerling et al., 1985). These compositional changes are most
intense along the perlitic fractures (Jezek and Noble, 1978;
Fisher and Schmincke, 1984).
Genesis of perlite
A debate continues over the origin of perlite and the
importance of hydration (Ross and Smith, 1955; Friedman
and Smith, 1958; Friedman et al., 1966) versus cooling
contraction (Marshall, 1961; Yamagishi and Goto, 1992).
The formation of perlite is favoured by hydration of rapidly
cooled glass (i.e. glass with a high degree of under cooling)
either during cooling or later at low temperatures (Friedman
et al , 1966; Noh and Boles, 1989; Drysdale, 1991). However,
it is also possible that perlitic fractures form in response to
strain inherited from rapid cooling contraction, during the
conversion of melts to glass, and associated volume changes
(Ross and Smith, 1955; Friedman et al., 1966; Davis and
McPhie, 1996).
FIGURE. 5.2 | Sequence of palagonite alteration and zeolite cementation stages in phonolitic glass fragments (after Brey and Schmincke, 1980, in Fisher and
Schmincke, 1984). (A) Glassy shards, perhaps with montmorillonite rim cements. (B) Hydration and development of perlitic fractures accompanied by partial
dissolution and alteration of glass shards to gel-palagonite. (C) Complete dissolution and alteration of hydrated glass shards to gel-palagonite, accompanied by the
precipitation of zeolites on to glass surfaces. (D) Alteration of gel-palagonite to fibro-palagonite and precipitation of zeolites into open spaces.
SE AF L OOR-AN D BURIAL-RELATED ALTERATION | 1 0 1
A. Perlite in thin section
The glassy groundmass of this quartz latite exhibits
classical perlitic fractures comprising intersecting
and overlapping arcuate cracks. The perlitic fractures
enclose cores of unaltered and locally oxidised glass.
Arcuate glassy false shard textures occur where perlitic
fractures intersect (arrow). Amygdales have been filled
with zeolites. Plane polarised light.
Sample ET7-4, Wereldsend Formation, Pilchard Gorge,
Etendeka, Namibia.
B. Perlite in partly altered rhyolite
Well-developed perlitic fractures are abundant in this
partly glassy rhyolite. The perlitic fractures have been
lined with fine-grained, dark green to brown smectites,
enhancing the fracture pattern. Perlite cores have been
partly altered to smectites and zeolites. Amygdales have
been filled with cristobalite. Plane polarised light.
Sample 147582, Miocene Nishikurosawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
C. Relict perlite and amygdales in altered rhyolite
In this diagenetically altered rhyolite, relict perlitic
fractures are conspicuous where glass adjacent to the
fractures has been altered to dark green mixed layer
smectite-chlorite. Elsewhere in the pervasively zeolite
altered domains the perlitic fractures have been obscured.
The amygdales have been filled with layers of cristobalite
and fibrous chlorite. Plane polarised light.
Sample J6-735 m, Miocene Nishikurosawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
D. Relict perlite in altered basalt
In this hydro thermally altered jigsaw-fit basaltic breccia,
perlitic fractures are only weakly discernable due to
multiple overprinting alteration facies. The pervasive
sericite + quartz + pyrite and nodular carbonate
alteration facies obscure the perlitic fracture pattern.
Plane polarised light.
Sample 76833, Cambrian Que-Hellyer Volcanics, western
volcanosedimentary sequences, Mount Read Volcanics,
western Tasmania.
FIGURE 5.3 | Photomicrographs of fresh and altered perlite.
1 0 2 | CHAPTER 5
Alteration of perlite
Perlite commonly undergoes subsequent alteration to
diagenetic mineral assemblages that include smectite, Fe-
oxides, zeolites, K-rich gel-like glass, low-cristobalite, K-
feldspar, chlorite, sericite and carbonate (Noh and Boles,
1989). Alteration begins by dissolution of hydrated glass
and crystallisation of smectite, carbonate or Fe-oxides along
perlitic fractures (e.g. Noh and Boles, 1989). This commonly
accentuates the fracture pattern (e.g. Fig. 5.3B). As alteration
progresses, glass dissolution with continued precipitation
advances inwards and the perlitic fractures become diffuse and
indistinct (e.g. Fig. 5.3C, D and Allen, 1988). Dissolution of
remaining glassy cores is succeeded by formation of zeolites,
such as clinoptilolite or mordenite, or gel-like glass, which
are ultimately replaced by K-feldspar (e.g. Noh and Boles,
1989).
5.3 | DIAGENESIS (GLASS TO ZEOLITE
FACIES)
Diagenesis encompasses the low-temperature and low-
pressure alteration processes that occur during progressive
burial of sediments and rocks. It can be defined as the
processes (excluding weathering) that change their character
and composition, between the moment of deposition, and
the onset of metamorphism (Larsen and Chilingar, 1979).
Submarine diagenesis involves low-temperature processes,
ranging from bottom water temperatures up to crystallisation
of unequivocally metamorphic minerals such as laumontite,
wairakite, chlorite and pumpellyite (Winkler, 1979; Bohlke
et al., 1980). It is impossible to define a unique pressure and
temperature range that would characterise the transition
between diagenesis and metamorphism, because of the greatly
contrasting degrees of mineral stability that characterise
different rock types and the wide range of conditions under
which the common diagenetic minerals crystallise. Generally,
diagenesis in submarine settings occurs at pressures of 0.1 to
10 MPa (1 bar to 1 kbar) and temperatures ranging from 0 to
250C (Alt and Honnorez, 1984; Morrow and Mcllreath,
1990; Alt, 1995b). Temperatures and pore water salinities
increase, and seawater-rock ratios decrease with burial depth
(Hanor, 1979; Alt-Epping and Smith, 1997).
Submarine diagenesis encompasses compaction,
dissolution and leaching of components, precipitation of
new minerals, and recrystallisation in response to changes
in pressure, temperature and chemical conditions in the
subseafloor. New minerals directly replace glass, form mineral
overgrowths, fill primary and secondary pore spaces, and
form cements, all of which dramatically reduce the porosity
and permeability and promote lithification.
With increasing diagenesis, porosity and permeability
typically decrease. However, reversals in this trend can
occur during fracturing or if a major component of the rock
becomes under saturated and secondary porosity is formed by
dissolution. This can occur where deeply buried sediments are
infiltrated by fresh or brackish ground water, or can be due
to the release of water of crystallisation from clay minerals
(Morrow and Mcllreath, 1990).
The process of dissolution involves corrosion or leaching
of pre-existing phases (either glass or mineral phases), with or
without minor replacement by new minerals (Morrow and
Mcllreath, 1990). It is a complex process involving many
distinct reaction steps and pathways. It can modify glass and
most primary igneous minerals. Dissolution may ultimately
lead to the formation of secondary porosity (e.g. dissolution
vugs), replacement of glass and minerals, and development
of solution seams or stylolites (Amstutz and Park, 1967;
Marsaglia and Tazaki, 1992).
Despite changes in mineral assemblage, many pre-existing
textures (primary volcanic, high-temperature devitrification
and hydration textures) are preserved and sometimes enhanced
during diagenesis. Figure 5.4 shows some examples of textures
in unaltered volcanic rocks, and their diagenetically altered
and in some cases metamorphosed equivalents.
Submarine diagenesis may involve multiple stages or
episodes of diagenesis (Bohlke et al., 1980; Morrow and
Mcllreath, 1990). Diagenesis of most ancient sedimentary
successions involved repeated exposure to diagenetic realms
as they underwent cycles of subsidence and uplift. Generally,
however, the imprint of the first stages of diagenesis is
preserved because of the large initial porosity reduction and
lithification (Morrow and Mcllreath, 1990).
Diagenetic minerals
There are three main types of minerals typical of seafloor
weathering and diagenesis in volcanic successions: layered
silicates, zeolites and carbonates. Figure 5.5 provides estimates
of their formation temperatures.
Layered silicates
The layered silicates include clay minerals, mixed-layered
minerals, micas, chlorite and prehnite. The common clay
minerals in volcanic facies can be divided in to two groups:
(1) smectites (e.g. montmorillonite, nontronite andsaponite),
and (2) illite group clay minerals (e.g. celadonite, glauconite
and illite).
Smectites are swelling clay minerals that readily exchange
Ca and Na cations. They typically result from the alteration
of volcanic grains under alkaline conditions where Mg and
Ca ions are available (Deer et al., 1966). Smectites form rims
on glass surfaces, replace both felsic and mafic glass, and
pseudomorph glass shards and olivine crystals (Sheppard and
Gude, 1968; Schmincke and von Rad, 1976; Viereck et al.,
1982). Smectites initially forms blebs and web-like arrays on
glass surfaces, and become better crystallised as diagenesis
proceeds (Hein and Scholl, 1978). The term bentonite refers
to felsic tuff that is composed of almost pure smectite (Gary
etal., 1974).
In contrast, the illite group are K- and Al-rich minerals
that typically form in neutral to alkaline conditions from the
breakdown of feldspars and micas (Deer et al., 1966). They
typically occur as vesicle fill and pseudomorphs of felsic glass
shards and pumice (Schmincke and von Rad, 1976; Iijima,
1978). Celadonite and glauconite are less common than
illite.
SE AF L OOR-AN D BURIAL-RELATED ALTERATION I 1 0 3
A. Flow banding
This devitrified flow-banded plagioclase-phyric rhyolite
contains alternating dark and light flow bands. The dark
bands are dominantly obsidian, whereas the pale bands
contain fine spherulites and lithophysae.
Sample NG1, < 140 ka Ngongotaha lava dome, Hendersons
quarry, Rotorua, New Zealand.
B. This diagenetically altered and metamorphosed flow-
banded plagioclase-phyric rhyolite contains alternating
orange albite + quartz and grey sericite-rich bands. In
thin section, the orange bands contain relict spherulites,
whereas the grey bands are microcrystalline.
Sample 147481, Cambrian Central Volcanic Complex,
Mount Read Volcanics, Mount Block, western Tasmania.
C. Spherulites
In thin section, fresh spherulites consist of radial crystal
fibres; typically feldspar intergrown with cristobalite,
tridymite or clinopyroxene. Many of these spherulites
enclose plagioclase phenocrysts and are separated by
small cuspate lenses of dark brown obsidian. Plane
polarised light.
Sample NG4, <14O ka Ngongotaha lava dome, Hendersons
quarry, Rotorua, New Zealand.
D. Recrystallised spherulites in this greenschist facies
rhyolite are composed of albite, quartz and sericite.
Fine sericite trails preserve a radial pattern within the
spherulites. The boundaries between the spherulites are
marked by concentrations of sericite. Plane polarised
light.
Sample 147528, Cambrian Central Volcanic Complex,
Mount Read Volcanics, Mount Black, western Tasmania.
E. Tube pumice clasts
This unaltered, semi-consolidated, dacitic pumice
breccia contains glassy tube pumice clasts and plagioclase
crystals in a matrix of fine glass shards. The pumice \
clast pictured here displays a fine fibrous texture, which
may be preserved during subsequent alteration. Plane
polarised light.
Sample from the -1 Ma trachydacitic pumice breccias,
Efate Pumice Tormation, Vanuatu.
FIGURE 5.4 | Photographs of unaltered and diagenetically altered volcanic textures.
1 04 | CHAPTER 5
F. Pumice clasts in this diagenetically-altered and
metamorphosed rhyolitic pumice breccia preserve the
fine tube vesicle structure. The originally glassy vesicle
walls have been altered to albite + quartz + hematite, the
vesicles have been lined with sericite and filled with albite.
Plagioclase crystals in this sample have been completely
replaced by albite and hematite. Plane polarised light.
Sample 133815, Cambrian Hercules Pumice Formation,
Central Volcanic Complex, Mount Read Volcanics, Hercules
footwall, western Tasmania.
G. Many tube pumice clasts locally preserve round
vesicles adjacent to phenocrysts. In this diagenetically
altered pumice breccia, round and tube vesicles adjacent
to a cluster of plagioclase phenocrysts have been filled
with mordenite. As a result, the vesicles have retained
their shapes during burial compaction. Plane polarised
light.
Sample OH8-369 m, Miocene Onnagawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
H. Similarly, this pumice breccia, which has been
diagenetically altered and metamorphosed to greenschist
facies, contains round and tube vesicles adjacent to
hematite-altered plagioclase phenocrysts. The vesicles
(V) have beeen filled with sericite and albite. Plane
polarised light.
Sample 147499, Cambrian Kershaw Pumice Formation,
Central Volcanic Complex, Mount Read Volcanics, east
Hercules, western Tasmania.
I. Palagonitised rinds on clasts
The rim of this basalt clast has been altered to orange-
brown palagonite. Palagonite has also formed rims
around the vesicles in the clast. Plane polarised light.
Sample 131562, Tertiary Macquarie Plains volcanics,
Bushy Park, Tasmania.
J. The sericite + albite + hematite-altered rim (R) on this
basalt clast may be the metamorphosed equivalent of a
palagonite-altered rind. Plane polarised light.
Sample 147572, Cambrian Sterling Valley Volcanics,
Central Volcanic Complex, Mount Read Volcanics, Sterling
Valley, western Tasmania.
FIGURE 5.4 | Photographs of unaltered and diagenetically altered volcanic textures, cont.
SE AFLOOR-AN D BURIAL-RELATED ALTERATION | 1 0 5
Carbonates
FIGURE 5.5 | Temperature estimates for the growth of common diagenetic and
burial metamorphic minerals, and palagonite (data from Thompson, 1971; Seki,
1972; Merino, 1975; Grapes, 1976 ; Kastnerand Gieskes, 1976 ; Seyfried and
Bischoff, 1979; Bohlke et al., 1980; Munha et al., 1980; Boles, 1982; Viereck et
al., 1982; Jakobsson and Moore, 1986 ; Bish and Aronson, 1993; Ogihara, 1996 ;
Ylagan et al., 1996 ; Bodon and Cooke, 1998).
Zeolites
Zeolites are hydrous Al-silicates containing Na and Ca
(Table 5.1). The most common zeolites in marine settings
are clinoptilolite, mordenite, phillipsite and analcime
(Marsaglia and Tazaki, 1992). A variety of fibro-radiated and
bladed zeolites fill pore spaces, cement volcaniclastic particles
and replace glass in altered volcanic facies (Miyashiro and
Shido, 1970; Schmincke and von Rad, 1976). Most zeolites
precipitate in open space on to smectite or chlorite films or
occur as overgrowths on detrital grains such as plagioclase
crystal fragments (e.g. Schmincke and von Rad, 1976). Others
crystallise directly from glass via dissolution reactions with
smectite (e.g. Noh and Boles, 1989) and may pseudomorph
glass shards (e.g. Walton, 1975).
Diagenetic carbonates are dominantly calcite and dolomite.
They typically fill originally open spaces such as vesicles, occur
as cements in volcaniclastic facies (e.g. Hay, 1977), as spheroids
or nodules, and as euhedral crystals replacing palagonite (e.g.
Dimroth and Lichtblau, 1979), rock fragments, olivine and
plagioclase crystals.
Other diagenetic minerals
Other diagenetic minerals include silica phases (e.g. low-
cristobalite, opal CT, chert and quartz), Fe-oxides (e.g.
hematite), Ti-rich minerals (e.g. leucoxene), anhydrite, pyrite,
epidote and feldspars (albite and K-feldspar). These mainly
replace glass, primary crystal phases and earlier alteration
minerals. Silica phases and feldspars also occur as overgrowths
on primary plagioclase and quartz crystals (e.g. Noh and
Boles, 1989; Tsolis-Katagas and Katagas, 1989).
Diagenetic zones
Diagenetic mineral assemblages commonly show a thick
vertical zonation (e.g. Fig. 5.6 and Section 5.5). Diagenetic
zones have been described by a number of authors in modern
and ancient submarine felsic to intermediate volcanic
successions (e.g. Iijima, 1974; Walton, 1975; Iijima, 1978;
Ratterman and Surdam, 1981; Sheppard et al., 1988;
Williams et al., 1989; Utada, 1991; Passaglia et al., 1995;
Ogihara, 1996). Sequences of diagenetic zones are between
500 m and 6 km thick, with individual altered zones varying
from a few metres to several kilometres in thickness. This
vertical zonation corresponds to progressive mineral reactions
that occur in response to changes in pore water chemistry and
temperature with depth of burial, and is very similar to burial
metamorphism (Coombs, 1954). Some altered zones may be
absent or combined.
Diagenetic zones in felsic volcanic successions
Diagenetic zones in felsic volcanic successions can be grouped
into four main zones (Table 5.2): (I) partially altered zones,
(II) alkali-rich zeolite zones, (III) late-stage zeolite + calcite
zones, and (IV) albite zones. At depth Zone IV may pass in to
a prehnite + pumpellyite zone, which represents the transition
to greenschist facies metamorphic zones (Iijima, 1974, 1978;
Utada, 1991).
Partially altered zones are characterised by silica and
clay minerals, they lack zeolites, contain unaltered and partly
altered glass, and unaltered primary minerals such as plagioclase
(Iijima, 1974, 1978). Alteration mineral assemblages are
dominated by smectites (commonly montmorillonite) + low-
cristobalite or opal-CT (Iijima, 1974, 1978; Walton, 1975;
Sheppard et al., 1988; Passaglia et al., 1995). Primary pore
spaces, such as vesicles, have typically been partially filled with
low-cristobalite, glassy clasts have been coated in thin films of
smectite, and some originally glassy shards and pumice clasts
altered to smectite. Coherent facies were relatively unaltered.
1 06 | CHAPTER 5
TABLE 5.1 | Common zeolites and their occurrences in submarine volcanic facies. Zeolite formulas are from Deer et al. (196 6 ).
A Na-rich late stage zeolite, which replaces earlier alkali zeolites in both coherent and clastic
volcanic facies of rhyolitic to basaltic composition (e.g. lijima, 1974; Ratterman and Surdam, 1981;
Torres et al., 1995)
Analcime Na[AISi
2
O
6
].H
2
O
Chabazite Ca[ AI
2
Si
4
0
12
].6 H
2
0 -
thomsonite NaCa
2
[(AI,Si)
5
O
10
]
2
.6 H
2
O
Clinoptilolite (Na,K)
4
CaAI
6
Si
30
O
72
.H
2
O
Heulandite (Ca,Na
2
)[AI
2
Si
7
O
18
].6 H
2
O
Laumontite Ca[ AI
2
Si
4
0
12
].4H
2
0
Mordenite (Ca, Na
2
,K
2
)[AI
2
Si
10
O
24
].7H
2
O
Phillipsite (1/2Ca,Na,K)
3
[AI
3
Si
5
O
16
].6 H
2
O
Wairakite CaAI
2
Si
4
0
12
.H
2
0
Restricted to mafic facies, typically replacing palagonite (e.g. Brey and Schmincke, 1980; Dimroth
and Lichtblau, 1979)
Occurs as a cement and replaces glass in felsic volcanic facies (e.g. Noh and Boles, 1989;
Ratterman and Surdam, 1981; Torres et al., 1995)
Occurs as cements in felsic volcaniclastic facies (e.g. Ratterman and Surdam, 1981)
A calcic zeolite, which occurs at depth in originally glassy felsic volcanic facies
Only derived from felsic volcanic facies and commonly coexists with smectite and silica phases
(i.e. opal, quartz, tridymite and cristobalite) (Noh and Boles, 1989; Ratterman and Surdam,
1981; Sheppard et al., 1988; Sheppard and Gude, 196 8; Torres et al , 1995; Tsolis-Katagas and
Katagas, 1989; Utada, 1970)
Occurs mainly in basaltic lavas and less commonly in volcaniclastic facies where it replaces
basaltic glass and palagonite (Taylor and Surdam, 1981), it commonly contains inclusions of Fe-
oxyhydroxides and smectites (Brey and Schmincke, 1980)
A common alteration product in basaltic facies at depth in modern geothermal systems (e.g. Boles,
1977; Hay, 1977)
Table 5.2 | Common diagenetic zones and their alteration mineral assemblages for thick submarine volcanic successions.
Zone I: partially altered zone
unaltered glass + smectite (montmorillonite) + low-cristobalite/opal-CT
Zone II: alkali-rich zeolite zone
(a) clinoptilolite + smectite (montmorillonite) + low-cristobalite/opal-CT
(b) Ca-clinoptilolite + mordenite + smectite + K-feldspar quartz
Zone III: late stage zeolite + calcite zone
(a) analcime + heulandite + clacite + phyllosilicate minerals (smectite,
chlorite, mixed layer minerals) + K-feldspar quartz pyrite
(b) analcime + laumontite + clacite + phyllosilicate minerals (illite,
chlorite, smectite) + K-feldspar + quartz
Zone IV: albite zone
albite + phyllosilicate minerals (prehenite, pumpellyite, chlorite,
sericite) + quartz K-feldspar laumontite calcite
Zone I: partially altered zone
unaltered glass + palagonite + smectite + illite + low-cristobalite/
adularia + Fe/Mn/Ti oxides + unaltered glass
Zone II: calcic-zeolite zone
phillipsite/chabzite + phyllosilicate minerals (chlorite, smectite,
sericite) + Fe/Mn/Ti-oxides K-feidspar
Zone III: late-stage zeolite zone
analcime natrolite ( heulandite laumontite) + chlorite + K-
feldspar + Fe/Mn/Ti-oxides calcite
Zone IV: epidote zone
epidote + chlorite + albite + calcite + sphene prehnite
SEAFLOOR-AND BURIAL-RELATED ALTERATION I 1 07
FIGURE. 5.6 | East-west schematic cross-sections showing the depth distribution of regional diagenetic zones and local
hydrothermal zones associated with the Kuroko deposits in the Green Tuff Belt, Japan (after lijima, 1974,1978). (A) Odate
Basin, Hokoroku district. (B) Odate to Hanawa Basin, Hokoroku district. (C) Diagenetic zones in the Neogene and Palaeogene
formations of Hokkaido. Traces of cross-sections A and B are shown on the regional map of the Hokuroku Basin (Fig. 5.15).
1 0 8 | CHAPTER 5
These early clay-rich zones are associated with minor
initial gains in K
2
O and Al
2
O
3
, and losses in Na
2
O and, to
lesser degrees, CaO and SiO
2
(Noh and Boles, 1989)
Alkali-rich zeolite zones are commonly characterised by
assemblages of clinoptilolite + mordenite + smectite (typically
montmorillonite, saponite or mixed-layer illite/smectite) +
low-cristobalite quartz opal K-feldspar. Opal-CT and
low-cristobalite occur in the upper parts of these zones, whereas
quartz and K-feldspar occur in the lower parts (Walton, 1975;
Sheppard et al., 1988). Plagioclase is rarely altered.
In general, clinoptilolite and mordenite has filled pore
spaces such as primary vesicles and dissolution voids. Glassy
clasts have been coated in montmorillonite or silica rims and
completely altered to zeolites (e.g. mordenite), low-cristobalite,
quartz, clay minerals and K-feldspar (Iijima and Utada, 1971;
Iijima, 1974, 1978). The glassy groundmass of lavas and sills,
and the cores of blocky clasts may have been partly altered.
Pumice-rich facies in these zones contain dark green, variably
flattened phyllosilicate-rich fiamme (e.g. saponite fiamme
in pumice breccia in the Hokuroku Basin, Iijima, 1974).
Tuffaceous mudstones may be rich in montmorillonite, low
cristobalite and quartz.
Alteration to alkali-zeolites and phyllosilicate minerals
resulted in whole-rock gains of MgO and Fe
2
O
3
, and losses
of SiO
2
, Na
2
O, K
2
O, and variable changes in CaO (Noh and
Boles, 1989; Tsolis-Katagas and Katagas, 1989; Passaglia et
al., 1995).
Late-stage zeolite + calcite zones are characterised by
mineral assemblages containing analcime and calcite (Iijima,
1974). In some successions, two late-stage zeolite + calcite
zones have been defined (e.g. Iijima, 1978): (a) analcime +
heulandite + calcite phyllosilicate minerals K-feldspar
quartz pyrite, and (b) analcime + laumontite + calcite
chlorite illite sericite K-feldspar. The phyllosilicate
minerals are typically smectites, chlorite and mixed-layer
minerals such as illite/smectite, saponite/chlorite or swelling
chlorite. These mineral assemblages may also contain relict
clinoptilolite and/or mordenite (Iijima, 1974; Walton, 1975;
Sheppard et al., 1988).
Plagioclase phenocrysts have remained unaltered or
have been analcime calcite altered. Analcime has replaced
mordenite- or clinoptilolite-altered felsic glass fragments
and pumice clasts. Saponite, smectite, chlorite and mixed-
layer mineral fiamme are typically common in pumice-rich
rocks. Calcite may occur as euhedral crystals, concretions or
veinlets.
Albite zones are commonly characterised by albite
+ laumontite calcite + prehnite chlorite sericite
pumpellyite quartz K-feldspar (Iijima and Utada, 1971;
Iijima, 1974). Plagioclase phenocrysts have been extensively
albitised, and albite + laumontite have replaced plagioclase
crystals, originally glassy shards and pumice clasts, and filled
pore spaces.
Mass gains in CaO, SiO
2
, Na
2
O, Sr and Ba in these zones
are consistent with seafloor albitisation (Boles and Coombs,
1977; Boles, 1982).
Diagenetic zones in mafic volcanic successions
Diagenetic mineral assemblages in mafic volcanic successions
contain palagonite, several species of calcic zeolites, Fe/Ti/Mn-
oxides and abundant clay minerals of the smectite-chlorite
series typically distributed in four zones (Table 5.2: Baragar et
al., 1979; Zhou and Fyfe, 1989; Utada, 1991).
Partially altered zones contain some fresh basaltic glass
and have alteration mineral assemblages of palagonite Fe/
Mn/Ti-oxides (e.g. maghemite or magnetite) + clay minerals
(smectites, illites and mixed-layer minerals) low-cristobalite.
Compositional changes include major gains of H
2
O, very
minor gains of K
2
O, FeO, TiO
2
and Cl and losses of Na
2
O,
A1
2
O
3
, SiO
2
and CaO (Baragar et al., 1977, 1979; Jakobsson
and Moore, 1986; Zhou and Fyfe, 1989).
Calcic-zeolite zones are characterised by phillipsite
or chabazite chlorite + smectite + Fe/Mn/Ti-oxides + K-
feldspar.
Whole-rock gains in MgO in these zones are consistent
with the formation of smectite, chlorite and other Mg-
silicates during diagenesis (cf. Hajash and Chandler, 1981;
Shiraki and Iiyama, 1990).
Late-stage zeolite zones are characterised by analcime
laumontite natrolite chabazite + heulandite mesolite +
chlorite + Fe/Mn/Ti-oxides + K-feldspar.
Epidote zones are characterised by epidote + chlorite +
albite + sphene calcite prehnite.
Genesis of diagenetic minerals and zones
In submarine volcanic facies, dissolution, cementation
and lithification begin shortly after deposition (<1 Ma);
however, major diagenetic changes develop through a series
of recognisable stages over tens of millions of years (Marsaglia
and Tazaki, 1992). The paragenesis from glass to smectites to
alkali zeolites may be explained by a sequence of hydration
and dissolution reactions in most glass bearing rocks (Fig.
5.7). Later reactions involve transitions from less stable to
more stable mineral assemblages, such as clinoptilolite to
Na-clinoptilolite + mordenite or K-rich gel-like glass to K-
feldspar (Noh and Boles, 1989).
There are four stages of clastic diagenesis after initial
hydration and oxidation (e.g. Fig. 5.8): (1) formation of
clay mineral rims on glassy surfaces, (2) partial to complete
dissolution of glass and compaction, (3) precipitation of
authigenic minerals, especially zeolites and calcite, in open
pore spaces, and (4) alteration and replacement of mineral
phases (Hay, 1963; Fisher and Schmincke, 1984; Pichler et
al., 1999). Stages two and three may overlap.
Stage 1 : coating surfaces
The initial stage of diagenesis in volcaniclastic facies is
characterised by the precipitation of thin rim cements, which
coat all originally glassy surfaces and some crystal surfaces
(Fig. 5.9A and B). Rim cements help to preserve shard and
clast outlines during subsequent replacement (e.g. Walton,
1975). Rim cements may be accompanied by dissolution of
intermediate to mafic glass, alteration of rhyolitic glass to clay
minerals and minor precipitation of calcite or clinoptilolite
cements (Fig. 5.9C and D: Marsaglia and Tazaki, 1992;
Torres et al., 1995).
In felsic volcaniclastic facies, the outer walls of most glass
shards and vesicles in pumice clasts are lined with thin films of
smectite, calcite, opal or rarely chlorite (e.g. Henneberger and
Browne, 1988; Sheppard et al., 1988; Noh and Boles, 1989;
Tsolis-Katagas and Katagas, 1989; Marsaglia and Tazaki,
1992; Torres et al., 1995). Only rarely are fine-grained, glassy
fragments such as shards and pumice completely replaced by
smectite. Smectite rims probably precipitate from alkaline
fluids during the dissolution of hydrated glass surfaces. This
may follow the reactions:
SE AFLOOR-AN D BURIAL-RELATED ALTERATION | 1 0 9
and high activities of Mg (Hay, 1978; Hajash and Chandler,
1981).
The initial stage of diagenesis in basalts involves the
palagonitisation of basaltic glass. Palagonitisation initiates on
glass surfaces, along fractures and around vesicles in a similar
way to the thin smectite coating on rhyolitic glass shards (e.g.
Zhou and Fyfe, 1989). Associated with palagonitisation is the
precipitation of Fe- and Ti-oxides, which coat all surfaces,
thermal contraction cracks, vesicles, and the palagonitisation
front (e.g. Dimroth and Lichtblau, 1979).
Stage 2: dissolution of glass and compaction
perlitic glass + 3.88K
+
+ 0.65H+ + 15.4H
2
O -
smectite + 9.5gel-like glass + 4.03Na
+
+ 0.25Ca
2+
+
10.55H
4
SiO
4
(Noh and Boles, 1989) (R5.2)
or
rhyolitic glass + Mg
2+
+ H
2
O
Na-Ca montmorillonite + SiO
2
+ Na
+
+ K
+
+ Fe
2+
(R5.3)
The formation of smectite and other Mg-silicates during
rhyolite-seawater interaction does not require significant gains
in alkalis, but is favoured by high ratios of H/Na and K/Ca,
Large-scale dissolution of glass and crystals is typically
accompanied and followed by the precipitation of authigenic
mineral cements and lithification after a few million years
(Marsaglia and Tazaki, 1992; Torres et al., 1995). The
dissolution of glass fragments, olivine and amphiboles occurs
rapidly at shallow burial depths prior to extensive cementation
and lithification (Smith, 1991; Marsaglia and Tazaki, 1992).
With increasing depths of burial, dissolution of feldspar
microlites and the glassy groundmasses of coherent facies
occurs (Marsaglia and Tazaki, 1992). In contrast, plagioclase
crystal fragments and phenocrysts undergo only minor
dissolution early in the diagenetic history.
Elements leached from the glass during dissolution
reactions are consumed by the formation of new minerals.
FIGURE. 5.7 | Flow diagrams showing the
successive development of alteration mineral
assemblages in volcanic glass during diagenesis.
(A) Alteration of silicic glass to day minerals,
zeolites and silicates (after Hay, 1978; lijima, 1978;
Utada, 1991). (B) Alteration of basaltic glass to
palagonite, clay minerals, zeolites and oxides
(Honnorez, 1978; lijima, 1978; after Brey and
Schmincke, 1980; Viereck et al., 1982; Fisher and
Schmincke, 1984).
FIGURE 5.8 | Schematic model for the microscopic textural evolution and reduction in porosity in non-welded pumice breccias during diagenesis (after Gifkins,
2001). (A) Stage 1: thin films of smectite (green lines) coat original surfaces, such as vesicle walls, crystals, shards and lithic clasts. (B) Stage 2: primary porosity
is filled, and originally glassy shards and vesicle walls replaced or partly replaced by zeolites (bronze), clay minerals (green) or carbonates. Zeolite or K-feldspar
overgrowths (orange) may develop on plagioclase crystals. (C) Stage 3: glass is dissolved, altered to clay minerals and compacted, producing phyllosilicate-rich
fiamme. Clays, zeolites and Fe-oxides precipitate synchronously with the dissolution of glass, forming stylolites. After compaction, more stable diagenetic or
metamorphic minerals replace any remaining glass and less stable minerals (i.e. Stage 4).
For example, Na released by the hydration and dissolution of
rhyolitic glass (reactions 5.2 and 5.3) may be consumed by
the precipitation of mordenite in vesicles (e.g. Fig.5.10E) or
dissolution vugs (Sheppard et al., 1988).
Dissolution may be accompanied by compaction that
reduces the pore space geometry by rotating grains, deforming
soft grains and crushing grains. This promotes lithification in
volcaniclastic facies by pressure welding clasts so that their
margins interpenetrate (Taylor et al., 1990; Marsaglia and
Tazaki, 1992). Generally, lithification of volcaniclastic facies
is associated with the formation of diagenetic clay mineral
or carbonate rims, which act as cohesive binders: cements.
However, Marsaglia and Tazaki (1992) in their study of
modern partially altered sandstones at ODP Site 788 (Japan)
suggested that lithification could be related to a combination
of compaction and brown glass dissolution. Sandstones in
a transitional zone, between the unlithified and cemented
zones, appeared to be lithified as a result of compaction and/
or pressure welding with minor cementation by phillipsite
and smectite/chlorite rim cements. The rim cements formed
where sufficient glass had dissolved to produce favourable
conditions for smectite or zeolite precipitation.
Compaction may also bend and flatten clasts, particularly
pumice clasts. Gifkins etal. (in press) suggested that compaction
during burial of clay-altered pumice clasts flattened the clasts
and modified tube vesicle structures resulting in bedding-
parallel phyllosilicate lenses (i.e. fiamme, Fig. 5.9G). They
also proposed that stylolites in pumice breccias in the Mount
Read Volcanics (western Tasmania) and the Green Tuff Belt
(Japan) resulted from the dissolution of soluble components,
particularly glass, and the precipitation of clays and Fe-oxides
during compaction (Fig. 5.9H).
Stage 3: filling pore space and cementation
Precipitation of low-cristobalite and zeolites as pore-fill
cements follows the early rim cements in both felsic and
mafic volcanic facies (e.g. Klein and Lee, 1984; Zhou and
Fyfe, 1989). Zeolites also fill vesicles and dissolution voids in
glass, and directly replace glass, forming shard pseudomorphs
or altering the glassy cores of perlite (Dimroth and Lichtblau,
1979; Noh and Boles, 1989; Tsolis-Katagas and Katagas,
1989; Passaglia et al., 1995). Many of the zeolites filling
vesicles have fibrous radiating textures (e.g. mordenite,
Fig. 5.10A and B), which may be partly preserved during
subsequent metamorphic recrystallisation (e.g. Fig. 5.IOC
and D). Dimroth and Lichtblau (1979) described fibro-
radial textures defined by a dusting of fine oxides in Archaean
basaltic hyaloclastite of the Noranda District (Canada), which
suggest the former presence of fibro-palagonite or zeolites.
The formation of alkali-rich zeolites as pore-fill cements
involves hydration and dissolution of glass by saline, alkaline
solutions (Ratterman and Surdam, 1981; Noh and Boles,
1989). For example, the formation of clinoptilolite from
hydrated felsic glass in reaction R5-4, which consumes Ca
and Si released during reaction R5.1:
perlitic glass + 0.1Ca
2
+ + 0.1H
4
SiO
4
+ 0.1 H
+
+ H
2
O -
clinoptilolite + 0.1K
+
+ 0.2Na
+
(R5.4)
The formation of rim and pore-fill cements results in a
lithified rock. Cements reduce the porosity, strengthen the
grain framework and reduce the amount of compaction.
The initial 35-40% porosity of a well-sorted sandstone can
be reduced to 15-20% by early clay mineral, carbonate or
zeolite cements (Helmold and van de Kamp, 1984). In the
clinoptilolite + mordenite zone, Henneberger and Browne
(1988) found the porosity of pumice breccias was reduced
by half, from 34-47% to 20-50%. Alteration in the quartz +
adularia zone further reduced porosity to 423%.
SEAFLOOR- AND BURIAL-RELATED ALTERATION I 1 1 1
A. Clay rim cement in pumice breccia
Smectite films on all glass and crystal surfaces record
the initial stage of diagenesis in this partially altered
rhyolitic pumice breccia. Green-brown smectite has
coated bubble-wall shards, plagioclase and quartz crystal
fragments, and lined vesicles. Some originally glassy
shards have been completely replaced by smectite;
however, larger clasts are still glassy (G).
Sample J6-295 m, Miocene Onnagawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
B. Clay-lined vesicles
Round and elongate vesicles (V) in this pumice clast are
coated in irregular, fine-grained, pale brown smectite
films. Small vesicles have been completely filled wih
smectite, larger vesicles are unfilled, and the originally
glassy vesicle walls have been altered to mordenite.
Sample FK5B, Miocene Tokiwa Formation, South Fossa
Magna, Green Tuff Belt, Odawara, Japan.
C. Pore-filling cements
In this pumice breccia sample, calcite cement binds the
unaltered glassy and partly calcite-altered tube pumice
clasts. Plane polarised light.
Sample Y2A, Quaternary Yali pumice breccia, Yali Island,
eastern Aegean, Greece.
D. In crossed nicols, the glassy pumice clasts are isotropic
and the calcite cement, calcite-filled tube vesicles, and
altered shards and pumice clasts are evident.
FIGURE 5.9 | Examples of textures that record the different steps in the evolution of pumice clasts during diagenesis.
1 1 2 I CHAPTER 5
E. Zeolite-filled vesicles in pumice
The smectite-lined vesicles (V) in this pumice clast have
been infilled with layered fibrous zeolites: mordenite
and clinoptilolite. Originally glassy shards have been
altered to smectite and vesicle walls to mordenite +
smectite. Fine-grained nodules of analcime overprinted
the mordenite and smectite altered tube pumice clast
(P). Plagioclase and quartz crystals are unaltered.
Sample OH8-537 m, Miocene Onnagawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
F. Clay-altered pumice clast
Other pumice ciasts may be completely altered to clay
minerals, like this dark green uncompacted smectite-
altered pumice. Shards and fine-grained ciasts in the
matrix have been altered to smectites (montmorillonite
and saponite) + mordenite.
Sample OH8-387 m, Miocene Onnagawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
G. Clay-altered and compacted pumice
During burial, lithostatic pressure may lead to the
flattening of soft clay-altered pumice ciasts. The mixed
layer smectite-chlorite fiamme (F) in this pumice
and lithic breccia roughly define a bedding-parallel
compaction fabric. The fiamme have a fibrous internal
texture, wispy terminations and flame-like shapes.
Some fiamme are also plagioclase porphyritic. They are
interpreted to be diagenetically altered and compacted
pumice ciasts.
Sample 147583, Miocene Nishikurosawa Formation,
Hokuroku Basin, Green Tuff Belt, Odate, Japan.
H. Dissolution fabrics in pumice breccia
The dissolution of glass commonly accompanies
compaction during diagenesis. Solution seams and
stylolites, like the one pictured here, are interpreted
to record the dissolution of soluble components. This
stylolite is an anastomosing sutured structure that
concentrates clay minerals and oxides.
Sample FK7, Miocene Wadaira Tuff Member, Tokiwa
Formation, South Fossa Magna, Green Tuff Belt, Wadaira,
Japan.
FIGURE 5.9 | Examples of textures that record the different steps in the evolution of pumice ciasts during diagenesis, cont.
SE AF L OOR-AN D BURIAL-RELATED ALTERATION | 1 1 3
A. Fibrous zeolites in vesicles B. In crossed nicols, radial extinction patterns accentuate the
Vesicles adjacent to this plagioclase crystal in a pumice clast fibrous nature of the vesicle-filling zeolites.
have been lined with smectite and filled with fibrous radiating
mordenite. Plane polarised light.
Sample 147580, Miocene Onnagawa Formation, Hokuroku
Basin, Green Tuff Belt, Odate, Japan.
The vesicles (V) in this altered pumice clast are faintly visible albite-filled vesicles mimic pre-existing fibrous textures.
in plane polarised light because they are lined with sericite.
Sample 133815, Cambrian Hercules Pumice Formation, Central
Volcanic Complex, Mount Read Volcanics, Hercules footwall,
western Tasmania.
E. Fibrous feldspar in perlite cores
In plane polarised light, perlitic fractures are conspicuous in
the groundmass of this altered plagioclase-phyric rhyolite.
Sample 147541, Cambrian Kershaw Pumice Formation, Central
Volcanic Complex, Mount Read Volcanics, Murchison Highway,
western Tasmania.
F. In crossed nicols, overlapping arcuate perlitic fractures
are defined by concentrations of sericite and radial fibrous
textures are preserved in the extinction pattern of the albite +
quartz + sericite-altered perlite cores (C).
FIGURE 5.10 | Photomicrographs of relict fibrous textures in vesicles and originally glassy domains in diagenetically altered fades.
1 1 4 I CHAPTER 5
Stage 4: alteration and replacement of mineral phases
Later reactions involve the dissolution and replacement
of earlier diagenetic phases, remaining gel-like glass, and
magmatic minerals such as plagioclase (Torres et al., 1995).
With time and/or increasing temperature, zeolite assemblages
are especially susceptible to replacement because zeolite
crystallisation is controlled by temperature, fluid pressure,
and rock and fluid composition. Transitions from unstable
zeolites (e.g. phillipsite, clinoptilolite and heulandite) to more
stable phases (e.g. mordenite, analcime, laumontite and K-
feldspar) are common. These changes reflect dehydration
reactions (Ratterman and Surdam, 1981; Noh and Boles,
1989; Smith, 1991):
2.75clinoptilolite + 0.75Na
+
+ 3.0H
4
SiO
4
-
Na-clinoptilolite + 2.25mordenite + 0.13K
+
+
0.27Mg
2+
+ 0.08H+ + 4.46H
2
O (R5.5)
Clinoptilolite and mordenite may also react with solution
to form analcime, adularia, quartz and calcite (Iijima, 1974).
mordenite + 2Na
+
+ CO
3
2
"
analcime + 6SiO
2
+ CaCO
3
+ 5H
2
O
clinoptilolite + Ca
2+
+ 2HCO
3
" analcime +
adularia + 5SiO
2
+ CaCO
3
+ CO
2 8H
2
O
(R5.6)
(R5.7)
Increasing diagenesis may favour the formation of calcic
zeolites, particularly in mafic facies, because of increased Ca/Na
activity ratios due to albitisation of calcic plagioclase (Utada,
1991). In addition, chlorite and epidote may crystallise.
Chlorite and epidote have been observed as direct alteration
products of dacitic to basaltic glasses, and clays at relatively
shallow depths (420 m) in modern geothermal regions
(White and Sigvaldason, 1962). The development of epidote
depends on the availability of Fe
3+
and is probably controlled
by the earlier formation of palagonite. Where palagonite is
absent the reaction of basaltic glass to form chlorite releases
Ca, which may be consumed by the precipitation of epidote
(Baragar et al., 1979). Chlorite and epidote are also typical
of low-temperature metamorphism and the growth of these
minerals may bridge the boundary between diagenesis and
metamorphism where glass and diagenetic clays are replaced
by phyllosilicates.
K-feldspar is common in diagenetically altered originally
glassy volcanic facies. Munhaetal. (1980) suggested that below
150C, Na in glass might be exchanged for K in seawater,
resulting in precipitation of K-feldspar and K-smectite.
However, K-feldspar has not been reported as a direct product
of glass alteration, thus some intermediate phases appear to
be required. Iijima and Hay (1968), Surdam and Sheppard
(1978) and Hay and Guldman (1987) recognised that K-
feldspar replaced earlier mordenite, analcime, clinoptilolite
and phillipsite. In contrast, Noh and Boles (1989) proposed
that K-feldspar crystallised from silicic glass via a series of
hydration and dissolution reactions, which included an
intermediate phase of K-rich gel-like glass.
dacitic glass + nH
2
O > perlitic glass + Na
+
+ OH~ (R5.8)
This is probably followed by either reaction R5.9 or
R5.10, which fixes Mg from seawater.
12.5perlitic glass + 3.88K+ + 0.65H
+
+15.4H
2
O -
smectite + 9.5gel-like glass + 4.03Na
+
+ 0.25Ca
2+
10.55H
4
SiO
4
2.79perlitic glass + 0.2Mg
2+
+ 0.2Fe
2+
+ 0.32H+ +
5.27H
2
O 1.27smectite + gel-like glass +
1.0Na
+
+ 0.12Ca
2+
+ 3.61H
4
SiO
4
+
(R5.9)
(R5.10)
The K liberated during zeolite forming reactions (e.g.
reaction R5.4) is then fixed in K-feldspar formation.
gel-like glass + 0.5K+ + 0.2H+ +0.2H
2
O
1.3K-feldspar + 0.1 Na
+
+ 0.1 Ca
2
0.1 Mg
2+
+ 0.1Fe
2+
+ 0.8H
4
SiO
4
+
(R5.ll)
Albite is common as a replacement product of plagioclase
and K-feldspar in rhyolitic to basaltic rocks (Munha et
al., 1980; Boles, 1982; Torres et al., 1995). Albitisation of
plagioclase occurs by dissolution and replacement (Boles,
1982; Shiraki and Iiyama, 1990; Torres et al., 1995). In many
cases, albitisation appears to have progressed preferentially
along plagioclase grain fractures and cleavage traces,
suggesting that fluid films can infiltrate the crystal along
lattice defects and planes of weakness, promoting albitisation
(Boles, 1982). Microlites in the glassy groundmass of slowly
cooled lavas and sills may serve as nuclei for the crystallisation
of albite (Dimroth and Lichtblau, 1979). The replacement of
plagioclase and K-feldspar by albite could reflect the exchange
of K in the rock with Na in seawater at greater depths and
temperatures (105-120C) (Iijima and Utada, 1972; Merino,
1975; Munha et al., 1980; Boles, 1982). The Na and Si
required for albite crystallisation are supplied by diffusion
from seawater and earlier diagenetic reactions (Boles and
Coombs, 1977; Boles, 1982).
2SiO
2
+ 0.5H
2
O + H
+
+ Na
+
+ Ca-plagioclase
albite + 0.5Al,Si,O
s
(OH)
4
+ Ca
2+
(R5.12)
Albite is typically riddled with minute inclusions of clays,
sericite and calcite, which form as by-products and consume
Ca and Al released during this reaction.
SE AF L OOR-AN D BURIAL-RELATED ALTERATION I 1 1 5
5.4 | REGIONAL METAMORPHISM
(ZEOLITE TO AMPHIBOLITE
FACIES)
Transition from diagenesis to regional
metamorphism
Diagenesis progresses gradually to regional metamorphism
with increasing temperature, pressure and, commonly, depth
of burial. Many sedimentologists consider that the boundary
between diagenesis and metamorphism occurs when a rock has
less than 5% interconnected pore space, whereas metamorphic
petrologists tend to define it by mineral assemblages that are
not stable in sedimentary environments (e.g. Coombs, 1954;
Blatt et al., 1972; Winkler, 1979; Turner, 1980; Morrow
and Mcllreath, 1990; Bevins and Robinson, 1992). These
definitions do not necessarily coincide. Diagenesis and
low-grade metamorphism can both have temperatures and
pressures in the 200-300C and less than 1 kbar range. In
fact, no clear distinction exists between the processes, pressure
and temperature conditions, fabrics (textures) and mineral
assemblages of diagenesis and low-grade metamorphism.
Diagenesis involves hydration of glass, compaction,
dissolution, cementation and minor recrystallisation: meta-
somatic processes involving minor chemical exchange
between the host facies and trapped fluid at low temperatures
(up to 250C). In contrast, metamorphism involves mainly
isochemical processes with substantial recrystallisation,
but chemical changes that are limited to dehydration and
decarbonation (Fyfe et al., 1958). During progressive burial
there is a stage when metasomatic reactions occur but the
temperatures are generally considered too high for diagenesis;
this transitional stage is referred to as burial metamorphism
(Coombs, 1954).
The transition from diagenesis to burial metamorphism
has been studied in andesitic to rhyolitic rocks in New Zealand
(e.g. Coombs, 1954; Coombs et al., 1959; Boles, 1974; Boles
and Coombs, 1975, 1977), Chile (e.g. Levi, 1970), the USA
(e.g. Dickinson, 1962; Sheppard and Gude, 1973) and Japan
(e.g. Utada, 1970; Iijima and Utada, 1972; Iijima, 1978), and
in basaltic rocks in Canada (e.g. Surdam, 1973; Kuniyoshi
and Liou, 1976), Australia (e.g. Smith, 1969; Hellman et
al., 1977; Smith et al., 1982) and Iceland (e.g. Viereck et al.,
1982).
Burial metamorphism
Burial metamorphism was defined by Coombs (1954) to cover
progressive mineral changes that can be directly correlated
with increases in temperature and burial depth in thick
sedimentary or volcanic successions. It is a form of regional
metamorphism that affects thick sedimentary or volcanic
successions in subsiding basins, where the basal parts attain
low-grade metamorphic conditions without the deformation
or folding typical of regional metamorphism.
Burial metamorphism, like diagenesis, rarely attains
equilibrium mineral assemblages, and penetrative deformation
fabrics are absent. Alteration minerals common to burial
metamorphism in submarine volcanic successions are:
zeolites (heulandite, stilbite, laumontite, analcime), prehnite,
pumpellyite, epidote, albite, K-feldspar, phyllosilicate minerals
(smectites, chlorite, celadonite, sericite), calcite, siderite,
quartz, apatite, sphene, pyrite and Fe-oxides (Surdam, 1973;
Boles and Coombs, 1977).
Burial metamorphic facies
A metamorphic facies is defined by a group of metamorphic
mineral assemblages occurring in spatially associated rock
types of diverse chemical composition, which are interpreted
to have formed during restricted temperature and pressure
conditions (Fig. 5.11).
Low-grade facies typical of burial metamorphism in
volcanic successions are characterised by hydrous minerals and
carbonates, whereas high-grades facies are typically coarser
grained and contain anhydrous and CO
2
-poor minerals.
The low-grade facies are (Table 5.3): zeolite, prehnite +
pumpellyite, pumpellyite + actinolite, lawsonite + albite +
chlorite, blueschist, and greenschist facies (Turner, 1980).
Although burial metamorphic facies are widespread,
they are commonly heterogenous with patchy and domainal
textures (e.g. epidote metadomains of Smith, 1968, 1974,
1977; Smith and Smith, 1976). This domainal style of
alteration may reflect mobilisation and local redistribution
of elements on a scale of centimetres to metres (e.g. the
loss of Fe
total
, Mg, Na and K and gain of Ca from epidote
metadomains balance losses and gains from the enclosing
albite domains, Smith, 1977), rather than the significant
addition of elements.
Burial metamorphic zones
Burial metamorphism is a progressive process that produces
a sequence of regionally extensive metamorphic zones (e.g.
Figs 3.13 and 5.12). Metamorphic zones are mappable groups
of rocks that have similar metamorphic grade. Adjacent
metamorphic zones, like altered zones, are separated by lines
of equal grade (isograds), which are delineated by the first
appearance of an index mineral or minerals within the same
rock type or composition.
FIGURE 5.11 | Pressure and temperature diagram showing the fields for
regional metamorphic facies (after Turner, 1980; Bevins and Robinson, 1992).
Boundaries between the fields are gradational. Abbreviations used are PA =
pumpellyite + actinolite and PP = prehnite + pumpellyite.
1 1 6 I CHAPTER 5
Table 5.3 | Common burial and regional metamorphic facies and mineral assemblages for submarine volcanic successions (from Coombs, 1954; Humphris
and Thompson, 1978; Turner, 1980; Sivell, 1984; Yardley, 1989).
Zeolite Zeolites (laumonite, analcime, heulandite, stilbite, natrolite, mesolite, wairakite) + mixed-layer clays +
quartz + calcite muscovite
Prehnite + pumpellyite Prehnite + pumpellyite chlorite + albite + quartz epidote calcite sphene rare garnet
Pumpellyite + actinolite Pumpellyite + actinolite + albite + chlorite + sphene + quartz + muscovite + calcite lawsonite
Lawsonite + albite + chlorite Lawsonite + albite + chlorite + quartz pumpellyite epidote actinolite sphene + muscovite calcite
Blueschist Glaucophanic amphibole + albite + actinolite/phengite + quartz epidote chlorite sphene
pumpellyite stilpnomelane calcite
Greenschist Actinolite + epidote + albite chlorite calcite tremolite talc + quartz sphene magnetite biotite
Amphibolite Hornblende + plagioclase epidote garnet biotite quartz + muscovite + calcite sphene
magnetite
Drilling in modern geothermal regions has revealed
patterns of low-grade metamorphic zones that can be related
directly to temperature and fluid composition at shallow
depths (e.g. Fig. 5.13: Coombs et al., 1959; White and
Sigvaldason, 1962; Vierecketal., 1982). It is generally assumed
that higher-grade rocks formerly had mineral assemblages
typical of lower grade zones, which were progressively altered
as metamorphism proceeded. Near isograds, index minerals
from the lower zones locally overgrow lower grade minerals.
For example, in basaltic lavas and tuffs in east Greenland near
the boundary between two zeolite facies burial metamorphic
zones, the analcime and mesolite + scolecite zones, vesicles
are lined with analcime and filled with mesolite (Neuhoff et
al., 1997).
The generalised sequence of burial metamorphic facies
with increasing depth is (e.g. Fig. 3.13): zeolite, prehnite +
pumpellyite, pumpellyite + actinolite, lawsonite + albite +
chlorite, blueschist, and greenschist facies (Coombs, 1954;
Coombs et al., 1959; Turner, 1980). Departures from this
pattern are common.
Zeolite facies
Coombs (1954) and Coombs et al. (1959) proposed that
regionally extensive zeolite facies in Triassic volcaniclastic rocks
of New Zealand bridged the transition between diagenesis and
conventional metamorphism. The zeolite facies embraces co-
existing assemblages of Ca + Al- and Na + Al-rich zeolites and
quartz (Coombs et al., 1959). No single mineral assemblage
is diagnostic of the zeolite facies as mineral assemblages are
sensitive to primary rock composition, fluid composition,
burial history and geothermal gradient. Typically the zeolite
facies contains heulandite, laumontite, analcime, quartz,
albite and smectites + prehnite and pumpellyite (Turner,
1980; Vierecketal., 1982).
Fyfe et al. (1958) and Coombs (1954), in studies of
Triassic submarine volcano-sedimentary rocks of Tarinagatura
Hills (New Zealand), divided the zeolite facies into three
mineral assemblages that correlate with increasing depth: (1)
heulandite + analcime + quartz (montmorillonite + celadonite
+ sphene), (2) laumontite + albite + quartz chlorite, and
(3) quartz + albite + adularia. However, at Hokonui Hills,
50 km east of Tarinagatura Hills, Boles and Coombs (1975)
found no correlation between mineral assemblage and depth
of the zeolite facies rocks. In contrast, Neuhoff et al. (1997)
divided the zeolite facies in the Tertiary flood basalts of east
Greenland into five diagenetic and burial metamorphic zones
based on the index minerals: (1) chabazite + thomsonite, (2)
analcime, (3) mesolite + scolecite, (4) heulandite + stilbite
and (5) laumonite (Fig. 5.12).
Genesis
The development of the burial metamorphic zones is complex.
In many cases, it is progressive through a number of stages
that may overlap with diagenesis. However, the transition
to burial metamorphism generally involves dehydration and
decarbonation accompanied by the release of silica, such as in
FIGURE 5.12 | Cross-section in the Borggraven region, east Greenland, showing the regional extent and vertical distribution of burial-
metamorphic zeolite zones (after Neuhoff et al., 1997). The thin discontinuous lines represent the dips of selected lavas in this section.
SEAFLOOR- AND BURIAL-RELATED ALTERATION | 1 1 7
F IG U R E 5 . 1 3 | Schematic cross-section showing the low-temperature altered
zones and average temperature measurements for the boundaries in the
Wairakei geothermal field, Taupo V olcanic Zone, N ew Zealand (modified from
Coombs et al . , 1 95 9, after Steiner, 1 95 3; Banwel l et al . , 1 95 7 ).
reactions R5.13 and R5.15, below (Coombs, 1954; Boles and
Coombs, 1977).
Rocks in the Tarinagatura Hills record two stages of
zeolite facies alteration: an upper heulandite + analcime
diagenetic stage and a lower laumonite + albite + quartz burial
metamorphic stage (Coombs, 1954). In the upper part of the
succession, volcanic glass has been replaced by heulandite
and, less commonly, analcime (Fig. 5.14). These zeolites co-
exist with newly crystallised quartz and fine-grained smectite.
In the lower part, analcime has been replaced by albite and
heulandite by laumonite + quartz. Smectite, chlorite, sericite
and mixed-layer minerals occur and prehnite and pumpellyite
appear as accessory minerals. These stages can be summarised
by the reactions:
heulandite > laumontite + quartz + 2H
2
O (R5.13)
analcime + quartz > albite + H
2
O (R5.14)
laumontite + calcite * prehnite + quartz +
H
2
O + CO
2
(R5.15)
Some successions lack textural evidence for shallow
diagenetic and zeolite zones as precursors to higher-grade
mineral assemblages, suggesting that metamorphism to higher
grades is not always progressive (e.g. Neuhoff et al., 1997) or
that early textures are destroyed.
FIGURE 5.14 | Down-hole mineral distributions at Tarinagatura Hills, New
Zealand (after Coombs, 1954).
1 1 8 I CHAPTER 5
5.5 | DIAGENETIC ALTERATION IN THE
HOKUROKU BASIN
The Middle Miocene Hokuroku Basin, part of the Green
Tuff Belt in northern Honshu, Japan, is the most frequently
cited example of diagenetic zones in a submarine volcanic
succession that hosts VHMS deposits. Iijima (1974, 1978),
Iijima and Utada (1971) and Utada (1991) described four
flat-lying, vertically stacked, zeolite-dominated altered zones
that grade into clay-rich hydrothermal zones proximal to the
Kuroko VHMS deposits (Fig. 5.6A and B).
Geological setting
The Hokuroku Basin is a 30 x 30 km submarine basin
containing a 3 to 6 km thick bimodal volcanic succession of
calc-alkaline rhyolites and tholeiitic basalts with some locally
abundant andesites (Figs 5.6 and 5.15: Dudas et al., 1983;
Urabe, 1987). Here, the Nishikurosawa and Onnagawa
Formations dominate the stratigraphy.
The lower Nishikurosawa or Hotakizawa Formation is up
to 650 m thick and includes intercalated basaltic lavas and
breccias, rhyolitic lavas, and laminated mudstone (Tanimura
et al., 1983). The upper Nishikurosawa Formation is a
thick (<400 m) succession of rhyolitic lavas, domes and
interbedded pumice-rich facies and mudstone (Fig. 5.16:
Ishikawa, 1983; Urabe, 1987; Yamagishi, 1987). The upper
Nishikurosawa Formation hosts the Kuroko VHMS deposits
and is conformably overlain by the Onnagawa Formation
(Nakajima, 1988).
The Onnagawa Formation comprises a sequence of
pumice-rich breccia, sandstone, siltstone and black mudstone
with abundant felsic synvolcanic intrusions and local basaltic
lavas (Fig. 5.16: Ohmoto and Takahashi, 1983; Tanimura
et al., 1983; Nakajima, 1988). The lower pumice breccia is
extensive and has been correlated across the Hokuroku Basin
(Urabe, 1987).
The volcanic succession is relatively undeformed but
has undergone regional diagenesis and local hydrothermal
alteration and mineralisation (Utada, 1970; Tanimura et al.,
1983). Generally the stratigraphy has a gentle dip with open,
N-S-trending folds (Tanimura et al., 1983).
Noquchi depression
LEGEND
Funakawa Formation
Onnagawa Formation
Nishikurosawa Formation
Daijima Formation
Rhyolite
Andesite
Dolerite
Quartz diorite
Other rocks
Drill holes in this study
Major mine
Township
Major fault
Railway
FIGURE 5.15 | Geology of the Hokuroku Basin showing the major lithostratigraphic units and inset the distribution of the Green Tuff Belt in Japan (after Sato, 1974;
Tanimura et al., 1983). The trace of Ijima's (1974) cross-section A (in Fig. 5.6 ) is represented by solid line AA', B by the dashed line BB' and our cross-section (Fig.
5.18) by the dotted line BC.
SE AF L OOR-AN D BURIAL-RELATED ALTERATION I 1 1 9
FIGURE 5.16 | Graphic log of drill core OH-8 in the Odate Basin, western Hokuroku Basin (Japan) showing lithology, bedforms, and diagenetic and hydrothermal
zones in pumice-rich facies of the Upper Nishikurosawa and Onnagawa Formations.
Alteration facies and zones
The regional diagenetic zones are, from top to base (Figs
5.6 and 5.17): (I) partially altered zone, (II) clinoptilolite +
mordenite zone (e.g. data sheets HK1 and 2), (III) analcime
zone (e.g. data sheets HK3 and 4), and (IV) laumontite +
albitezone (Iijima, 1974, 1978; Hay, 1978; Iijima, 1978).
The partially altered zone is distributed beneath the
Quaternary gravels and overlies the clinoptilolite + mordenite
zone in the eastern Odate Basin, western Hokuroku Basin
(Fig. 5. 6A). This zone has a maximum thickness of 60 m
and is characterised by well preserved volcanic textures,
the absence of zeolites and the presence of unaltered glassy
pumice clasts and plagioclase crystals (Iijima, 1974). Rocks in
this zone are pale grey. Glass shards and parts of pumice clasts
have been altered to montmorillonite and vesicles are empty
or have been partly filled with low-cristobalite.
The clinoptilolite + mordenite zone is 160250 m thick
and widely distributed in the shallower part of the Odate Basin
in the upper Onnagawa Formation. Regionally it overlies
the analcime zone, but in the east of the Odate Basin it is
repeated below the analcime + heulandite zone (Figs 5.6B and
5.17). In the upper part of this zone, some glass shards and
pumice clasts are unaltered (e.g. data sheet HK1). Typically
the surfaces of shards, pumice clasts and vesicles have been
coated in a thin film of smectite or low-cristobalite. Vesicles
were filled sequentially with mordenite and clinoptilolite.
Originally glassy shards and clasts have been altered to smectite
(montmorillonite, saponite or mixed-layer illite/smectite)
mordenite. In the lower part of this zone, dark green saponite
1 2 0 I CHAPTER 5
FIGURE 5.17 | Schematic cross-section of the western
Hokuroku Basin (Japan) showing lithology and altered
zones. The locations of the six data sheets are marked
on the section.
or mixed-layer smectite/chlorite and pale green mordenite
domains are common in the pumice-rich rocks (e.g. data
sheet HK2). Plagioclase crystals are typically unaltered.
The analcime zone is at least 7 km wide and 150-200 m
thick, and occurs in the Upper Nishikurosawa and lower
Onnagawa Formations (below the Ml mudstone). It is
approximately equivalent to the ore position, grading into
the hydrothermal montmorillonite zone and overlying the
sericite + chlorite zone (e.g. data sheets HK5 and 6) associated
with the Kuroko VHMS deposits (Figs 5.6 and 5.17). In the
east of the Odate Basin, an analcime + calcite zone occurs
within the clinoptilolite + mordenite zone (Fig. 5.6A: Iijima,
1974). However, it is different from the regional analcime
zone, containing disseminated pyrite and calcite concretions
and veinlets, and is considered to result from hydrothermal
alteration related to Kuroko mineralisation (Yoshida and
Utada, 1968; Iijima, 1974).
The regional analcime zone is characterised by analcime,
which has completely replaced some shards, and rounded
crystals of analcime that have replaced clinoptilolite- and
mordenite-altered shards and pumice clasts (e.g. data sheets
HK 3 and 4). The internal structure of analcime-altered
uncompacted tube pumice clasts is not as well preserved
as those in the clinoptilolite + mordenite zone. Plagioclase
crystals are unaltered or have been partly analcime and calcite
altered. Dark green saponite, saponite + chlorite and chlorite
fiamme are common in this zone. They are interpreted as
altered and compacted pumice clasts (Gifkins et al., in press).
Analcime-filled solution seams are common in the lower part
of this zone.
The albite + laumontite zone occurs beneath the sericite
+ chlorite zone at depth in the Odate Basin (Fig. 5.6). It is
characterised by albite + laumontite calcite chlorite +
sericite (Iijima and Utada, 1971). Laumontite has replaced
plagioclase crystals, originally glassy shards and pumice clasts
and filled pore space (Iijima, 1974). Albite has replaced
plagioclase crystals.
Genesis of altered zones
The four zeolite zones in the Hokuroku Basin are interpreted
to have formed during submarine diagenesis of the mainly
glassy felsic volcanic succession (Iijima and Utada, 1971;
Iijima, 1974, 1978; Utada, 1991). The alteration pattern
is interpreted to have formed beneath the seafloor, while
sedimentation continued and the rocks were progressively
buried. Ptygmatic folds in near vertical calcite veinlets in
SEAFLOOR-AND BURIAL-RELATED ALTERATION | 1 2 1
Diagenetic zones
ZONE!
I | Partially altered zone lacking zeolites
ZONE II
I I Clinoptilolite + mordenite zone
ZONE III
Analcime + calcite zone
Hydrothermal zones
^B Montmorillonite and transitional zones
Bi Sericite + chlorite zone with plagioclase
I | Sericite + chlorite zone lacking
plagioclase
f.-' .'] Gravel
I ? I Dacitic lava
I T l Basaltic lava/sill
Massive sulfide ore
i i Felsic volcanic facies
FIGURE 5.18 | Model for the development and
deformation of the altered zones in the Hokuroku
Basin from the Middle Miocene to Recent
(after lijima, 1974). (A) Late Nishikurosawa
stage (Middle Miocene). (B) Onnagawa stage
with deposition of the M1 mudstone. (C) Late
Funakawa stage (Late Miocene). (D) Recent.
Zone III are considered to result from compaction that took
place when the volcanic facies were most deeply buried.
Thus the diagenetic zones probably formed before the end
of deposition of the upper Miocene Funakawa Formation,
which overlies the Onnagawa Formation (lijima, 1974).
The development of the diagenetic zones can be described in
four stages (Fig. 5.18): late Nishikurosawa stage, Onnagawa
stage, late Funakawa stage and recent (lijima, 1974). During
these stages, mineralogical changes progressed with depth as
successive reactions between volcanic glass and interstitial
modified seawater occurred, originally forming zeolites and
then albite as the temperature and pressure increased (Utada,
1991). These mineralogical changes were accompanied by
textural and compositional changes.
During the late Nishikurosawa stage (Fig. 5.18A) in the
middle Miocene, felsic glass was hydrated and began to alter
to montmorillonite and low-cristobalite (lijima, 1974). This
resulted in a shallow, regionally extensive partly altered zone.
Submarine hydrothermal activity associated with VHMS
mineralisation also commenced (lijima, 1974).
The clinoptilolite + mordenite zone formed during the
late Nishikurosawa and early Onnagawa stage (Fig. 5.18A
and B), when partly altered facies were buried to a depth
of approximately 100 m (lijima, 1974). This increased the
alkalinity of the fluid resulting in the dissolution of felsic glass
and the precipitation of alkali clinoptilolite and mordenite
in vesicles, interstitial voids and dissolution cavities.
Simultaneous reactions altered glassy clasts and tube pumice
walls to saponite. Reaction rates were possibly accelerated as a
result of hydrothermal fluids circulating within the succession
increasing the geothermal gradient and concentrations of Na
and K in the fluid. By the end of the Onnagawa stage the
hydrothermal sericite + chlorite and montmorillonite zones
associated with the ore deposits had formed (lijima, 1974).
During the late Funakawa stage (Fig. 5.18C), clinoptilolite
and mordenite reacted with solution to form analcime,
adularia, quartz and calcite (Ogihara, 1996). As a result, the
analcime + heulandite zone was superimposed on the deeper
part of the clinoptilolite + mordenite zone (lijima, 1974). The
Funakawa stage corresponds to the time of deepest burial and
compaction of altered pumice clasts to form fiamme (lijima,
1974). The diagenetic analcime + heulandite zone formed
contemporaneously with hydrothermal zones surrounding
the Kuroko VHMS deposits (lijima, 1978; Ohmoto, 1978).
Since the Funakawa stage, the altered zones have been
mildly deformed and eroded (Fig. 5.18D: lijima, 1974).
1 2 2 | CHAPTER 5
Subtle, patchy mordenite + smectite-chlorite alteration facies HK1
Sample No.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
J6 -294
subtle, patchy mordenite + smectite-
chlorite
clinoptilolite + mordenite zone
Yoneshiro River
Onnagawa Formation
Green Tuff Belt
pumice breccia
plagioclase + quartz
tube pumice clasts, bubble-wall shards,
crystal fragments, non-vesicular volcanic
clasts
rhyolite
graded bed
syneruptive, mass-flow emplaced pumice
breccia
partly glassy, mordenite + saponite +
montmorillonite + smectite-chlorite +
K-feldspar + pyrite
saponite films in vesicles, mordenite
saponite filled vesicles and pore space,
smectite-chlorite fiamme, disseminated
pyrite
patchy
excellent
subtle
early
Weak, pervasive mordenite + smectite alteration facies
SE AFL OOR-AN D BURIAL-RELATED ALTERATION | 1 2 3
HK2
Sample No.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
OH8-36 9
weak, pervasive mordenite + smectite
clinoptilolite + mordenite zone
Odate city
Onnagawa Formation
Green Tuff Belt
pumice + lithic breccia
plagioclase + quartz
tube pumice clasts, bubble-wall shards,
crystal fragments, non-vesicular volcanic
clasts
rhyolite
graded bed
syneruptive, mass-flow emplaced pumice
breccia
mordenite + smectite-chlorite + K-feldspar
+ calcite + pyrite + glauconite
smectite films in vesicles, mordenite
smectite filled vesicles and pore space,
mordenite-altered glass shards and
vesicle walls, smectite-chlorite fiamme,
disseminated pyrite, microcrystalline lithic
clasts
pervasive
excellent
weak
early
diagenetic
1 2 4 | CHAPTER 5
Subtle, pervasie smectite-chlorite + mordenite + analcime alteration facies HK3
Sample No.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
OH8-511
subtle, pervasive smectite-chlorite +
mordenite + analcime
analcime zone
Odate city
Onnagawa Formation
Green Tuff Belt
pumice breccia
plagioclase + quartz
tube pumice clasts, crystal fragments
rhyolite
massive
syneruptive, mass-flow emplaced pumice
breccia
smectite-chlorite + mordenite + analcime +
sericite + pyrite
analcime solution seams, smectite-chlorite
fiamme, mordenite filled vesicles, analcime
replacing mordenite + smectite-altered
pumice clasts
pervasive
good
subtle
early
diagenetic
Weak, pervasive analcime + mordenite alteration facies
SEAFLOOR- AND BURIAL-RELATED ALTERATION | 1 2 5
HK4
Sample No.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
OH8-537
weak, pervasive analcime + mordenite
analcime zone
Odate city
Onnagawa Formation
Green Tuff Belt
pumice breccia
plagioclase
tube pumice clasts, bubble-wall shards,
crystal fragments
rhyolite
graded bed
syneruptive, mass-flow emplaced pumice
breccia
analcime + mordenite + clinoptilolite +
smectite-chlorite + pyrite + sericite
mordenite and clinoptilolite filled
vesicles, mordenite and analcime altered
vesicle walls, analcime overgrowths on
plagioclase crystal fragments
pervasive
moderate
weak
early
diagenetic
1 2 6 | CHAPTER 5
Strong, pervasive quartz + sericite alteration fades HK5
Sample No.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
OH8-794
strong, pervasive quartz + sericite
sericite + chlorite zone
Odate city
Nishikurosawa Formation
Green Tuff Belt
pumice + lithic breccia
plagioclase
volcanic clasts
rhyolite
graded bed
syneruptive, mass-flow emplaced pumice
breccia
quartz + K-feldspar + sericite + chlorite+
pyrite
pseudomorphs after plagioclase crystals
and clasts in pervasive crystalline matrix,
quartz-filled dissolution vugs
pervasive
poor
strong
early
hydrothermal
moderate, pervasive sericite + chlorite alteration facies
SE AFL OOR-AN D BURIAL-RELATED ALTERATION | 1 2 7
HK6
Sample No.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
HO20-485
moderate, pervasive sericite + chlorite
sericite + chlorite zone
near Fukazawa deposit
Nishikurosawa Formation
Green Tuff Belt
pumice + lithic breccia
nil
clasts?
rhyolite
graded bed
synemptive, mass-flow emplaced pumice
breccia
chlorite + sericite + pyrite + montmorillonite
dissolution vugs after crystals, sericite +
chlorite fiamme, disseminated pyrite
pervasive
poor
moderate
early
hydrothermal
1 2 8 | CHAPTER 5
5.6 | DIAGENETIC ALTERATION IN THE
MOUNT READ VOLCANICS
Gifkins (2001) recognised two regionally developed
Cambrian diagenetic zones (albite zone and epidote zone)
within the northern Central Volcanic Complex in the
Mount Read Volcanics. Although the rocks currently have
mineral assemblages consistent with greenschist facies
metamorphism, diagenetic alteration facies were identified
based on combinations of mineral assemblages, overprinting
relationships, textures, distribution, alteration intensity and
whole-rock geochemistry. The original diagenetic mineral
assemblages were inferred from local relict textures and by
comparison with younger diagenetically altered volcanic
successions (Gifkins and Allen, 2001).
Locally, the diagenetic alteration facies merge with
hydrothermal alteration facies in the margins of the Rosebery
and Hercules VHMS systems (Allen, 1997). Diagenetic and
hydrothermal alteration facies are interpreted to have had
similar timing (e.g. Fig. 3.20), and the hydrothermal system
may have contributed heat and fluid to intensify the diagenetic
system (Gifkins, 2001).
Geological setting
For a detailed description of the geology of the Mount Read
Volcanics refer to Section 1.5. The northern Central Volcanic
Complex is exposed in an approximately 30 by 6 km area
located north and west of the Henty fault, and east of the
Rosebery fault (Fig. 1.5). It includes three compositionally
and texturally different formations (Fig. 5.19): the Sterling
Valley Volcanics, the Mount Black Formation, and the
Kershaw (or Hercules) Pumice Formation (Gifkins, 2001).
The Sterling Valley Volcanics (>1.5 km thick) are composed
of dacitic to basaltic lavas and sills, and polymictic mafic
volcaniclastic facies interpreted as resedimented syneruptive
hyaloclastite, autobreccia, pillow lava and scoria. The Mount
Black Formation is a laterally extensive (>20 km), thick
succession (>1.6 km) of mainly feldspar-phyric massive, flow-
banded and autobrecciated lavas, domes, cryptodomes and
synvolcanic sills (e.g. data sheets CVC1 CVC5 and CVC6).
The Kershaw Pumice Formation, which conformably
overlies the Mount Black Formation, is a laterally extensive
(>16 km), relatively thick (>800 m) succession dominated
by non-welded pumice breccia (e.g. data sheets CVC3 and
CVC4), pumice-rich sandstone and shard-rich siltstone, with
lesser proportions of pumice-lithic clast-rich breccia and
sandstone, and massive, flow-banded and brecciated rhyolitic
and dacitic lavas and intrusions (e.g. data sheet CVC2). The
upper part of the Kershaw Pumice Formation and the base of
the overlying White Spur Formation host the Rosebery and
Hercules VHMS deposits.
Abundant spherulites, lithophysae, micropoikilitic texture
and relict perlite indicate that volcanic rocks in the northern
Central Volcanic Complex were initially partly crystalline and
partly glassy (Gifkins and Allen, 2001).
Alteration facies and zonation
Regionally distributed diagenetic albite and epidote zones
formed before, or were synchronous with, stylolitic S,
compaction foliation. The alteration intensity is generally
weak with volcanic textures and albite-altered plagioclase
crystals preserved. Locally the distribution and intensity of the
diagenetic alteration facies is patchy, reflecting the complexity
of the original volcanic facies (Fig. 5.20).
The albite zone is characterised by pervasive albite + quartz
+ sericite (e.g. data sheets CVC2 and CVC5), domainal albite
+ quartz + sericite with sericite + hematite chlorite (e.g. data
sheet CVC3) and pervasive sericite (e.g. data sheet CVC4)
alteration facies. It is thick (>2 km) and encompasses the
Kershaw Pumice Formation and most of the Mount Black
Formation (Fig. 5.21). The albite + quartz + sericite-rich facies
are associated with minor increases in SiO
2
, CaO, Na
2
O and
total mass, and decreases in K
2
O and A1
2
O
3
consistent with
seafloor albitisation (cf. Boles and Coombs, 1977; Boles,
1982). The sericite + hematite + chlorite alteration facies is
associated with minor increases in K
2
O and A1
2
O
3
consistent
with the conversion of silicic glass to clay minerals (cf. Noh
and Boles, 1989; Passaglia et al., 1995). The abundance of
hematite may reflect the oxidation of Fe
3+
during alteration of
glass to clays (e.g. Klein and Lee, 1984).
The epidote zone is characterised by pervasive albite +
quartz + sericite, pervasive chlorite + sericite, pervasive chlorite
+ epidote and domainal chlorite + epidote with albite + quartz
+ sericite (e.g. data sheet CVC6) alteration facies. The epidote
zone is less extensive than the albite zone and is restricted to the
Mount Black Formation and Sterling Valley Volcanics at the
stratigraphic base of the northern Central Volcanic Complex,
adjacent to the Henty fault (Fig. 5.19). In the epidote zone,
chlorite + sericite and chlorite + epidote altered felsic rocks
have gained MgO, consistent with the formation of smectite,
chlorite and other Mg-silicates during diagenesis (cf. Hajash
and Chandler, 1981; Shiraki and Iiyama, 1990).
Genesis of alteration facies
The epidote zone occurs in the core of the regional anticline
in the Sterling Valley, suggesting that it is associated with the
deepest stratigraphic level in the northern Central Volcanic
Complex: the lower Mount Black Formation and Sterling
Valley Volcanics (Gifkins, 2001). The change from the albite
zone to the epidote zone with stratigraphic depth is consistent
with diagenetic alteration zonation (cf. Iijima, 1974, 1978).
Thick (>1 km) diagenetic zones with high-temperature
mineral assemblages (albite + quartz + sericite and chlorite
+ epidote) suggest that they developed in response to a high-
grade diagenetic alteration system that involved an elevated
geothermal gradient (cf. Utada, 1991).
Albite + quartz + sericite, sericite + hematite chlorite,
and sericite alteration facies are the metamorphosed
equivalents of diagenetic alteration facies that coated original
surfaces, filled primary porosity and replaced glass in the
northern Central Volcanic Complex prior to or synchronous
with diagenetic compaction. Thin films of sericite, carbonate
and hematite replaced clays that had coated original glassy
surfaces at the onset of diagenesis. Albite + quartz + sericite,
SE AFLOOR-AN D BURIAL-RELATED ALTERATION | 1 2 9
FIGURE 5.19 | Geology of the northern Mount Read Volcanics in western Tasmania, showing the major lithostratigraphic units and altered
zones in the northern Central Volcanic Complex (after Gifkins, 2001). Locations of the six data sheets are marked on the map.
FIGURE 5.21 | Schematic cross-section of the northern Central Volcanic
Complex stratigraphy and altered zones, western Tasmania (after Gifkins, 2001).
FIGURE 5.20 | Detailed cross-section in the Rosebery hanging wall (western Tasmania) showing the complex distribution of volcanic and alteration fades (after
Gifkins and Allen, 2001).
SEAFLOOR- AND BURIAL-RELATED ALTERATION | 1 3 1
St age 1: Onset of regi onal synvolcanic di agenesi s
Thin films of sericite, hematite and calcite coat original surfaces, including
vesicle walls, plagioclase crystals, shards and fractures. These films are
the metamorphic equivalents of low-temperature smectite, palagonite and
calcite rim cements. This early stage probably involved interaction with
seawater trapped in the volcanic succession. Modified seawater may have
been expelled from the succession in response to overburden pressure,
and migrated towards the seafloor as diffuse unfocused flow.
St age 2: Di agenesi s and synchr onous hydrothermal
alt erat i on and mi nerali sat i on
Zeolite or clay mineral cements began to fill primary pore spaces, vesicles
and perlitic fractures. Subsequently, these were extensively replaced
by K-feldspar or albite and chlorite. Zeolitisation probably occurred at
temperatures between 40 and 100C. Locally, hydrothermal fluids altered
the succession. Hydrothermal fluid flow was unfocussed and in places
ponded beneath the coherent facies of sills and lavas. The Rosebery and
Hercules VHMS deposits and their altered halos are interpreted to have
formed during this stage at temperatures greater than 300C.
St age 3: Cont i nui ng di agenet i c alt erat i on and
compact i on synchr onous wi t h deposi t i on of t he Whi t e
Spur Formation
Dissolution and alteration of glass to clays, sericite and chlorite occurred
synchronous with compaction. Replacement of earlier zeolites by K-
feldspar occurred below 150C, albitisation of plagioclase phenocrysts and
albite replacement of K-feldspar occurred at temperatures between 100
and 190C. Large volumes of fluid were probably displaced as a result of
compaction under the weight of the accumulating White Spur Formation.
Rapid and variable sedimentation rates may have over-pressured the
pore fluid, promoting lateral fluid flow along permeable layers. Weak
hanging wall alteration developed during continued hydrothermal alteration
associated with the formation of the Rosebery deposit.
Stage 4: Transition between di agenesi s and regi onal
met amorphi sm
More stable, higher-temperature mineral assemblages replaced remaining
glass, phenocrysts and early alteration minerals. Chlorite + epidote
alteration facies developed at depth in both mafic and felsic volcanic
facies: probably at high (>200C) temperatures.
St age 5: Devonian met amor phi sm and def or mat i on
Greenschist facies mineral assemblages and tectonic fabrics overprinted
diagenetic and hydrothermal alteration facies. Deformation modified pre-
existing volcanic and alteration textures and produced folds, faults and
shear zones. The distribution of syn-S
2
alteration facies suggests that
metamorphic fluid migration was restricted to regional structures such
as faults and shear zones. Mineral assemblages in intermediate and
mafic rocks in the Mount Read Volcanics indicate that the peak regional
metamorphic temperature was between 370 and 450C.
FIGURE 5.22 | Model for the post-depositional evolution of the northern Central Volcanic Complex, western Tasmania (after Gifkins, 2001). Schematic cross-
sections are not to scale.
1 3 2 | CHAPTER 5
chlorite + sericite and sericite + hematite + chlorite replaced
zeolites and clays that filled pore space and altered glass, prior
to and synchronous with diagenetic compaction. In pumice-
rich facies, a bedding-parallel stylolitic foliation reflects
the dissolution of glass during compaction and fiamme are
interpreted as diagenetically altered and flattened pumice
clasts (Gifkins et al., in press). Diagenetic alteration involved
significant mineralogical and textural changes but only minor
changes in composition consistent with the interaction of
rhyolitic and basaltic glass with seawater during burial.
The chlorite + epidote alteration mineral assemblage may
be transitional between diagenesis and burial metamorphism.
It developed after lithification and compaction but pre-dated
regional deformation associated with peak metamorphism.
The chlorite + epidote facies replaced earlier clay or chlorite
+ sericite-rich facies and filled any remaining pore space.
Negligible absolute and total mass changes associated with
chlorite + epidote alteration suggest that it grew in response
to increasing temperature with increasing depth of burial late
in the diagenetic history (Gifkins, 2001).
Mineral assemblages in these diagenetic zones reflect the
reaction of glass with interstitial fluid at elevated temperatures.
The albite zone probably formed at temperatures between
100 and 190C (cf. Iijima and Utada, 1971; Thompson,
1971; Merino, 1975; Munha et al., 1980; Boles, 1982).
The epidote zone is characterised by chlorite + epidote,
chlorite + sericite and albite + quartz + sericite indicating
formation at temperatures of at least 200C (cf. Seki, 1972;
Kristmannsdottir, 1976).
The regional diagenetic alteration and metamorphism of
the northern Central Volcanic Complex can be described in
five successive stages (Fig. 5.22): (1) the onset of diagenesis;
(2) formation of diagenetic cements, and synchronous
hydrothermal alteration and mineralisation; (3) diagenetic
alteration and compaction synchronous with emplacement
of the White Spur Formation; (4) replacement of early
diagenetic minerals and remaining glass by more stable mineral
assemblages; and (5) regional Devonian metamorphism and
deformation.
SEAFLOOR-AND BURIAL-RELATED ALTERATION | 1 33
Subtle, pervasive albite + quartz + chlorite alteration facies
Least-altered rhyoiite
CV C1
Sample no.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
133921
subtle, pervasive albite + quartz + chlorite
albite zone
Mount Black
Mount Black Formation
Central V olcanic Complex
massive, plagioclase-phyric rhyolite
plagioclase
porphyritic, micropoikilitic
rhyolite
massive
coherent facies
albite + quartz + sericite + chlorite +
hematite
albite + quartz sericite chlorite
pseudomorphs of plagioclase,
micropoikilitic albite + quartz, interstitial
chlorite, disseminated hematite
pervasive
excellent
subtle
pre-S
2
diagenetic
Hand specimen photograph
Geochemistry
SiO
2
74.58 K
2
O 4.34 Cu 2 Al 56
TiO
2
0.27 P
2
O
5
0.03 Pb 3 CCPI 22
AI
2
O
3
13.85 S <0.01 Zn 17 Ti/Zr 5.98
Fe
2
O
3
2.08 Total 100.32 Th 22
MnO 0.01 Zr 270
MgO 0.38 Rb 136 Nb 17
CaO 0.13 Sr 96 Y 41
Na
2
O 3.54 Ba 988
Photomicrograph (ppl)
1 3 4 | CHAPTER 5
Weak, pervasive albite + quartz * sericite alteration facies CVC2
Sample no.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
147407
weak, pervasive albite + quartz + sericite
albite zone
120R-438.5 m
Kershaw Pumice Formation
Central Volcanic Complex
jigsaw fit, monomictic, plagioclase-phyric
rhyolite breccia
plagiociase
porphyritic, perlitic fractures, jigsaw fit
clasts
rhyolite
massive
in situ hyaloclastite
albite + quartz + sericite > chlorite + pyrite
> calcite
albite calcite pseudomorphs of
plagiociase, microcrystaiiine groundmass,
calcite veins, chlorite filled perlitic fractures
pervasive
moderate
weak
pre-S
2
diagenetic
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
74.01 K
2
O 1.82 Cu 2 Al 26
TiO
2
0.23 P
2
O
5
0.03 Pb 4 CCPI 29
AI
2
O
3
12.21 S 0.01 Zn 19 Ti/Zr 5.31
Fe
2
O
3
2.17 Total 100.6 1 Th 12
MnO 0.09 Zr 258
MgO 0.49 Rb 76 Nb 16
CaO 2.41 Sr 113 Y 36
Na
2
O 4.07 Ba 513
SEAFLOOR- AND BURIAL-RELATED ALTERATION | 1 3 5
Moderate, domainal albite + quartz + sericite with sencite + hematite chlorite
alteration facies
CVC3
Sample no.
Alteration facies
Alteration zone
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
147410
moderate, domainal albite + quartz +
sencite
albite zone
1 2 0R-5 2 4 .5 m
Kershaw Pumice Formation
Central V olcanic Complex
graded, plagioclase-phyric pumice breccia
plagioclase
tube pumice clasts, fiamme, plagioclase
crystal fragments, blocky rhyolite clasts
rhyolite
normally graded
syn-eruptive, mass-flow-emplaced pumice
breccia
albite + quartz + sericite + chlorite +
hematite + calcite
sericite fiamme, hematite styioiites, albite
veins, recrystallised albite + quartz +
sericite pumice clasts and matrix, albite +
sericite + calcite altered plagioclase
domainal
poor
moderate
pre-S
2
diagenetic
Geochemistry
SiO
2
TiO
2
AI
2
O
3
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
76 .08
0.19
10.6 6
1.6 7
0.09
0.47
2.39
4.71
K
2
O
P
2
O
5
S
Total
Rb
Sr
Ba
0.88
0.03
0.01
99.93
34
144
280
Cu
Pb
Zn
Th
Zr
Nb
Y
1
4
27
10
210
13
37
Al
CCPi
Ti/Zr
16
26
5.43
Hand specimen photograph Photomicrograph (xn)
1 3 6 | CHAPTER 5
Weak, pervasive sericite alteration facies CVC4
Sample no.
Alteration facies
Alteration zone
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
147552
weak, pervasive sericite
albite zone
Pieman Road
Kershaw Pumice Formation
Central V olcanic Complex
massive, plagioclase-phyric pumice
breccia
plagioclase
tube pumice ciasts, bubble wall shards,
plagioclase crystal fragments, fiamme
rhyolite
normally graded
syn-eruptive, mass-flow-emplaced pumice
breccia
sericite + albite + calcite + chlorite +
hematite
sericite fiamme, hematite stylolites,
disseminated calcite rhombs, albite +
sericite altered pumice ciasts and shards
pervasive
good
weak
pre-S
2
diagenetic
Hand specimen photograph
Geochemistry Geochemistry
SiO
2
70.91 K
2
O 3.16 Cu 4 Al 48
TiO
2
0.31 P
2
O
5
0.07 Pb 2 CCPI 36
AI
2
O
3
14.08 S 0.01 Zn 48 Ti/Zr 7.41
Fe
2
O
3
2.78 Total 99.75 Th
MnO 0.07 Zr 251
MgO 0.77 Rb 124 Nb 13
CaO 1.6 6 Sr 87 Y 28
Na
2
O 2.6 8 Ba 786
Photomicrograph (ppl)
SEAFLOOR- AND BURIAL-RELATED ALTERATION | 1 3 7
I
Subtle, pervasive aibite + quartz + chlorite alteration facies
Least-altered dacite
CVC5
Sample no.
Alteration facies
Alteration zone
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
147435
subtle, pervasive aibite + quartz + chlorite
aibite zone
MBD4 -1 8 .4 m
Mount Black Formation
Central V olcanic Complex
massive, plagioclase + hornblende-phyric
dacite
plagioclase, hornblende
porphyritic, glomeroporphyritic clusters,
micropoikilitic
dacite
massive
coherent facies
aibite + quartz + chlorite + epidote
aibite + quartz micropoikilitic groundmass
with interstitial chlorite + epidote, aibite
pseudomorphs of plagioclase, epidote +
chlorite altered hornblende
pervasive
excellent
subtle
pre-S
2
diagenetic
Geochemistry
SiO
2
6 7.53 K
2
O 3.95 Cu 4 Al 47
TiO
2
0.52 P
2
O
5
0.13 Pb 4 CCPI 41
AI
2
O
3
14.51 S 0.01 Zn 51 Ti/Zr 14.48
Fe
2
O
3
4.37 Total 99.51 Th 15
MnO 0.06 Zr 216
MgO 1.3 Rb 102 Nb 12
CaO 2.38 Sr 242 Y 34
N a, 0 3.56 Ba 958
Hand specimen photograph Photomicrograph (ppl)
1 3 8 | CHAPTER 5
Moderate, domainal chlorite + epidote alteration facies CVC6
Sample no.
Alteration facies
Alteration zone
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
147557
moderate, domainal chlorite + epidote
epidote zone
Pieman Road
Mount Black Formation
Central V olcanic Complex
jigsaw fit, monomictic plagioclase +
homblende-phyric dacite breccia
plagioclase, hornblende
glomeroporphyritic, perlitic fractures,
jigsaw-fit clasts
dacite
massive
in situ hyaloclastite
albite + quartz + chlorite + epidote
microcrystalline groundmass with domainal
albite + quartz and chlorite + epidote
facies, plagioclase phenocrysts albite or
chlorite epidote altered, hornblende
altered to chlorite + epidote
domainal
good
weak
pre-S2
diagenetic
Geochemistry Geochemistry
SiO
2
6 7.93 K
2
O 2.05 Cu 10 Al 31
TiO
2
0.59 P
2
O
5
0.13 Pb 3 CCPI 44
AI
2
O
3
14.32 S 0.01 Zn 28 Ti/Zr 17.6 8
Fe
2
O
3
4.57 Total 99.6 5 Th
MnO 0.06 Zr 201
MgO 1.33 Rb 41 Nb 12
CaO 2.6 7 Sr 151 Y 31
Na
2
O 4.97 Ba 826
Hand specimen photograph Photomicrograph (ppl)
1 39
6 | SYNVOLCANIC INTRUSION-RELATED
ALTERATION
The spatial and genetic associations between intrusions
and altered zones are widely appreciated in porphyry and
epithermal districts (Lowell and Guilbert, 1970; Titley, 1982;
Henley and Brown, 1985). Similar relationships also exist
in VHMS districts, where synsedimentary or synvolcanic
intrusions are commonly altered and surrounded by halos of
altered rocks. In some VHMS districts (e.g. Snow Lake and
Sturgeon Lake, Canada), there are spatial associations between
synvolcanic intrusions and broad-scale, semi-conformable
altered zones and clusters of VHMS deposits in the overlying
successions (Spooner and Fyfe, 1973; Campbell et al., 1981;
Gibson and Watkinson, 1990; Galley, 1993; Hannington et
al., 2003a). It has been suggested that synvolcanic intrusions
were heat sources (Spooner and Fyfe, 1973; Ohmoto and
Rye, 1974; Solomon, 1976; Cathles, 1977; Franklin et al.,
1981; Polya et al., 1986; Galley, 1993; Large et al., 1996),
and perhaps also volatile and metal sources (Urabe and Sato,
1978; Stanton, 1990; Yang and Scott, 1996; Hannington et
al., 1999) for subseafloor hydrothermal systems that formed
altered zones and VHMS deposits.
Synvolcanic intrusive sills, cryptodomes, dykes and sub-
volcanic plutons are volumetrically important in submarine
volcanic successions (Polya et al., 1986; McPhie and Allen,
1992; Doyle and Huston, 1999; Galley, 2003). They may be
composite intrusions of variable volumes up to 1000 km
3
,
typically emplaced at depths up to 4 km below the seafloor
(Nielsen et al., 1981; Galley, 2003; Whalen et al., 2004).
Intrusions and intrusion-related altered zones that significantly
post-date volcanism are also common in ancient submarine
volcanic successions; however, they are not the focus of this
chapter.
Alteration can occur within intrusions (deuteric and
local hydrothermal alteration), locally in the immediate
host rocks (contact alteration) or regionally in the host
succession (regional hydrothermal alteration) (Fig. 6.1). This
chapter describes the role of intrusions in generating regional
hydrothermal systems, regional hydrothermally altered zones,
altered zones within intrusions and contact altered zones
around both small-volume, near-seafloor and larger, deeper
intrusions in submarine volcanic successions. The final section
presents a case study of contact altered zones associated with
the Darwin Granite in the southern Mount Read Volcanics,
western Tasmania. The recognition of altered zones related
to synvolcanic intrusions can provide insights into fluid-flow
and thermal histories of VHMS districts, and thereby assist
mineral exploration.
FIGURE 6 .1 | A cartoon of the variety of altered zones associated with synvolcanic intrusions. (A) A deuteric altered zone within
the top of a large volume intrusion. (B) A fracture-controlled hydrothermally altered zone at the margins of an intrusion and in the
surrounding host rocks. (C) Contact-altered zones around synvolcanic sills emplaced into unconsolidated sediment immediately
below the seafloor. (D) Concentric contact-altered zones around a large volume intrusion emplaced at depth. (E) Regional
hydrothermally altered zones related to emplacement of a subvolcanic pluton. (F) Afootwall alteration pipe beneath a VHMS deposit.
1 4 0 I CHAPTER 6
6.1 | THE ROLE OF INTRUSIONS IN
GENERATING HYDROTHERMAL
SYSTEMS
The most active hydrothermal systems are those related to
magma-induced thermal anomalies (Alt, 1999; Butterfleld,
2000). The magma chamber provides heat to overlying
strata and active volcanism contributes heat from its eruptive
products, intrusions and feeder dykes. The transfer of heat and
mass away from the intrusion may occur by either conduction
only, or conduction and infiltration. Conduction generally
involves only minor diffusion of elements, although Weaver et
al. (1990) suggested that at near solidus temperatures vapour-
phase expulsion may produce local mineral and chemical
variations (loss of Na, halogens and REE) in volcanic glass.
In contrast, conduction accompanied by infiltration and
circulation of hot fluid can remove heat from the magmatic
system much faster than conduction alone, and effectively
transport elements considerable distances, up to hundreds of
kilometres, through the succession.
Thermal metamorphism related to the shallowemplacement
of synvolcanic intrusions in dry successions typically results
in limited alteration with little or no mass transfer. In rare
cases, magmatic fluids exsolved from the crystallising magma
hydrothermally alter dry host facies. Vapour-phase expulsion
of some elemental species as complexes (e.g. fluoride,
chloride, hydroxide, sulfide and carbon dioxide) may result
in minor losses as glassy rocks devitrify, and glassy clasts may
be welded by elevated temperatures in the contact zones (e.g.
Christiansen and Lipman, 1966).
The effects of intrusions emplaced into water-saturated
successions are very different because water mobilises heat
and soluble elements. Trapped seawater in submarine volcanic
successions is heated by intrusions, initiating convection and
metasomatic alteration in the overlying succession. Thus,
almost all intrusion-related alteration in submarine volcanic
successions involves some degree of metasomatism by
magmatic fluid, modified seawater, or both.
Subseafloor regional hydrothermal systems
Studies of the petrology, geochemistry and oxygen isotopes
of hydrothermally altered volcanic and plutonic rocks
from ophiolite complexes provide insight into subseafloor
hydrothermal systems, fluid generation and circulation, and
the role of intrusions (e.g. Lydon and Jamieson, 1984; Alt
et al., 1986; Gillis and Robinson, 1990; Bettison-Varga et
al., 1992; Kelley et al., 1992). The convection cell model for
hydrothermal systems and the formation of VHMS deposits
is based on observations from VHMS deposits and the upper
part of the Cretaceous Troodos Massif in Cyprus, where
hydrothermal convection was driven by emplacement of late,
high-level gabbro stocks into the fractured and permeable
crust (e.g. Spooner et al., 1974; Lydon and Jamieson, 1984;
Bettison-Varga et al., 1992). This model involves the circulation
of seawater in approximately 10 km diameter cells to depths
of 3-5 km within the crust (Fig. 6.2). Initially, increased
temperatures in the host succession drive dehydration and
decarbonation reactions, and fluids migrate away from the
intrusion. Buoyant heated connate seawater rises through the
permeable volcanic succession, drawing down cold seawater,
which is heated as it descends. In this way, magma drives
convective circulation of seawater between the seafloor and
the intrusion (Norton, 1984; de-Ronde et al., 1994; Galley,
2003). Fluid flow is focused along joints, fractures and faults
formed during extension or in response to intrusive pressures
(Bettison-Varga et al., 1992). Alternatively, the multi-tiered
convection model involves a high-temperature (450-700C)
cell, which circulates recycled modified seawater in plutonic
rocks at depth, overlain by a low-temperature (350-400C)
cell (Gregory and Taylor, 1981; Norton et al., 1984; Alabaster
and Pearce, 1985; Kelley et al., 1992).
Submarine hydrothermal systems comprise three parts:
a down-flow or recharge zone; a high-temperature reaction
zone; and an up-flow or discharge zone (Fig. 6.3: Spooner
and Fyfe, 1973; Alt, 1999). The locations of the recharge and
discharge zones are commonly controlled by faults (Schardt et
al., in press). Seawater percolates down through the recharge
zone, and is slowly heated and chemically modified by low-
temperature reactions (White, 1970; Gibson et al., 1983;
Galley, 1993; Alt, 1999). The reaction zone is a porous
reservoir near the heat source where heated seawater reacts
with the host rocks, exchanging some elements (Norton,
1984; de-Ronde et al., 1994; von Damm, 1995; Butterfield,
2000; Schardt etal., in press). Hot buoyant hydrothermal fluid
(modified seawater) ascends rapidly to the seafloor through
the discharge zone, which is characterised by cooling of the
fluid, alteration of the host rock, and mineral precipitation
(Skirrow and Franklin, 1994; Schardt et al., in press). The
rising hydrothermal fluid cools by adiabatic decompression,
conductive heat loss, and mixing with cold seawater in the
shallow subsurface (Mottl, 1983; Butterfield, 2000). In well-
FIGURE 6 .2 | Simple convection cell model for the genesis of
the Cyprus VHMS deposits (modified after Heaton and Sheppard,
1977, and Spooner, 1977, in Lydon and Jamieson, 1984).
established hydrothermal systems, the discharge zone may be
focused, intensely altered and veined. Surface discharge onto
the seafloor may produce high-temperature (150350C)
features such as black smokers (Goodfellow and Franklin,
1993; Rona et al., 1993). The temperature of the discharging
fluids on the seafloor initially increases, and then gradually
decreases to ambient temperatures, in a time scale of 100 to
10,000 years (Ohmoto, 1996).
Regional hydrothermal systems are interpreted to be
related to large volume intrusions, as the volume of circulating
fluid in a hydrothermal system theoretically cannot be greater
than the volume of the intrusion (Cathles, 1981). However,
small-volume near-seafloor intrusions, which are unlikely to
generate significant hydrothermal systems, may be related
to larger plutons or stocks at depth that were capable of
generating hydrothermal convection (e.g. Bettison-Varga et
al., 1992).
6.2 | REGIONAL ALTERED ZONES
ASSOCIATED WITH INTRUSIONS
The products of regional-scale hydrothermal alteration
systems in ancient submarine volcanic successions are
recorded by cross-cutting recharge and discharge zones, and
broad, regional-scale, semi-conformable altered zones or
reaction zones (Galley, 1993).
Recharge zones
Very little is known about altered zones associated with
recharge. They are rarely recognised except in studies of
modern crustal alteration beneath mid-ocean ridges (e.g.
Mottl, 1983; Saccocia et al., 1994; Alt, 1999), hydrothermal Discharge ZOneS
SYNV OLCANIC INTRUSION-RELATED ALTERATION | 1 4 1
alteration studies in ophiolites (e.g. Schiffman et al., 1987;
Schiffman and Smith, 1988), and studies of O- and S-isotope
compositions in ancient hydrothermally altered systems (e.g.
Cathles, 1993; Davidson and Kitto, 1997). Rocks in modern
recharge zones are pervasively altered at low to moderate
temperatures. At less than 150C, oxidation, the fixation of
alkalis (mainly Ca and Na), and Mg-metasomatism produces
sericite, hematite and clays (Alt, 1999). At higher temperatures
(150-350C) anhydrite precipitates, alkalis are leached and
Mg is consumed by chlorite in the rock (Alt, 1999).
Schiffman and Smith (1988) proposed that the distribution
of epidosites in the Troodos ophiolite represent areas of
high-temperature alteration involving high fluid-rock ratios.
Epidosites are granoblastic, fine- to medium-grained rocks,
with little or no relict igneous textures, composed of epidote,
quart and chlorite. They are inferred to record reaction zones
in which circulating modified seawater reacted with host
rocks to form metal-rich hydrothermal fluids, and appear
diagnostic of the up-welling and deep recharge parts of the
hydrothermal system beneath VHMS deposits. Co-incident
whole-rock O-isotope patterns support their formation in
proximal recharge zones and up-flow conduits beneath VHMS
deposits. Regionally extensive, depth-dependent 6
18
O profiles
in the sheeted dyke complex reflect oxygen exchange during
prograde regional hydrothermal alteration involving diffuse
down-welling of cold seawater (Fig. 6.4). However, surfaces of
equal whole-rock 6
18
O are not horizontal but nearly vertical
in the central epidosite zone. This suggests up-flow of hot
modified seawater within the epidosite zone.
The spatial association between gabbro intrusions and
the epidosite zones in the sheeted dyke complex indicates a
genetic link between the emplacement of these intrusions and
focused high-temperature hydrothermal up-flow (Richardson
et al., 1987; Bettison-Varga et al., 1992).
Discordant footwall alteration pipes and feldspar-destructive
zones that directly underlie VHMS deposits are widely
interpreted as discharge zones through which metal-bearing
hydrothermal fluid ascended to the seafloor (Sangster,
1972; Large, 1977; Lydon, 1984; Galley, 1993; Skirrow and
Franklin, 1994; Brauhart et al., 1998). They are characterised
FIGURE 6.3 | Model of an active geothermal system illustrating the recharge,
reaction or reservoir and discharge zones. Seawater is drawn down in broad
recharge zones or along faults and reacts at increasing temperatures. High-
temperature reactions (>350C) occur in the reaction zone above a subvolcanic
intrusion and hot (>300C) buoyant fluids rise towards the surface in focused or
diffuse discharge zones (modified after Alt, 1995a). Not to scale.
FIGURE 6.4 | Cross-section of the Solea graben, Troodos ophiolite, Cyprus,
showing surfaces of equal whole-rock d
1 8
0. Regionally these surfaces are sub-
horizontal, but in the central epidosite zone they are nearly vertical indicating
up-flow of hot, modified seawater during convection. Modified after Schiffman
and Smith (1988).
1 4 2 | CHAPTER 6
by Mg-Fe enrichment and Na-Ca depletion and assemblages
that include chlorite, sericite, quartz or rare talc (Lydon,
1984; Eastoe et al., 1987; Skirrow and Franklin, 1994;
Brauhart et al., 1998). The characteristics and compositional
changes associated with discordant footwall alteration pipes
are discussed in Section 7.3.
Although discordant altered zones typically cut across
the regional, deep, semi-conformable altered zones (Galley,
1993; Brauhart et al., 1998), in some successions, they grade
laterally into deep, semi-conformable altered zones (Skirrow
and Franklin, 1994; Hudak et al., 2000). Gibson et al.
(2000) suggested that whether or not deep, semi-conformable
altered zones are cut by or transitional with pipe-like altered
zones, depends on whether the host succession (footwall)
is dominated by coherent volcanic or volcaniclastic facies
respectively, or timing of alteration.
Deep, semi-conformable altered zones
Since they were first discussed by Franklin et al. (1981), deep,
semi-conformable altered zones have been documented in
the footwall beneath VHMS deposits in a variety of districts
including: Matagami, Snow Lake, Noranda and Sturgeon
Lake districts in Canada; Bersglagen and Skellefte districts in
Sweden; Iberian pyrite belt in Spain and Portugal; Troodos
Ophiolite Complex in Cyprus; Panorama district in Australia;
and the Sirohi district in India (MacGeehan, 1978; Gibson et
al., 1983, 2000; Lagerbald and Gorbatschev, 1985; Galley,
1993; Skirrow and Franklin, 1994; Tiwary and Deb, 1997;
Brauhart et al., 1998; Bailes and Galley, 1999; Hannington et
al., 2003a, 2003b). They have not been documented in eastern
Australia, possibly because of structural complexities. Thus
the following discussions on deep, semi-conformable altered
zones are largely based on Canadian examples. Figure 6.5
depicts the characteristics and typical zonation of deep, semi-
conformable altered zones in the documented examples.
Deep, semi-conformable altered zones typically extend
for up to 20 km laterally and 1-4 km depth beneath paleo-
seafloors and VHMS deposits (Gibson et al., 1983, 2000;
Cathles, 1993; Galley, 1993; Skirrow and Franklin, 1994).
They comprise vertically stacked, sub-horizontal altered zones
(Galley, 1993; Skirrow and Franklin, 1994). Generally these
are (Fig. 6.5): an upper background K-Mg metasomatic zone;
a transitional Na-Mg metasomatic zone; a central silicified
zone; and a basal Ca-Fe metasomatic and base metal-leaching
zone (Galley, 1993). In many systems only one or two of
these altered zones are recognised. The alteration minerals in
the semi-conformable altered zones reflect the primary host
rock composition, bulk-rock composition established during
synvolcanic hydrothermal alteration, and the subsequent
metamorphic grade (Paradis et al., in press). In greenschist
facies metamorphosed felsic rocks, these zones are typically,
from base to top: albite or carbonate zone, silica or sericite
zone and sericite or chlorite zone (Gibson et al., 1983; 2000).
In mafic rocks, the zones are: albite zone, silica zone and
clinozoisite/epidote + quartz zone (Galley, 1993; Skirrow
and Franklin, 1994; Gibson et al., 2000; Hannington et al.,
2003a). At higher metamorphic grades, such as in the Snow
Lake District, mineral assemblages can include kyanite,
staurolite, sillimanite, chlorite, biotite, quartz, plagioclase
cordierite, amphibole, epidote and garnet (Paradis et al.,
1993, in press; Bailes and Galley, 1999).
The semi-conformable altered zones are interpreted to
be synvolcanic because they have undergone the same degree
of tectonic deformation as the surrounding rocks, have
prograde mineral assemblages, are spatially associated with
VHMS deposits, and are commonly truncated by unaltered
synvolcanic intrusions (Gibson et al., 1983; Paradis et al.,
1993; Skirrow and Franklin, 1994). At Snow Lake, Paradis
et al. (1993) recognised that deep, semi-conformable altered
zones were superimposed on low-temperature (possibly
diagenetic) altered zones and also cross cut by discordant
feldspar-destructive zones associated with VHMS deposits.
PROCESS AND
COMPOSITIONAL
CHANGES
K-Mg metasomatism
+ Mg, K, Fe
-N a, Ca, Cu, Pb, Zn, Si
Na-Mg metasomatism
+ Na, Mg
- Ca, Fe, Zn.Cu
Si
Silicification or sericitisation
+ Si, Na
-F e, Mg, Mn,Zn
Ca
Ca-Fe metasomatism
+ Ca
- Mg, Mn, Na, K
Fe, Si
ASSEMBLAGE IN
MAFIC ROCKS
ASSEMBLAGE
IN FELSIC ROCKS
Mg clays + chlorite
+ zeolites +
Fe-oxides
K-feldspar
Albite + quartz +
sericite +
Mg-chlorite calcite
Quartz + albite
Clinozoisite/epidote
+ quartz
actinolite
carbonate
Mg-clays + zeolites
cristobalite adularia
analcime
K-feldspar
Albite + quartz + sericite
+ Mg-chlorite
Quartz albite sericite
Sericite + quartz
Mg-chlorite or chloritoid
+ Fe-chlorite
FIGURE 6 .5 | A schematic compilation of regional-scale, deep, semi-conformable altered zones and their characteristics. There is a progression, with increasing
depths in submarine volcanic successions, from the background Mg-K metasomatic zone to a transitional Na-Mg metasomatic zone characterised by feldspar
alteration, a central silicified zone and a basal Ca-Fe metasomatic and base-metal leaching zone that typically includes epidote or chlorite, After Gibson et al. (1983),
Galley (1993), Skirrow and Franklin (1994), and Brauhart etal. (1998).
They suggested that regional hydrothermal alteration post-
dated the onset of diagenetic alteration, and pre-dated footwall
alteration associated with hydrothermal alteration and
mineralisation. In some Canadian examples and at Panorama
in western Australia, deep, semi-conformable altered zones are
gradational with discordant footwall alteration pipes suggesting
that regional hydrothermal alteration was synchronous with
the VHMS-related alteration (Gibson et al., 1999).
Deep, semi-conformable altered zones are commonly
spatially and temporally associated with subsurface syn-
volcanic intrusions (Galley, 1993, 2003). These intrusions
can be individual granitic or porphyritic plutons or sheeted
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 4 3
dyke swarms (e.g. Gibson et al., 1983; de-Ronde et al., 1994;
Brauhart et al., 1998). The tops of the subvolcanic intrusions
and associated dykes may be included in the basal semi-
conformable altered zone (e.g. Galley, 1993; Brauhart et al.,
1998).
One of the best-documented examples of the spatial
association between a subvolcanic intrusion, regional-scale
semi-conformable altered zones and VHMS deposits comes
from the Panorama district in Western Australia. Discordant
chlorite + quartz zones directly beneath the VHMS deposits,
are spatially associated with feldspar-destructive sericite
+ quartz zones in the top of the Strelley Granite pluton
FIGURE 6 .6 | Geology and altered zones within part of the Strelley succession, Panorama district, Western Australia (modified after
Brauhart et al., 1998).
1 4 4 | CHAPTER 6
FIGURE 6.7 | Schematic section of the geology and altered zones in the
Kangaroo Caves footwall succession, Panorama district, Western Australia
(modified after Brauhart et al., 1998). See Figure 6 .6 for legend.
(Figs 6.6 and 6.7; Brauhart et al., 1998). Faults bounding
the discordant chlorite + quartz zone in the footwall of the
Kangaroo Caves deposit (Fig. 6.7) controlled the distribution
of the feldspar-destructive sericite + quartz zone in the Strelley
Granite (Brauhart et al., 1998).
The variations in alteration mineral assemblage down
through the semi-conformable altered zones correspond to
geochemical gradients in which there are gradual decreases
in the Mg/Ca, Mg/Na and Na/Ca ratios of the altered rocks
with increasing depths (Galley, 1993). Oxygen-isotope
compositions suggest that the altered rocks are
18
O enriched
with respect to unaltered volcanic rocks (Munha and Kerrich,
1980; Barringa and Kerrich, 1984). The geochemical
gradients and O-isotope data are consistent with metasomatic
alteration resulting from the interaction of volcanic rocks
with seawater (Muehlenbachs and Clayton, 1972; Lagerbald
and Gorbatschev, 1985; Cathles, 1993).
Although modified seawater is interpreted to be the main
component, magmatic fluids may have also contributed to
the hydrothermal fluid (Lagerbald and Gorbatschev, 1985).
The spatial association between altered zones and subsurface
intrusions suggests a genetic link where intrusions may have
provided heat and or fluid to the hydrothermal system.
The extent and intensity of deep, semi-conformable
altered zones implies that very large volumes of fluid must have
reacted with the host volcanic rocks (Skirrow and Franklin,
1994). In the Snow Lake district Skirrow and Franklin
(1994) estimated that approximately 1.1 x 10
7
metric tons
of SiO
2
was added by a minimum 12 km
3
of hydrothermal
fluid at 12 km depth. Interactions between large volumes
of modified seawater and volcanic successions at depth are
supported by geochemical and geophysical research at active
ocean spreading ridges (e.g. Spooner and Fyfe, 1973; Bischoff
andDickson, 1975).
Deep, semi-conformable altered zones are characterised
by mineral assemblages that reflect the reactions of glass
and both primary and secondary minerals with seawater at
temperatures up to 400C (Galley, 1993).
Background K-Mg metasomatic zones
These zones are often described as the least-altered or
diagenetically altered zones. At low temperatures (50-140C)
in the shallow subseafloor, the interaction of abundant
seawater with the volcanic succession produces Mg-K-rich
alteration assemblages (Seyfried and Bischoff, 1977; Galley,
1993). In felsic rocks these mineral assemblages include
adularia and Mg-smectite, whereas in mafic rocks they are
dominated by zeolites and Mg-smectite. Seawater becomes
enriched in Si, Fe
3+
, Mn and lesser amounts of Ca, Mg and
sulfur (Seyfried and Bischoff, 1977).
Current mineral assemblages reflect the regional
metamorphic grade. For example, at Snow Lake the dia-
genetically altered zone is characterised by quartz + biotite +
garnet, Fe
2
O
3
, MgO and K
2
O gains and CaO, Na
2
O, Cu, Pb,
Zn losses (Paradis et al., in press). These compositional changes
are consistent with low-temperature seawater-dominated
diagenesis of felsic volcanic facies to clays and zeolites (Section
5.3). The current mineral assemblage reflects the overprint of
amphibolite facies metamorphism. In the Panorama district
the background alteration mineral assemblage includes
feldspar + calcite ankerite + quartz + pyrite sericite
consistent with greenschist facies metamorphism of clays and
zeolites in felsic volcanic rocks (Brauhart et al., 1998).
Transitional zone or Na-Mg metasomatic zones
With increasing stratigraphic depth there is a transition from
K-rich zones to Na-rich zones (Munha et al., 1980; Munha
and Kerrich, 1980; Lagerbald and Gorbatschev, 1985;
Schiffman and Smith, 1988; Brauhart et al., 2001). The Na-
Mg metasomatic zones are characterised by the occurrence
of feldspar, usually albite. Greenschist facies assemblages
typically include chlorite, sericite, albite, epidote and quartz in
mafic rocks, and albite, quartz, sericite chlorite carbonate
(calcite or dolomite) in felsic rocks (Gibson et al., 2000). In
the Panorama district, felsic rocks in the feldspar zone have the
assemblage K-feldspar or albite + sericite + quartz + ankerite +
leucoxene pyrite (Brauhart et al., 1998).
The transition to Na-rich zones reflects the behaviour of Na
and K in seawater at elevated temperatures. Between 140 and
200C there is a transition between K- and Na-metasomatism
(Seyfried and Bischoff, 1977). Munha et al. (1980) suggested
that at lower temperatures (<150C), Na in glass is exchanged
for K in seawater, resulting in precipitation of K-rich zeolites
and possibly K-feldspar. At higher temperatures, K in the rock
is exchanged for Na in seawater, resulting in the formation
of albite. Thus at moderate temperatures (140300C),
metasomatic reactions between modified seawater and the
volcanic succession result in Na-Mg alteration assemblages
(Seyfried et al., 1988). Regardless of the rock type, alteration
mineral assemblages include Mg-smectite + chlorite + quartz
+ albite, and compositional changes are Na
2
O and MgO
gains, and CaO, Zn and Cu losses (Gibson et al., 2000). The
removal of Mg from seawater lowers the pH of the fluid and
seawater evolves from a moderately alkali, Mg-K-Na-SO
4
-
rich fluid to a hot acidic Si-Na-Ca-rich hydrothermal fluid
(Bischoff and Seyfried, 1978; Seyfvied et al., 1988).
Silicified zones
In greenschist facies felsic and some mafic rocks, the central
altered zone is typically silicified, with assemblages of quartz
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 4 5
+ plagioclase or albite (Skirrow and Franklin, 1994; Gibson
et al., 2000). In some mafic rocks, the central zone is sericitic,
dominated by sericite + quartz chlorite (Gibson et al.,
2000). Central silicified or sericite zones overprint regional
albite zones (Galley, 1993). At Snow Lake the silicified zone
is spatially and temporally associated with VHMS deposits
and is zoned laterally from a silica zone to epidote and Fe-
Mg-metasomatic zones (amphibolite grade; garnet + chlorite
biotite staurolite) (Skirrow and Franklin, 1994). Silicified
zones are typically spatially associated with synvolcanic
intrusions and the intensity and pervasiveness of alteration
increases with proximity to the intrusions (Skirrow and
Franklin, 1994; Paradis et al., in press).
Silicified zones commonly contain patches of quartz +
feldspar-altered rock, quartz-altered clasts in volcaniclastic
facies, and quartz veins (e.g. Fig. 6.8A: Gibson et al., 1983;
Skirrow and Franklin, 1994). The patches of quartz + feldspar-
altered rock are restricted to flow-top breccias, and flow-
A. Central silicified zone
Moderate, patchy quartz alteration in this andesite from
the central silicified zone resulted in a fine-grained, pale
rock, which resembles a rhyolite.
Amulet Formation, Noranda district, Buttercup Hill,
Canada.
B. Epidote + quartz zone
This approximately one metre-wide patch of epidote
+ quartz alteration facies in the upper Amulet andesite
has an irregular shape typical of patchy alteration in the
basal episite + quartz zone. The groundmass has been
pervasively epidote + quartz altered.
Amulet Formation, Noranda district, Canada.
C. Epidote + quartz zone
Amygdales in this patch of epidote + quartz-altered
andesite from the basal epidote + quartz zone have
amoeboid shapes and were lined with Fe-oxides and
filled with epidote + quartz.
Amulet Formation, Noranda district, Canada.
FIGURE 6.8 | Photographs from deep semi-conformable alteration zones in
the Noranda district, Canada.
1 4 6 | CHAPTER 6
banded and vesicular lavas (Gibson et al., 1983). In mafic
volcanic rock the originally glassy groundmass, elsewhere
typically altered to chlorite, is altered to quartz in this zone
(Gibson et al., 2000). This led to intensely silicified andesites
in the Noranda sequence being misinterpreted as rhyolite: the
Amulet rhyolite (Gibson et al., 1983).
With increasing depth, seawater carries larger amounts of
Si as Si solubility increases with temperature and pressure,
and is enhanced in NaCl solutions or where the fluid is in
contact with free Si or glass (Kennedy, 1950; Fournier,
1985). The Si-rich hydrothermal fluid is rapidly heated
beyond the temperature of the quartz solubility maximum:
340-400C at pressures below 900 bars (Fig. 6.9: Kennedy,
1950; MacGeehan, 1978; Skirrow and Franklin, 1994). The
result is gains in SiO
2
and Na
2
O, due to the precipitation of
silica within pore spaces and albitisation, and losses in FeO,
MgO, CaO, K
2
O, MnO and other metals from the volcanic
rocks (Gibson et al., 1983, 2000; Lagerbald and Gorbatschev,
1985; Galley, 1993; Skirrow and Franklin, 1994). The Fe
3+
,
Mg and possibly Zn leached from the silicified zone may
have been transported laterally away from this environment,
thereby producing semi-conformable Fe-Mg-metasomatised
zones (Skirrow and Franklin, 1994).
A second silicified zone is common directly beneath the
seafloor in the Snow Lake, Noranda and Matagami Lake
districts, where it is directly overlain by exhalites. This near-
seafloor, silicified zone is related to low-temperature silici-
fication during the hydrothermal alteration and devitrification
of glass in cooling pillow basalts and andesites (Skirrow and
Franklin, 1994; Galley et al., 2002).
FIGURE 6.9 | Calculated solubilities for quartz in water up to 900C at
pressures between 200 and 1000 bars (after Fournier, 1985). The shaded area
outlines the conditions for retrograde solubility.
Basal Ca-Fe metasomatic zones
Mineral assemblages in basal semi-conformable altered zones
are dependent on the host-rock composition and porosity. In
felsic rocks, basal zones are typically sericite or chlorite zones,
whereas in mafic rocks they are clinozoisite or epidote +
quartz zones (Gibson et al., 2000; Hannington et al., 2003a).
Sericite zones are characterised by sericite + quartz Mg-
chlorite assemblages (Gibson et al., 2000). Chlorite zones
are characterised by chloritoid + Fe-chlorite Fe-carbonate
assemblages (Gibson et al., 2000). Epidote + quartz zones
are characterised by epidote + quartz + calcite + actinolite +
chlorite assemblages (Galley, 1993).
Two alteration textures are persistent in epidote + quartz
zones: pervasive and patchy. Pervasive epidote + quartz occurs
as selective pervasive replacement of plagioclase phenocrysts
or the groundmass by epidote, crystallisation of fine quartz
patches in the groundmass (e.g. Fig. 6.8B), and replacement
of Fe-Ti-oxide grain rims by sphene (Skirrow and Franklin,
1994). Patchy epidote + quartz occurs as less than 1 cm to
2 m irregular ovoids or amoeboid patches that infill vesicles
and gas cavities within mafic lavas (e.g. Fig. 6.8C: Gibson et
al., 1983; Skirrow and Franklin, 1994). These patchy textures
are similar to the epidote + quartz metadomains described in
spilites by Smith (1968, 1974, 1977; Smith et al., 1982).
These basal zones may grade into the discordant
(discharge) altered zones that cut through the overlying semi-
conformable and background altered zones (Brauhart et al.,
1998).
Epidote + quartz zones are enriched in CaO and Sr and
depleted of MgO, Na
2
O, K
2
O, FeO MnO, Ba and base
metals (MacGeehan, 1978; Gibson et al., 1983; Richardson et
al., 1987; Schiffman and Smith, 1988; Skirrow and Franklin,
1994). Unlike the Canadian examples, the epidote + quartz
zone in the Panorama district does not appear to have been
the source of leached base metals (Brauhart et al., 2001).
At high temperatures (300-500C), Ca-Fe-S-base metal-
rich hydrothermal fluid reacts with the volcanic succession
and possibly also with parts of the subsurface intrusion
forming mineral assemblages typical of the basal Ca-Fe-
metasomatic zones. Experimental work suggests that epidote
+ quartz alteration involved modified seawater (Mg-depleted,
Ca-Na-K-Cl fluid) at temperatures of 35O-5OOC and low
water-rock ratios of less than three (Gibson et al., 2000).
This zone is interpreted to represent the high-temperature
interaction between modified seawater and the host volcanic
facies to form metal-rich hydrothermal fluid at the deepest part
of the hydrothermal convection system. Hence, it represents
the roots of up-welling fluid discharge zones (Galley, 1993).
Alternatively, Smith (1968, 1977) interpreted these
district-scale zones of albitised basalt with Ca-rich epidote +
quartz and pumpellyite + quartz metadomains to have formed
during heterogenous burial metamorphism where local fluid
flow promoted redistribution of elements. In some cases, he
noted that the alteration was focussed suggesting that it was
related to local hydrothermal systems and subseafloor fluid
circulation.
Altered zones as part of a regional hydrothermal
system
Deep, semi-conformable altered zones superficially resemble
regional diagenetic or metamorphic facies. This is because
they are regionally extensive, vertically stacked altered
zones with mineral assemblages similar to those formed
during high-temperature diagenesis, regional greenschist
facies metamorphism and hydrothermal seafloor alteration
(Galley, 1993; Paradis et al., 1993). Discriminating between
these processes and their products is difficult in submarine
volcanic successions. The differences are essentially related to
timing, temperatures, and fluid-rock ratios. In reality, there is
a progression from diagenesis to isochemical metamorphism
with increasing temperature and depth during burial (Fig.
6.10A). Porosity, permeability and fluid-rock ratios decrease
with depth in diagenetic-metamorphic systems, thereby
inhibiting the degree and pervasiveness of metasomatism at
temperatures above 150C. Typically once the temperature
and pressure realm of metamorphism has been reached, the
porosity and permeability of the host succession has been
dramatically reduced, fluid flow inhibited and metasomatic
reactions ceased. Gibson et al. (2000) suggested seawater-
dominated diagenesis might also progress to deep regional
hydrothermal alteration with increasing temperature and
depth in shallow subseafloor hydrothermal systems (Fig.
6.1 OB). Deep regional hydrothermal alteration is interpreted
to involve metasomatic reactions between seawater and
the volcanic succession at temperatures transitional with
diagenesis and greenschist facies metamorphism (i.e. 150
400C) (Galley, 1993). Although the processes of diagenesis
and deep regional hydrothermal alteration are very similar,
and both involve reactions between seawater (or modified
seawater) and volcanic successions at increasing temperatures
and depths, deep, semi-conformable altered zones are
inconsistent with the diagenetic-metamorphic system. They
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 4 7
have anomalous mineral assemblages and alkali contents for
igneous rocks (e.g. Fig. 6.11; Hughes, 1973), which suggest
metasomatic rather than metamorphic origins (Gibson et al.,
1983; Galley, 1993).
Gibson et al. (1983) documented a vertically stacked
sequence of altered zones in the Noranda sequence, from top
to bottom: albite zone (spilites), silicified zone, and epidote +
quartz zone (Fig. 3.16). This is consistent with a progression
from low-temperature diagenesis to moderate- and high-
temperature metasomatism with depth in the stratigraphy.
Munha and Kerrich (1980) referred to this process of
temperature and hence depth dependent metasomatism
as 'hydrothermal metamorphism', a term that reflects the
combined processes that operated in the subseafloor.
In areas of volcanic and hydrothermal activity, it is probable
that there is a spectrum of alteration between diagenesis and
hydrothermal alteration where these processes operate in
combination. Hydrothermal activity in the depositional basin
would accelerate and intensify the process of diagenesis by
contributing additional fluid and heat, and by promoting
convection (Iijima, 1974; Marsaglia and Tazaki, 1992).
Deep, semi-conformable altered zones are assumed to
be the products of hydrothermal alteration within regional
subseafloor hydrothermal systems (Gibson et al., 1983;
Galley, 1993). These hydrothermal systems involve the large-
scale convection of modified seawater through the permeable
volcanic successions (Spooner and Fyfe, 1973; Galley, 1993).
The distribution of altered zones and spatial association with
subsurface intrusions suggests that subsurface intrusions,
augmented by heat from the cooling volcanic succession, may
be the driving force for hydrothermal convection (Campbell
et al., 1981; Lesher et al., 1986; Cathles, 1993; Galley, 1993;
Skirrow and Franklin, 1994). Where the upper contacts of
subsurface intrusions are sub-parallel to the volcanic-strata,
the overlying isotherms are also semi-conformable with the
strata and progressive temperature-dependent seawater-rock
FIGURE 6 .10 j The relationships between fluid convection, diagenesis, metamorphism and regional hydrothermal alteration in submarine volcanic successions that
host VHMS deposits. (A) Diagenetic-metamorphic system, where there is a progression from diagenesis to isochemical metamorphism with increasing temperature
and depth in the subseafloor. This transition reflects the maximum depth to which seawater circulates and reacts with the host rocks. (B) Diagenetic-hydrothermal
system, where a subsurface intrusion promotes deep circulation of fluid via the recharge, reservoir and discharge zones. The depth progression from diagenesis to
regional hydrothermal alteration (deep, semi-conformable alteration) is dependent on temperature and fluid circulation.
1 4 8 | CHAPTER 6
FIGURE 6.11 | Alkali ratios for altered andesite samples from the Amulet
rhyolite, Noranda district, Canada (after Gibson et al., 1983). Fields for the
primary and metasomatised (albite-altered) andesites and basalts are from
Hughes (1973).
reactions form a series of semi-conformable altered zones
(Galley, 1993).
In axial mid ocean ridge hydrothermal systems, down-
welling seawaters traverse extremely steep temperature
gradients in the upper crust, from less than 50C near the
seafloor to more than 250C at 1-2 km depth (Mottl, 1983).
Thus, vertically stacked deep, semi-conformable altered
zones result from metasomatic reactions that take place at
progressively higher temperatures with depth in the succession
(Galley, 1993). The decrease in pervasiveness of alteration,
from widespread nearly uniform diagenesis to more restricted
and patchy silicification and Ca-Fe metasomatism, may reflect
decreasing permeability with depth.
The distribution, relative timing, and spatial association
with VHMS deposits suggest a genetic link between regional
hydrothermal alteration and mineralisation. Some authors
have proposed that the deep, semi-conformable altered zones
acted as reservoirs from which metals and sulfur were leached
(e.g. Gibson et al., 1983; Lagerbald and Gorbatschev, 1985;
Galley, 1993; Skirrow and Franklin, 1994). As such they
represent much larger exploration targets than the discordant
altered footwall zones (Galley, 1993).
6.3 | ALTERED ZONES WITHIN
INTRUSIONS
Intrusions are commonly modified by deuteric or hydrothermal
alteration associated with emplacement and may subsequently
undergo diagenesis, regional metamorphism or hydrothermal
alteration.
Deuteric alteration
Deuteric alteration, also referred to as autohydration or
autometamorphism, is the alteration of recently crystallised
magma by trapped magmatic fluid exsolved from the same
cooling magma (Honnorez et al., 1979; Destrigneville et al.,
1991). It has been recorded in intrusions and lavas in both
subaerial and submarine successions (Honnorez et al., 1979;
Bohlke et al., 1980; McConnell et al., 1995). Sederholm
(1929) originally defined deuteric alteration as the alteration
that takes place 'in direct continuation of the consolidation
of the magma' and thus it is considered a magmatic alteration
process. It is the earliest alteration style and is a short-
lived process, typically occurring as intrusions cool from
temperatures of several hundred degrees centigrade (Ade-Hall
etal., 1968; Honnorez etal., 1979). Small volume synvolcanic
intrusions typically cool too rapidly to experience deuteric
alteration (cf. Gromme et al., 1969). In contrast, granitoids
and large-volume sills may have altered zones in their upper
parts resulting from reactions between rising magmatic fluids
and the cooling intrusions (e.g. Fig. 6.1A).
Deuteric textural changes are minimal (Wilshire, 1959).
Phenocrysts, particularly feldspars and mafic minerals, such as
pyroxene or olivine, may be pseudomorphed by amphibole,
chlorite or smectite (Fuller, 1938; Bohlke et al., 1980;
Destrigneville et al., 1991). Open spaces, such as vesicles,
mariolitic voids, and quench fractures, are lined or filled
with smectite, zeolites, carbonate, biotite, chlorite and oxides
(Wilshire, 1959; Furbish and Schrader, 1980; Destrigneville
et al., 1991). High-Ti minerals, such as titanomagnetite, are
oxidised and altered to low-Ti minerals, such as ilmenite
hematite (Butler and Burbank, 1929; Ade-Hall et al., 1968;
Surdam, 1968; Sherwood, 1988).
Deuteric alteration does not involve major chemical
composition changes; some components may be locally
redistributed at sub-millimetre scales or undergo oxidation
state changes (e.g. Fe
2+
/Fe
3+
ratio, Scott and Hajash, 1976)
that may alter rock thermomagnetic properties (Ade-Hall et
al., 1968; Sherwood, 1988). The changes are quantitatively
unimportant when compared to the products of long-lived
diagenesis and hydrothermal alteration and may be difficult
to distinguish from those of other alteration styles.
Hydrothermal alteration
Altered zones within synvolcanic intrusions may also result
from reactions with seawater or modified seawater circulating
through the intrusion, either during the prograde high-
temperature stage of hydrothermal activity or during cooling.
If the hydrostatic pressure is high enough (at sufficient depths)
seawater will be forced into thermal contraction fractures in
the cooling intrusion (Burnham, 1979). The time interval for
seawater-intrusion interaction may be limited by the rapid
development of a local intrusion-related hydrothermal system
in the host succession, which would result in the lithification
and filling of primary pore space inhibiting fluid flow.
Thereafter, episodic seawater-intrusion interaction would
occur only if the fluid pressure exceeds the tensile strength
causing the altered rock adjacent to the contact to fracture
(Secor, 1965, in Fournier 1985; Phillips, 1973; Henley and
McNabb, 1978). Alteration follows the advancing front
of brittle fracturing to deeper and deeper levels within the
intrusion (Burnham, 1979; Giggenbach, 1997).
The resulting altered zones may be pervasive, occur along
cooling fronts or more commonly as selective-pervasive
alteration adjacent to fractures or veins. Alteration minerals
fill vesicles, miarolitic voids and fractures, cement hydraulic
breccias, and pseudomorph primary magmatic minerals
(Mevel and Cannat, 1991; Gillis et al., 1993; Kelley and
Gillis, 1993; Nehlig, 1993; Davidson, 1998; Galley, 2003).
Polya et al. (1986) and Davidson (1998) described
proximal zones of hydrothermal alteration within the
Cambrian Murchison Granite, western Tasmania. They
described narrow, texturally destructive intense K-feldspar
zones, associated with calcite veins, irregularly distributed
in the margins of the granite and patchy selective-pervasive
chlorite zones with chlorite pseudomorphs after biotite and
hornblende or patches of chlorite pyrite sphene.
The significant mineralogical, compositional and
isotopic changes associated with proximal hydrothermal alter-
ation within intrusions are consistent with seawater-rock
interaction (Gregory and Taylor, 1981; Stakes and Taylor,
1992; Cathles, 1993; Galley, 2003). Galley (2003) identified
three types of early hydrothermal and magmatic alteration
within subvolcanic intrusions in the Snow Lake, Noranda and
Sturgeon Lake districts. The earliest, a greenschist alteration
facies in quartz diorite intrusions, is manifest as pervasive
epidote + quartz, and epidote + actinolite + quartz + albite
+ magnetite sulfides, which replaced primary minerals and
infilled miarolitic cavities, vesicles and fractures. Mass change
calculations suggest CaO, Sr, Pb and CO
2
were gained and
Fe
2
O
3
, MgO, Cu, Zn, Mo, Na
2
O, K
2
O and Ba lost. Galley
(2003) interpreted this facies to be the product of high-
temperature hydrothermal-magmatic alteration resulting from
emplacement of quartz-diorite intrusions into a seawater-
saturated succession. The second chlorite-rich alteration facies
is characterised by quartz + chlorite + sericite and chlorite
+ sulfide-filled fractures and vein selvages. It is most intense
near the margins of intrusions and directly beneath VHMS
deposits. The chlorite-rich zones gained Fe
2
O
3
, MgO, Cu
Pb, K
2
O, Ba, and Zn, and lost CaO, Na
2
O, Sr SiO
2
, Ba
and CO
2
and are interpreted to result from hydrothermal
alteration (Galley, 2003). Overprinting both of these zones
is a biotite-rich alteration facies associated with silicification,
and Cu-Mo-rich veins and breccia, which is interpreted as
a magmatic alteration facies associated with late stage dykes
(Galley, 2003).
Hydrothermal alteration may also result from the
absorption of fluid from and assimilation with sediment
inclusions incorporated into the magma as it was emplaced
into wet unconsolidated sediment. Wilshire and Hobbs
(1962) described hydrothermal alteration in the margin of
a peperitic latite intrusion in a submarine volcaniclastic
succession, near Port Kembla in New South Wales. Alkali
feldspar + chlorite + carbonate-rich zones coincide with
abundant sediment inclusions and quench fractures in the
margin of the intrusion, and the sedimentary inclusions have
been chlorite zeolites carbonate altered. The altered latite
gained Na
2
O and volatiles, and lost SiO
2
, A1
2
O
3
, Fe
2
O
3
, K
2
O,
MgO and CaO, whereas the sediment inclusions lost Na
2
O
and volatiles, and gained CaO and MgO K
2
O.
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 4 9
6.4 | CONTACT ALTERED HALOS
AROUND INTRUSIONS
All magmatic intrusions transfer heat; they have thermal
impacts on the enclosing rocks or sediments and may induce
compositional changes. Contact alteration is used here as a
non-genetic term referring collectively to the processes of
contact metamorphism and contact hydrothermal alteration.
Contact or thermal metamorphism involves changes in
rock texture and mineralogy of the immediate host rock as a
result of temperature increase (Yardley, 1989). The increased
temperature drives dehydration and decarbonation reactions,
and fluid migration away from the intrusion (Blatt et al.,
1972; Manning and Bird, 1991). Only small volumes of
H
2
O- and CO
2
-rich fluids are generated from these reactions
and thus the metasomatic effect of contact metamorphism
is negligible (Rose and Burt, 1979). Contact metamorphism
typically results in only local remobilisation but extensive
static recrystallisation of existing minerals or components.
Contact hydrothermal alteration involves a substantial
volume of heated fluid, typically comprising trapped seawater
and pore fluid with or without magmatic fluid derived from
the intrusion, which circulates through and reacts with the
host facies (MacGeehan, 1978; Taylor and Forester, 1979;
Polya et al., 1986; Galley, 2003). This promotes textural,
mineralogical and compositional changes in the host facies.
In submarine volcanic successions, abundant trapped
seawater means that isochemical thermal metamorphism is
rare. In addition, interaction between hot magma and wet
unconsolidated sediment can result in: peperitic contacts,
fluidisation of sediment (e.g. Schmincke, 1967; Kokelaar,
1982), fluid expulsion, induration (e.g. Einsele et al., 1980),
secondary welding (e.g. Ito et al., 1984, in Kano, 1989),
brecciation of host rock, local or regional hydrothermal
alteration, quenching of the intrusion and magma-host rock
assimilation (e.g. Wilshire and Hobbs, 1962; Puffer and
Benimoff, 1997; WoldeGabriel et al., 1999).
Contact altered zones
Contact altered zones are spatially associated with intrusion
margins and may surround the intrusion as halos or aureoles.
Successive contact altered zones reflect progressive changes
in temperature or temperature and chemical conditions in
the host succession away from the intrusion (Rose and Burt,
1979; Einsele et al., 1980; Yardley, 1989).
Contact altered halos may vary in thickness from a few
millimetres at the margins of thin, shallow synvolcanic sills
(e.g. Einsele, 1985; Boulter, 1993; Skirrow and Franklin,
1994) to several kilometres around large subvolcanic plutons
or intrusive complexes (e.g. Boulter, 1993; Schweitzer and
Hatton, 1995; Galley, 2003). They may comprise one low-
grade altered zone, a sequence of roughly concentric altered
zones, a series of asymmetric altered zones or overprinting
altered zones (Fig. 6.12).
Grades and mineral assemblages of contact altered zones
vary considerably, reflecting: temperature and compositional
differences between the host succession, the intrusion, and any
fluid; duration of the alteration system; emplacement depth;
1 5 0 | CHAPTER 6
FIGURE 6.12 | Cartoons of the variety of contact altered zones around
intrusions. (A) Cross-section showing two sills that were emplaced into wet
unconsolidated sediment at shallow levels beneath the seafloor. The sills both
have single low-grade indurated zones, which have a lower porosity than the
surrounding host turbidites (after Einsele et al., 1980). (B) Schematic cross-
section through roughly concentric zeolite and clay zones around a granite
emplaced into felsic volcanic fades in the Green Tuff Belt, Japan. These zones
are, from the intrusion outwards: a zeolite zone, devitrified zone, and least-
altered zone (after Utada, 1991). (C) Schematic section of the asymmetric
altered halo around the Rustenberg Layered Suite intrusions in the Rooiberg
felsic volcanic rocks of the Bushveld Complex, Africa (after Schweitzer and
Hatton, 1995). Above the intrusion is a thick (>1.4 km) halo comprising biotite
hornfeis and quartz + sericite + albite zones, which are enriched in K
2
O, MgO
and base metals. Beneath the intrusion is a thinner (<400 m) granoblastic zone,
in which primary volcanic textures are overprinted by metamorphic textures
without significant compositional changes. Schweitzer and Hatton (1995)
postulated that the reason for the asymmetrical zonation was that heated fluid
convected freely above the intrusion and hydrothermal alteration dominated,
whereas buoyant convection was inhibited beneath the intrusion acting as a seal,
and thermal metamorphism dominated. (D) Map view of the prograde olivine,
pyroxene and actinolite + chlorite zones associated with emplacement of the
Skaergaard intrusion into mafic volcanic rocks, east Greenland (after Manning
and Bird, 1991,1995). In the outer pyroxene zone and adjacent to fractures
in the pyroxene and olivine zones, high-temperature mineral assemblages
are overprinted by actinolite + chlorite, suggesting retrograde metamorphism
occurred as temperatures dropped and cooler hydrothermal fluids migrated
inwards through fractures.
rate of cooling; and the subsequent regional metamorphic
grade. Although it is difficult to generalise about the
mineralogy of contact altered zones, there are a few indicator
minerals that are almost exclusively generated by intrusion-
related hydrothermal alteration (i.e. minerals associated with
magmatic systems such as biotite, diaspore, fluorite, kaolinite,
magnetite, pyrophyllite, rutile, topaz and tourmaline).
Typically altered halos associated with small-volume
intrusions emplaced at shallow depths below the seafloor
comprise low-grade altered zones adjacent to the intrusion,
which grade into partially altered zones at the peripheries.
Low-grade altered zones may be manifest as indurated
sediment (e.g. Einsele et al., 1980; Kano, 1989), fused glass
(e.g. Ross and Smith, 1960; Smith, I960; Christiansen and
Lipman, 1966; Schmincke, 1967; McPhie and Hunns,
1995), devitrified glass (e.g. Schweitzer and Hatton, 1995;
WoldeGabriel et al., 1999), palagonite or clay minerals in
mafic volcanic rocks (e.g. Upton and Wadsworth, 1970;
Jakobsson, 1972; 1978) or zeolite and clay minerals in felsic
volcanic rocks (e.g. Iijima, 1978; Utada, 1991).
In contrast, high-grade altered zones tend to be associated
with large volume subvolcanic intrusions and include high-
temperature (up to 1000C) mineral assemblages. For
example, Seki et al. (1969) reported five high-grade altered
zones around a large intrusion in the Neogene Green Tuff Belt,
Japan. From the contact to the margin they were: amphibole
zone, actinolite zone, pumpellyite + prehnite + chlorite zone,
laumonite + mixed-layer chlorite zone, and clinoptilolite +
vermiculite zone.
SYNV OLCANIC INTRUSION-RELATED ALTERATION I 1 5 1
Indurated or fused zones
Thin contact altered zones of fused or secondary welded
volcanic glass are common adjacent to intrusions and lavas
in subaerial volcanic successions (e.g. Ross and Smith, I960;
Smith, 1960; Christiansen and Lipman, 1966; Schmincke,
1967; WoldeGabriel et al., 1999), and also occur around
intrusions in ancient submarine volcanic successions (e.g. Ito
etal., 1984, in Kano, 1989; McPhie and Hunns, 1995).
Christiansen and Lipman (1966) used the term fused for
the induration and deformation of glassy clasts resulting from
heating by adjacent lava, but emphasised that the term should
not be taken to imply that melting (fusion) had occurred. They
described altered subaerial tuffs adjacent to the Combs Peak
rhyolite lavas and domes near Fortymile Canyon, southern
Nevada. Three altered zones were developed parallel to the
lava contact: an outer red zone characterised by the oxidation
of glass, a middle indurated or partially fused zone and an
inner densely fused zone characterised by fiamme and eutaxitic
texture (Figs 6.13 and 6.14). In this case, eutaxitic texture was
interpreted to result from the re-heating and accompanying
load compaction of originally glassy pumice clasts in tuffs
beneath the lava as a result of its emplacement (Christiansen
and Lipman, 1966). The minimum temperature required for
this partial welding is 535C (Smith, 1960).
Typically, indurated or fused zones closely parallel lava
or intrusion contacts and may be several millimetres to
tens of metres thick (e.g. Christiansen and Lipman, 1966;
Einsele, 1985; Keating and Geissman, 1998). They are
commonly associated with thin (< 1100 m) intrusions that
have peperitic or irregular contacts indicating emplacement
into wet unconsolidated sediments (Kokelaar, 1982; Branney
and Suthren, 1988; McPhie and Hunns, 1995; Keating and
Geissman, 1998). Induration of sediment adjacent to contacts
and around juvenile clasts in peperite is typically accompanied
by changes in colour associated with thin (cm scale) carbonate,
quartz or Fe-oxide altered halos (Fig. 6.15A: Schmincke,
1967; Kokelaar, 1982; Kano, 1989; Hunns and McPhie,
1999). The most significant textural changes in this zone are
contact-parallel fiamme and eutaxitic textures in pumice-rich
facies (Fig. 6.15B: McPhie and Hunns, 1995).
Devitrified zones
Devitrified zones are characterised by high-temperature
devitrification textures such as spherulites, lithophysae and
micropoikilitic texture (Christiansen and Lipman, 1966;
McPhie and Hunns, 1995). It is important to note that
devitrification textures generated from re-heating of glassy
volcanic facies by intrusions are indistinguishable from those
formed during first cooling (Lofgren, 1971a, 1971b).
Narrow zones oriented parallel to intrusion contacts
may be completely or partially devitrified (e.g. Keating and
Geissman, 1998; WoldeGabriel et al., 1999). They may
overprint fused zones. For example, Christiansen and Lipman
(1966) described superposition of three devitrified zones on to
three fused zones in bedded rhyolitic tuffs (Fig. 6.14): an outer
porous glassy zone, a middle dense glassy zone (vitrophyre
that is commonly perlitic), and an inner crystalline zone with
microlites, spherulites and lithophysae.
FIGURE 6.13 | Distribution of the fused zone adjacent to the Combs Peak
rhyolite, near Fortymile Canyon, southern Nevada (modified after Christiansen
and Lipman, 196 6 ).
FIGURE 6.14 | Idealised relationships between the Combs Peak rhyolite
(Nevada), the three fused zones and overprinting devitrified zones (modified
after Christiansen and Lipman, 196 6 ).
Compositional changes associated with devitrification
are usually negligible. WoldeGabriel et al. (1999) found that
devitrification in felsic volcaniclastic rocks in a 10m thick
contact zone around a basaltic intrusion at Grants Ridge,
New Mexico, was associated with minor gains in K
2
O and
losses in H
2
O, Na
2
O, F, Fe
2
O
3
. The margins of the intrusion
were slightly enriched in SiO
2
, K
2
O and P
2
O
5
and depleted
in Fe
2
O
3
. They concluded that the thermal effects of the
intrusion induced devitrification, dehydration and vapour-
phase expulsion in the contact zone. Vapour-phase expulsion
of fluoride, chloride, hydroxide, sulfide, and CO
2
from silicic
glass may have been responsible for the subtle chemical
variations during devitrification (cf. Weaver et al., 1990).
Zeolite, clay or palagonite zones
Low-temperature altered zones characterised by palagonite,
zeolite and clay minerals in mafic volcanic rocks (e.g. Upton
and Wadsworth, 1970), and zeolite and clay minerals in felsic
volcanic rocks (Utada, 1991) are common around shallow
synvolcanic intrusions in submarine volcanic successions.
Mineral assemblages in these zones typically reflect the
host rock compositions. For example, altered zones around
granitoids in the felsic volcanic rocks of the Green Tuff
Belt contain calcic zeolites (Iijima, 1978; Utada, 1991). In
contrast, palagonite dominates altered zones around dykes in
1 5 2 | CHAPTER 6
A. Indurated siltstone in peperite
The irregular clasts of indurated and silicified siltstone
(grey) are mixed with feldspar-phyric rhyolite (green)
clasts in this peperitic contact between rhyolite and
siltstone. Away from the rhyolite contact, the host
siltstone is green-grey, but fades to cream or pale
green silicified siltstone in a zone about 1-2 cm wide
adjacent to the rhyolite clasts in the peperite. This local
colour change and silicification result from the thermal
metamorphism of the unconsolidated silt in contact
with hot rhyolite.
Early Permian Berserker beds, Mount Chalmers district,
Queensland.
B. Fused pumice breccia
Well-developed fiamme (F) and eutaxitic texture
characterise the fused zone in this pumice breccia
immediately adjacent to a rhyolitic sill. Away from the
rhyolite, fiamme in the pumice breccia are indistinct
and parallel to bedding, whereas in the fused zone
they parallel the pumice breccia-rhyolite contact. The
fiamme and eutaxitic texture result from the partial
welding and compaction of glassy pumice clasts during
heating associated with emplacement of the rhyolite
(McPhie and Hunns, 1995).
Early Permian Berserker beds, Mount Chalmers district,
Queensland.
Figure 6.15 | Photographs of hand specimens from the indurated and fused zones adjacent to rhyolite sills near the IVIount Chalmers VHMS deposit, Queensland.
submarine basaltic hyaloclastite at Surtsey (Jakobsson, 1972,
1978; Jakobsson and Moore, 1986), and chabazite, analcime,
thomsonite, mesolite, phillipsite and natrolite characterise the
zeolite zone associated with a swarm of sills in basaltic lavas
and breccias at Piton des Neiges volcano on Reunion Island
(Lacroix, 1936; Upton and Wadsworth, 1970). Zeolites
in these zones may be accompanied by chlorite, epidote,
carbonate and clay minerals.
Compositional changes in the zeolite, clay or palagonite
zones include K
2
O and MgO gains, and SiO
2
and CaO losses
(Hart, 1969; Thompson, 1973; Honnorezetal., 1979). These
are consistent with low-temperature (<150C) reactions
with seawater promoting oxidation, fixation of alkalis, and
exchange of seawater-Mg for rock-Ca to form smectite (Alt,
1999).
Greenschist facies zones
Synvolcanic granitoids, large composite intrusions and
clusters of sills or dykes commonly have altered zones with
epidote-, chlorite-, sericite-, biotite- or K-feldspar-bearing
mineral assemblages characteristic of greenschist facies meta-
morphism (Polya et al., 1986; Boulter, 1993; Neuhoff et al.,
1997; Galley, 2003).
For example, Schweitzer and Hatton (1995) described
a 1.4 km thick greenschist facies aureole above the mafic
Rustenburg Layered Suite in the felsic Rooiberg volcanic
rocks of the Bushveld Complex (Fig. 6.12C). The asymmetric
aureole contains a biotite hornfels zone (immediately above
the Rustenburg Layered Suite), and an overlying quartz
+ sericite + albite zone, which grades up into least-altered,
devitrified volcanic rocks. In the quartz + sericite + albite
zone, hornblende or chlorite replaced mafic phenocrysts, and
quartz + chlorite + epidote replaced the glassy groundmass.
Primary compositions may be significantly modified
in greenschist facies zones. They are commonly enriched
in K
2
O and MgO and depleted in CaO, Fe
2
O
3
, Na
2
O and
MnO (Schweitzer and Hatton, 1995; Large et al., 1996).
The behaviour of SiO
2
is variable. The mineralogical and
compositional changes reflect high-temperature (300-450C)
seawater-rock interactions similar to some proximal altered
zones associated with VHMS ore deposits (Galley, 2003).
Silicified zones
Silicified zones are typically pale grey in colour and can be
massive pervasive or patchy in texture, filling vesicles and
fractures, or cementing breccias (Humphris and Thompson,
1978; Skirrow and Franklin, 1994; Gifkins and Allen, 2001).
They comprise chalcedony, cristobalite, quartz, quartz +
feldspar, or quartz + sericite dominated alteration mineral
assemblages.
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 5 3
Skirrow and Franklin (1994) described 10 cm to 2 m
thick silicified contact aureoles associated with unaltered
plagioclase- and quartz + plagioclase-phyric porphyry dykes
in the submarine volcanic rocks beneath the Chisel Lake
VHMS deposit in the Snow Lake district. The weakly silicified
mottled zones consist of irregular light grey patches of quartz
+ plagioclase + hornblende + magnetite biotite, which
coalesce into massive quartz + plagioclase rock in intensely
silicified zones.
Compositional changes include gains in total mass and
SiO
2
, which may be accompanied by gains in K
2
O or Na
2
O,
and losses in Fe
2
O
3
, MgO, CaO and Zn (MacGeehan, 1978;
Skirrow and Franklin, 1994; Gifkins and Allen, 2001).
Silicification is a common feature of hydrothermal
alteration and incorporates both the addition of Si (largely
as vein infill) and the redistribution of Si that was originally
in glass or cristobalite (Henley and Ellis, 1983). Circulating
heated seawater can leach Si from the intrusion or felsic
volcanic glass in the host succession, resulting in a solution
supersaturated with Si. Silica precipitation from this solution
can occur by several mechanisms: cooling by conduction or
mixing, decompression associated with boiling, heating into
the temperature range for retrograde Si solubility, or a pH
change (Dickson and Potter, 1982; Fournier, 1985). The
solubility of Si generally increases with increasing temperature
(Fig. 6.9); however, if a supersaturated solution is heated at
constant pressure (<900 bars) it will either boil or reach a
solubility maximum and may precipitate quartz upon further
heating (Fournier, 1985). A supersaturated saline fluid may
precipitate quartz at temperatures between 300 and 55OC
(Fournier, 1985). Thus Si-saturated seawater would deposit
quartz on encountering temperatures greater than 300C
in the intrusion or the immediate host rocks adjacent to
the intrusion. MacGeehan (1978) proposed this process,
of Si leaching from volcanic glass and heating of the fluid
into the retrograde solubility temperature range, to explain
silicification in pillow basalts adjacent to synvolcanic gabbro
sills in the Matagami district.
Genesis of contact altered zones
Contact altered zones may develop adjacent to an intrusion
as heat is transferred from the cooling intrusion and heated
modified seawater reacts with the host succession (Fig. 6.16).
Vapour or fluid exsolved from the crystallising magma may
contribute both heat and elements to the hydrothermal fluid
(Norton, 1984). Hydrothermal fluid temperatures are partly
determined by the depth of emplacement, volume of the
intrusion and the temperature and volume of contributed
magmatic fluid (Polyaetal., 1986; Eastoeetal., 1987; Cathles,
1993; Galley, 2003). For example, two active hydrothermal
systems are recognised in the Guaymas Basin (Geiskes et al.,
1982; Kastner, 1982). One is a low temperature (<300C)
hydrothermal system associated with the emplacement
of shallow sills into unconsolidated sediments below the
seafloor. The other is a deep high-temperature hydrothermal
system associated with dykes or magma chambers that fed the
overlying sill complexes.
FIGURE 6,16 | Development of a contact metamorphic-hydrothermal system
in a submarine volcanic succession after the emplacement of a synvolcanic
intrusion. (A) Initial fluid expulsion and migration away from the intrusion as
heat from the intrusion drives dehydration and decarbonation reactions in the
host succession. A combination of thermal metamorphism and hydrothermal
alteration, by seawater and magmatic volatiles and fluid, may produce a contact
altered zone. Seawater heated by the intrusion is buoyant and rises towards the
seafloor either by diffuse flow or along fractures and faults. (B) In response, cold
seawater is drawn down and heated in the vicinity of the intrusion, promoting
hydrothermal convection and alteration between the intrusion and the seafloor.
(C) Hydrothermal convection collapses as the intrusion cools. Cold seawater
may be drawn down along fractures to produce proximal zones of hydrothermal
alteration within the intrusion, and retrograde zones that overprint higher
temperature contact altered zones adjacent to the intrusion.
The volume of an intrusion influences the temperature and
longevity of the alteration system. Large volume intrusions
(e.g. plutons and thick sills) influence the temperature of the
host rocks and the alteration system for longer than smaller
volume intrusions (e.g. synvolcanic sills, cryptodomes and
dykes). A small volume sill (-30 m thick) may cause the
temperature at the sill-sediment contact to rise as high as
400C (Einsele et al., 1980). However, calculations suggest
1 5 4 | CHAPTER 6
that the temperature at the contact will drop below boiling
within five years. This will significantly reduce convection
and remaining heat will be lost mainly by conduction through
the contact. Generally relatively small volume intrusions
have thermal effects restricted to several metres or hundreds
of metres from the contacts (Utada, 1973). Although the
contact temperatures may be high, high-temperature altered
zones are rare because the isotherms dip sharply away from
small volume intrusions (Reyes, 1990).
In contrast, large volume intrusions, such as the Skaergaard
intrusion in east Greenland, which had an estimated volume
of 180 km
3
, may take 500,000 years to cool to ambient
temperatures (Norton and Taylor, 1979; Norton, 1984).
They may result in thick, high-grade contact metamorphic-
hydrothermal altered zones (Seki et al., 1969) and may drive
regional convection of modified seawater.
6.5 | CONTACT ALTERED ZONES
ASSOCIATED WITH THE DARWIN
GRANITE
Cambrian granites along the eastern margin of the Mount
Read Volcanics (Fig. 1.5) are extensively altered and
surrounded by concentric altered zones (Polya, 1981; Polya et
al., 1986;Eastoeetal., 1987; Abbott, 1992; Large etal., 1996;
Davidson, 1998; Wyman, 2001). Well-developed K-feldspar,
chlorite and sericite zones have been mapped around the
margin of the Darwin Granite and its northward extension in
the southern Mount Read Volcanics (Fig. 6.17: Jones, 1993;
Large et al., 1996; Wyman, 2001).
The Darwin Granite alteration halo is of particular
interest because of its close spatial and possibly temporal
relationships with several small tonnage but high-grade Cu-
Au prospects (Fig. 6.17: Jones, 1993; Large et al., 1996).
These prospects occur in the Central Volcanic Complex along
the western margin of the granite and above its northern
subsurface projection from Mount Darwin towards Mount
Lyell. The deposit styles vary systematically with increasing
distance from the granite: from Fe-oxide veins, stockworks
of pyrite + chalcopyrite hematite magnetite and quartz +
pyrite + chalcopyrite veins, disseminated pyrite + chalcopyrite
covellite, to veins containing quartz, bornite, chalcopyrite
and hematite (Wyman, 2001). Large et al. (1996) suggested
that the Darwin Granite provided heat, metals and magmatic
fluid to form VHMS deposits in the southern Mount Read
Volcanics, such as those in the Mount Lyell field.
This section summarises the setting, altered zones and
genesis of the Darwin Granite system and presents data sheets
of typical alteration facies (DG1 to DG6).
FIGURE 6.17 | Geological map of the Jukes-Darwin area in the southern Mount
Read Volcanics (western Tasmania), showing the limited surface extent of the
Darwin granite and the thick hydrothermally altered halo around the granite
(modified after Wyman, 2001). The locations of the six data sheets are shown.
Geological setting
The Darwin Granite is an I-type magnetite series equigranular
granitoid pluton dominated by pink granite intruded by
subordinate white granite, microgranite and quartz porphyry
phases (Wyman, 2001). The surface extent of the pluton is
approximately 5 x 1 km (Fig. 6.17). However, modelling
of gravity and aeromagnetic data along the eastern margin
of the Mount Read Volcanics has suggested that a semi-
continuous body of granite extends subsurface approximately
100 km northwards to the Murchison Gorge (Leaman and
Richardson, 1989; Payne, 1991; Large et al., 1996).
In the Darwin-Jukes area, the Central Volcanic Complex
includes feldspar-phyric dacite, quartz + feldspar-phyric
rhyolite (e.g. data sheet DG1), pumice breccia, tuffaceous
sandstone, blocky rhyolite breccia, and minor sedimentary
facies (Jones, 1993; Wyman, 2001). A thick, columnar
jointed, micropoikilitic or spherulitic rhyolite hosts the
Jukes Cu-Au Prospect and altered zones. The emplacement
age of Darwin Granite is constrained to the Cambrian as it
intruded the Middle Cambrian Central Volcanic Complex,
and both the granite and Central Volcanic Complex are
unconformably overlain by late Middle Cambrian Tyndall
Group volcaniclastic rocks, which contain pebbles of granite
near Mount Darwin (Corbett, 1979, 1981, 1992; Jones,
1993; Wyman, 2001).
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 5 5
Alt er at i on facies and zonat i on
The altered zones associated with the Darwin Granite cover a
15x3 km area that extends north to Jukes Prospect (Wyman,
2001). The altered zones at Mount Darwin and Jukes Prospect
represent two parts of the same hydrothermal system: those
adjacent to the granite at Mount Darwin record the lateral
extent of the altered halo, whereas those at Jukes Prospect
occur at least 1 km above the north-plunging pluton (Fig.
6.18).
Immediately adjacent to the granite at Mount Darwin is
a thin (1020 m) K-feldspar biotite (now chlorite) hornfels
zone (Table 6.1). This grades outward into a 400 m wide
K-feldspar + chlorite zone, which locally hosts coarse
breccias with magnetite tourmaline matrices (Jones, 1993;
Wyman, 2001). The K-feldspar + chlorite zone grades into
a 300 m thick chlorite + magnetite zone (Wyman, 2001).
A discontinuous wedge-shaped silicified zone separates
the chlorite + magnetite zone from the outer K-feldspar +
quartz zone. At Mount Darwin the K-feldspar + quartz zone
occurs between 800 and 1000 m from the granite contact.
At Jukes Prospect the K-feldspar + quartz zone is the central
altered zone and is enclosed in chlorite, sericite and regional
diagenetic-metamorphic zones. It is associated with Cu-Au
mineralised rocks (Doyle, 1990; Large et al., 1996). The
peripheral sericite zone merges with the regional diagenetic-
metamorphic albite + sericite zone (Wyman, 2001).
K-feldspar <20 K-feldspar chlorite Pervasive, hornfels
biotite hornfels (after biotite) quartz
sulfides
K-feldspar + 400 K-feldspar + quartz Intense Magnetite tourmaline fill in SiO
2
, AI
2
O
3
, K
2
O and net DG2
chlorite + chlorite sericite hydraulic breccias and veins mass gains
hematite (after
magnetite) Na
2
O losses
Silicified ' Quartz sericite pyrite Strong to Texturally destructive, Large gains in SiO
2
and DG3
hematite intense cryptocrystalline and net mass
microcrystalline
K-feldspar + 200 K-feldspar + quartz Strong to Moderately texturally K
2
O, SiO
2
and Fe
2
O
3
DG4
quartz + sericite + chlorite + intense destructive, pervasive, gains
pyrite magnetite cryptocrystalline ,
chalcopyrite pseudomorphs plagioclase, Na
2
O losses
veins and vein envelope
Chlorite + 300 Chlorite + sericite + Moderate Moderate preservation, chlorite Fe
2
O
3
, MgO and K
2
O DG5
magnetite magnetite dolomite to intense pseudomorphs plagioclase, gains
apatite pyrite, domainal replacement of
chalcopyrite veins groundmass and matrix SiO
2
, AI
2
O
3
, Na
2
O and
textures, infill in hydraulic net mass losses
breccias and veins, vein
envelope
Sericite Sericite chlorite Weak to Sericite partially to completely K
2
O gain and Na
2
O DG6
pyrite strong replaces plagioclase and
is disseminated in the CaO losses
groundmass and matrix
1 5 6 | CHAPTER 6
FIGURE 6.18 | Schematic cross-section of the Darwin Granite and altered
zones, illustrating relationships between the Mount Darwin and Jukes Prospect
alteration systems, and surface maps presented in the data sheets (Large et al.,
1996 ; Wyman, 2001).
Genesis of the alteration system
The proximity of intense hydrothermal alteration mineral
assemblages to the granite contact and the northward
extending cupola region above the buried pluton support
the interpretation that the hydrothermal system was driven
by heat from the intrusion (Eastoe et al., 1987; Large et al.,
1996; Wyman, 2001). Overprinting relationships between K-
feldspar, sericite and chlorite facies indicate multiple alteration
stages in which low-temperature mineral assemblages
overprinted initial high-temperature assemblages (Wyman,
2001).
Initial sericite and chlorite alteration assemblages were
associated with fracturing and vein formation around and
above the granite. As the fracture system evolved, the Jukes
hydrothermal system became part of the discharge zone. The
well-defined zones of hydrothermal alteration are interpreted
to have formed from diffuse circulation of hydrothermal fluid
through the volcanic rocks. Fluid access was enhanced by
hydrothermal brecciation, and intense K-feldspar alteration
and silicification was confined to the fracture zone above the
granite (Wyman, 2001).
Magnetite and tourmaline veins and breccias in the K-
feldspar-rich zones immediately adjacent to the granite
contact, and in the centre of the Jukes alteration system,
demonstrate that magmatic-hydrothermal fluids were exsolved
during crystallisation (Large et al., 1996). The mass changes
(i.e. gains in K
2
O, Fe
2
O
3
, Ba, Sr, Cu, Mo, W, Th and U) in
the altered zones adjacent to the granite are consistent with
hydrothermal alteration of the volcanic rocks by magmatic
fluid (Wyman, 2001). These fluids mixed with modified
seawater in reaction zones around the hotter portions of the
discharge zones to form K- and Fe-rich alteration assemblages
above the granite (Wyman, 2001). The K-feldspar, chlorite
and sericite zones in the Jukes area all show depletions in
Na
2
O and CaO, which reflect the breakdown of plagioclase
and mafic minerals during alteration by modified seawater.
Large et al. (1996) suggested that the distribution,
composition and zonation of alteration facies around the
Darwin Granite, regional zonation of metals with respect to
the granite, distribution of Cu-Au-rich VHMS deposits in the
Mount Lyell field and pre-Tyndall Group timing of both the
granite and mineralisation support a genetic link between the
granite and these deposits. Furthermore, magnetite + apatite
pyrite veins in the Prince Lyell deposit are similar to those
adjacent to and within the Darwin Granite and are consistent
with magmatic fluid contributing to their formation. These
authors proposed a model for the genesis of the Mount Lyell
Cu-Au and related Pb-Zn-Cu massive sulfide deposits (Fig.
6.19), which involves seawater convection deep into the
Central Volcanic Complex where it mixed with Fe, Cu, Au and
P-rich magmatic fluids exsolved from the granite. The mixed
magmatic-seawater hydrothermal system produced altered
zones, magnetite veins (e.g. Jukes Prospect) and subseafloor
Cu-Au deposits (e.g. Prince Lyell) close to the granite, and
contemporaneous Pb-Zn-Cu massive sulfide deposits on the
seafloor (e.g. Lyell-Comstock).
FIGURE 6.19 | Model of the zones of hydrothermal alteration and ore deposits
in the Mount Lyell field and their relationship to the Darwin Granite, southern
Mount Read Volcanics, western Tasmania (after Large et al., 1996 ).
Weak, regional, pervasive albite + sericite alteration facies
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 5 7
DG1
Hand specimen photograph Photomicrograph (ppl)
Sample no. 143401
Alteration fades weak, regional, pervasive albite + sericite
Location Jukes Road
Formation Central V olcanic Complex
Succession Mount Read V olcanics
V olcanic facies massive, feldspar + quartz-phyric rhyolite
Relict minerals feldspar (3%, 2 mm), quartz ( 1 %,
<1 mm)
Relict textures micropoikilitic, porphyritic
Primary composition rhyolite
Lithofacies columnar jointed, massive
Interpretation sill
Alteration minerals albite + sericite + chlorite + hematite
Alteration textures recrystallised groundmass, sericite
pseudomorphs after feldspar, pervasive-
selective sericite > chlorite, disseminated
hematite
Distribution regional
Preservation good
Alteration intensity weak
Timing early
Alteration style regional diagenetic and metamorphic
Geochemistry
SiO
2
76 .17 Na
2
O 2.29 Rb 129 Zr 280
TiO
2
0.31 K
2
O 3.6 6 Sr 22 Nb 13
AI
2
O
3
13.30 P
2
O
5
0.04 Ba 822 Y 39
Fe
2
O
3
3.35 S 0 Cu 4
MnO 0.10 Total 100.00 Pb 20 Al 6 4
MgO o.6 7 (
voLfree
> Zn 39 CCPI 38
CaO 0.11 Th 21 Ti/Zr
6
. 6
1 5 8 | CHAPTER 6
Intense, pervasive K-feldspar + chlorite alteration facies DG2
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
143278
intense, pervasive K-feldspar +
chlorite
Mount Darwin
Central V olcanic Complex
Mount Read V olcanics
massive, feldspar + quartz-phyric
rhyolite
quartz + plagioclase (5 %, 1-2 mm)
porphyritic
rhyolite
massive
unknown
quartz + K-feldspar > chlorite + sericite
> hematite + pyrite
plagioclase replaced by sericite,
recrystallised groundmass of quartz
+ K-feldspar > chlorite + sericite,
disseminated hematite + pyrite,
selective chlorite
local
none
intense
syn- to post-intrusion
proximal intrusion-related
hydrothermal alteration
Hand specimen photograph Photomicrograph (ppl)
Geochemistry ^
2 2 3
SiO
2
71.35 Na
2
O 0.30 Rb 259
N b 1 g
TiO
2
0.27 K
2
O 9.27 Sr 80
y 46
AI
2
O
3
14.59 P
2
O
5
0.04 Ba 246 3
Fe
2
O
3
2.96 S 0 Cu 8
A] g7
MnO 0.01 Total 99.72 Pb 31
c c p | 2J
Mg O 0. 92 (vol. free) Zn 8 9
T i / Z r 7 3
CaO 0.01 Th 25
Strong, foliated quartz + sericite + pyrite alteration facies
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 5 9
DG3
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution/zonation
Preservation
Alteration intensity
Timing
Alteration style
143237
strong, foliated quartz + sericite + pyrite
Slate Spur
Central V olcanic Complex
Mount Read V olcanics
feldspar + quartz-phyric rhyolite schist
quartz + plagioclase (15%, 1-5 mm)
porphyritic, microcrystalline, partly
spherulitic and possibly perlitic
rhyolite
foliated
unknown
quartz + sericite + pyrite + hematite
plagioclase replaced by sericite,
pervasive microcrystalline groundmass
of quartz + sericite + pyrite, hematite
stylolites, quartz overgrowths,
schistosity
local
poor
strong
syn- to post-intrusion
proximal intrusion-related hydrothermal
alteration
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
81.77 Na
2
O 1.54 Rb 145 Zr 142
TiO
2
0.17 K
2
O 4.51 Sr 35 Nb 11
AI
2
O
3
10.23 P
2
O
5
0.03 Ba 889 Y 33
Fe
2
O
3
1.37 S 0 Cu 3
MnO 0.01 Total 99.92 Pb 16 Al 76
MgO 0.29 (vol. free) Zn 22 CCPI 20
CaO 0.01 Th 19 Ti/Zr 7.2
1 6 0 | CHAPTER 6
Intense, pervasive K-feldspar + sericite alteration facies DG4
Hand specimen photograph Photomicrograph (ppl)
Sample no. 14336 0
Alteration fades intense, pervasive K-feldspar + sericite
Location Jukes Road
Formation Central V olcanic Complex
Succession Mount Read V olcanics
V olcanic facies massive, feldspar + quartz-phyric rhyolite
Relict minerals plagioclase (2 %, 2 mm), quartz (1 %, 1
mm)
Relict textures micropoikolitic, porphyritic
Primary composition rhyolite
Lithofacies columnar jointed, massive
Interpretation sill
Alteration minerals K-feldspar + quartz + sericite + chlorite +
pyrite + magnetite
Alteration textures pervasive recrystallised groundmass of
K-feldspar + quartz + sericite > chlorite,
cleavage defined by sericite, plagioclase
altered to K-feldspar > chlorite +
sericite, K-feldspar overgrowths,
chlorite pseudomorphs of feldspar
microphenocrysts
Distribution/zonation local
Preservation moderate
Alteration intensity intense
Timing syn- to post-intrusion
Style intrusion-related hydrothermal alteration
Geochemistry
SiO
2
73.74 K
2
O 7.6 0 Rb 16 1 Zr 257
TiO
2
0.27 P
2
O
5
0.04 Sr 38 Nb 11
AI
2
O
3
12.10 S 0 Ba 2449 Y 40
Fe
2
O
3
5.51 Total 99.97 Cu 215
MnO 0.03 (vol. free) Pb 17 Al 98
MgO 0.53 Zn 57 CCPI 42
CaO 0.01 Th 17 Ti/Zr 6 .3
Na
2
O 0.13
Strong, pervasive sericite alteration facies
SYNVOLCANIC INTRUSION-RELATED ALTERATION | 1 6 1
DG5
Hand specimen photograph Photomicrograph (ppl)
Sample no. 14336 6
Alteration fades strong, pervasive sericite
Location Jukes Road
Formation Central Volcanic Complex
Succession Mount Read Volcanics
Volcanic fades massive feldspar + quartz-phyric
rhyolite
Relict minerals quartz (2%, 1 mm), plagioclase (2%,
3 mm), hornblende (<1%)
Relict textures spherulitic, glomeroporphyritic
Primary composition rhyolite
Lithofacies massive
Interpretation sill
Alteration minerals chlorite > K-feldspar + sericite > pyrite
+ magnetite
Alteration textures pervasive K-feldspar + quartz +
chlorite, chlorite + sericite + magnetite
pseudomorphs after plagiociase,
chlorite pseudomorphs after
hornblende, recrystallised spherulites
Distribution local
Preservation moderate
Alteration intensity intense
Timing syn- to post-intrusion
Alteration style intrusion-related hydrothermal
alteration
Geochemistry
SiO
2
72.77 Na
2
O 0.15 Rb 16 7 Zr 28 4
TiO
2
0.37 K
2
O 5.06 Sr 18 Nb 13
AI
2
O
3
13.23 P
2
O
5
0.05 Ba 1323 Y 27
Fe
2
O
3
7.11 S 0 Cu 234
MnO 0.04 Total 99.97 Pb 6 Al 97
MgO 1.15 (vol. free) Zn 114 CCPI 59
CaO 0.05 Th 19 Ti/Zr 7.8
1 6 2 | CHAPTER 6
Strong, pervasive sericite alteration facies
DG6
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
143400
strong, pervasive sericite
Jukes Road
Central V olcanic Complex
Mount Read V olcanics
massive feldspar + quartz-phyric rhyolite
plagioclase (2%, 3 mm), quartz ( 1 %,
1 mm)
porphyritic, spherulitic?
rhyolite
massive
unknown
sericite + K-feldspar + quartz + hematite
+ pyrite
plagioclase replaced by sericite,
recrystallised groundmass of sericite +
K-feldspar + quartz, hematite stylolites,
disseminated pyrite, weakly developed
cleavage
local
poor
strong
syn- to post-intrusion
intrusion-related hydrothermal alteration
Geochemistry
SiO2
TiO2
AI A
Fe
2
O
3
MnO
MgO
CaO
75.23 Na
2
O 0.99 Rb 173 Zr 278
0.30 K
2
O 5.18 Sr 22 Nb 13
13.52 P
2
O
5
0.05 Ba 1232 Y 41
3.6 7 S 0 Cu 4
0.02 Total 100.00 Pb 3 Al 86
0.99 (vol. free) Zn 35 CCPI 41
0.04 Th 21 Ti/Zr 6 .5
Hand specimen photograph Photomicrograph (ppl)
I 1 6 3
7 | LOCAL HYDROTHERMAL ALTERATION
RELATED TO VHMS DEPOSITS
An understanding of the mineralogical and chemical zonation
of hydrothermally altered rocks around submarine massive
sulfide deposits is vitally important to both ore genesis studies
and to assist and focus mineral exploration. It has led to a
vast amount of literature on hydrothermal alteration around
VHMS deposits, which various workers have summarised
(e.g. Franklin et al., 1981; Barriga et al., 1983; Urabe et al.,
1983;Lydon, 1984; 1988; Large, 1992; Madeisky and Stanley,
1994; Barrett and MacLean, 1994b; Galley, 1995; Carvalho
et al., 1999; Sanchez-Espana et al., 2000, 2002; Gemmell and
Herrmann, 2001; Large et al., 2001c).
Since the pioneering work in the 1950s and 1960s, when
geologists first recognised the critical link between volcanic-
magmatic processes and massive sulfide genesis (e.g. Stanton,
1955, 1959; Oftedahl, 1958; Gilmour, 1965; Horikoshi,
1969), it became widely accepted that these deposits form
on the seafloor from hydrothermal activity generated during
periods of local quiescence between volcanic eruptive cycles
(Sangster, 1972; Solomon, 1976; Franklin et al., 1983;
Ohmoto and Skinner, 1983). The discovery of seafloor
black smokers and related sulfide chimneys on the present-
day seafloor has further stimulated research and contributed
to an improved understanding of ore forming processes in
the ancient deposits (Rona and Scott, 1993). Over the last
15 years, many researchers have questioned whether all
massive sulfide deposits form by exhalation on the seafloor.
Although it has been recognised for some time that stringer
zones and the lower parts of some massive sulfides formed by
replacement (e.g. Large, 1997), several authors now consider
that replacement of particular volcanic units below the
seafloor maybe be a key process for massive sulfide formation
(e.g. Barriga and Fyfe, 1988; Khin Zaw and Large, 1992;
Allen, 1994b; Bodon and Valenta, 1995; Hannington et al.,
1999; Doyle and Allen, 2003).
This chapter is not a summary of VHMS ore genesis, but
it highlights the features of alteration halos associated with
VHMS deposits, particularly in the Mount Read province,
western Tasmania, and fits the deposits and their alteration
halos in to the broad range of ore deposits that are found
in volcanic and volcano-sedimentary successions. In the later
part of this chapter we provide descriptions, including data
sheets depicting typical alteration facies, of examples from the
Mount Read province and Mount Windsor Subprovince in
eastern Australia that illustrate the range in alteration styles
and zonation associated with the spectrum of submarine
volcanic-hosted base metal ores.
7.1 COMMON FEATURES OF VHMS
DEPOSITS
VHMS deposits display the following features:
They are hosted by submarine volcanic or volcano-
sedimentary successions.
They are the same age as the host volcanic succession
(i.e. the deposits are approximately synvolcanic and/or
synsedimentary).
The host rocks vary from coherent to clastic volcanic or
sedimentary facies and range in composition from basalt
through andesite and dacite to rhyolite.
Most deposits are hosted in thin volcaniclastic units
(<100 m thick) between major volcanic formations.
The economic parts of the deposits typically comprise
massive sulfide, principally pyrite, subordinate sphalerite,
chalcopyrite and galena. The term massive implies greater
than 80 wt% sulfides (Sangster, 1972).
Massive sulfide lenses are commonly, but not always,
aligned parallel to volcanic strata.
Stringer (or stockwork) sulfide zones commonly underlie
the massive sulfides and may contain economic Cu grades.
Metal contents and metal ratios vary considerably. Deposits
include Cu-rich, Au-rich, Cu-Zn, and polymetallic (Cu-
Zn-Pb-Ag-Au) types, but all contain more Zn than Pb.
Ore metals are typically vertically zoned within sulfide
deposits from Cu at the stratigraphic base to Zn, Pb, Ag,
Au and Ba in general order towards the top. Nevertheless,
there are many exceptions to this zonation pattern and
some deposits have no Ba.
Intense hydrothermal alteration of the footwall volcanic
rocks stratigraphically below the massive sulfide, to
chlorite, sericite and quartz is common. By comparison,
the hanging wall rocks are weakly altered or unaltered.
Over 700 VHMS deposits have been recorded around the
world. They range in size from less than 100,000 tonnes to
1 6 4 | CHAPTER 7
over 510,000,000 tonnes (RioTinto, Iberian pyrite belt). The
top 50 deposits, in terms of tonnes of contained Cu + Zn + Pb
metal, are listed in Figure 7.1. Most of these deposits are from
seven major VHMS provinces or districts (Fig. 7.2), from
oldest to youngest: Abitibi belt in Canada, Skellefte district in
Sweden, Mount Read province in Australia, Bathurst mining
camp in Canada, Southern Urals in Russia, Iberian pyrite belt
in Spain and Portugal, and Hokuroku district in Japan.
Cu+Zn+Pb metal content (million tonnes)
10 40
FIGURE 7.1
Pb tonnes.
The 50 largest VHMS deposits in terms of contained Cu + Zn +
FIGURE 7.2 | Locations of the major V HMS provinces around the world.
7.2 | HYDROTHERMAL ALTERATION
HALOS ASSOCIATED WITH VHMS
DEPOSITS
Hydrothermally altered zones proximal to VHMS deposits
may include footwall alteration pipes, stratabound altered
footwall zones and altered hanging wall zones.
Previous studies (e.g. Franklin et al., 1983; Lydon,
1988) emphasised the pipe-like hydrothermal altered zones
in the footwalls of many massive sulfide deposits. They are
common in Archaean deposits in the Abitibi belt in Canada
(e.g. Sangster, 1972) and in the Miocene Kuroko deposits
of the Hokuroku district in Japan (e.g. Urabe et al., 1983).
However, in other districts such as the Mount Read province,
Mount Windsor Subprovince and Lachlan Fold Belt in
eastern Australia, the Iberian pyrite belt, and the Bathurst
mining camp, well-defined alteration pipes are less common,
and stratabound altered footwall zones dominate (Large,
1992). These two styles of altered footwall zones are described
below, with emphasis on the footwall alteration pipes, because
they have received considerable attention from researchers
and their alteration mineral zonation and genesis are better
understood.
Footwali alteration pipes
Figure 7.3 is a schematic cross-section of the geology, sulfide
and alteration zonation related to a typical VHMS deposit.
It is based on our understanding of the Hellyer deposit in
the Mount Read province (Gemmell and Large, 1992), but
also incorporates information on other Australian (Large,
1992), Canadian (Franklin et al., 1981; Lydon, 1988; Lentz
and Goodfellow, 1996), Japanese (Date et al., 1983; Urabe
et al., 1983) and Spanish-Portuguese deposits (Barriga et
al., 1983; Leistel et al., 1998; Sanchez-Espana et al., 2000).
Immediately below the thickest part of the massive sulfide
ore the footwall alteration pipe, which may be oval in plan
but is more commonly elongate along a synvolcanic fault,
contains a concentric series of altered zones. These are, from
the centre of the pipe outwards: siliceous core zone, chlorite
zone, sericite zone, and albite zone, which grades into least-
altered volcanic rocks.
FIGURE 7.3 | Cross-section of idealised mineralisation and alteration zonation
patterns in a footwall alteration pipe beneath a typical VHMS deposit (modified
after Gemmell and Large, 1992; Lydon, 1997). (A) Sulfide mineral zones and
geology. (B) Hydrothermally altered zones.
Siliceous core zones
The siliceous core zones are composed of quartz + pyrite and
quartz + pyrite + sericite chlorite assemblages (e.g. data
sheets HE6, RB4, TH4, WT7, HR8, HN6 and HN7). They
may not always be present, and have only been described
from a few deposits (e.g. Hellyer, Gemmell and Large, 1992;
Brunswick No. 12 and other deposits in the Bathurst mining
camp, Zhang et al., 2003). Siliceous core zones are the most
intensely altered rocks in the centre of the pipes, and are
commonly intersected by networks of pyrite + chalcopyrite
stringer veins (Fig. 7.3). All primary rock textures within
these zones have been completely destroyed due to the
intensity of alteration. In some cases, quartz-rich alteration
assemblages have overprinted earlier chlorite-rich alteration
assemblages creating pseudobreccia textures. Mass-change
calculations indicate that gains of 50-100 g/100 g, mainly
due to Si addition, are common within siliceous core zones
(Gemmell and Large, 1992, Fig. 11). Lentz and Goodfellow
(1996) reported SiO
2
gains of up to 300% in the siliceous
core zone in the centre of the alteration system below the
Brunswick No. 12 massive sulfide deposit. The siliceous core
represents the zone of maximum hydro thermal fluid flow and
highest temperatures.
Chlorite zones
Chlorite zones are dominated by chlorite (>50 wt% and
commonly >80 wt%), with subordinate quartz + pyrite +
sericite carbonate (e.g. data sheets HE4, RB5 and HR7).
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 1 6 5
These zones are typically fine grained, dark and massive,
preserving no volcanic or sedimentary textures. In many
deposits these zones host pyrite + chalcopyrite stringer
veins (e.g. Woodlawn deposit in eastern Australia and most
Noranda district and Iberian pyrite belt deposits). Chlorite
zones are commonly deformed during tectonic events,
resulting in chlorite schist zones, which may be strung-out
or dislocated from the massive sulfide ore (Sangster, 1972).
Studies of chlorite composition from Canadian, Australian
and Japanese deposits indicate that the inner chlorite zones
are dominated by Mg-rich chlorite, with a general increase
in Fe/Mg ratio passing from the inner chlorite zone to the
outer edge of the sericite zone (e.g. Riverin and Hodgson,
1980; Urabe et al., 1983; Paulick et al., 2001). Nevertheless,
reverse trends have been recorded, where chlorite becomes
more Fe-rich towards the core (e.g. Eastoe et al., 1987; Lentz
et al., 1997). Variations in chlorite composition are discussed
in more detail in Section 4.2.
Sericite zones
Sericite zones surround the inner chlorite zones and are
characterised by assemblages of sericite + chlorite + quartz +
carbonate + pyrite (e.g. data sheets HE3, RB3, WT3, WT6,
HR6 and HN5). At Hellyer, rocks in the sericite zone are
strongly to intensely altered, with up to 70 wt% sericite
and sparse relict primary textures. In other deposits, the
alteration intensity in the sericite zone decreases towards the
outer margin, where altered rock grades into the least-altered
footwall rocks. In many cases, the sericite zones are laterally
extensive and merge with stratabound altered zones away
from the central pipe (e.g. Mount Chalmers, Large and Both,
1980). Minor disseminated sphalerite or stockwork Zn may
occur in the sericite zone, whereas Cu-enrichment is more
common in the chlorite zone.
Albite zones
Some authors have described weakly altered zones of albite
+ chlorite sericite that surround the main sericite zones
(e.g. Iijima, 1974; Green et al., 1981; Urabe et al., 1983;
Relvas et al., 1997; Goodfellow and McCutcheon, 2003).
Although albite zones have not been widely described or
accepted in all districts they are discussed here because of
their significance to mineral exploration. It may be difficult to
distinguish hydrothermal albite facies from the background
diagenetic alteration facies, as within albite zones primary
volcanic textures are commonly preserved. In the Hokuroku
district this zone has an albite + sericite + chlorite assemblage
(Iijima, 1974; Urabe et al., 1983). In some deposits of the
Iberian pyrite belt an outermost halo of Na-sericite has been
recognised (Relvas et al., 1997). In the Bathurst mining camp,
Goodfellow and McCutcheon (2003) described an outermost
altered zone of albite + Mg-rich chlorite surrounding the
footwall alteration pipe. They described an increase in patchy
albite, which replaced K-feldspar phenocrysts, in proximity
to the pipe. Similar albite-altered rocks have been recognised
adjacent to the Hercules footwall alteration pipe in the Mount
Read province (Large et al., 1996). Further research is required
1 6 6 | CHAPTER 7
to characterise the features of these outermost albite + chlorite
+ sericite altered zones and to distinguish them from regional
diagenetic albite alteration facies.
Variations in alteration zonation
The three main hydro thermally altered zones (siliceous core,
chlorite and sericite) are not present in all footwall alteration
pipes beneath VHMS deposits, and in some cases additional
zones, such as carbonate or talc zones, exist. Some variations
from the idealised alteration pipe model outlined in Figures
7.3 and 7.4A are:
FIGURE 7.4 | Different patterns of mineral zonation in footwall alteration pipes.
(A) Generalised model. (B) Hokuroku district model. (C) Noranda district and
Iberian pyrite belt model. (D) Mattagami district model.
The Kuroko deposits do not have an inner chlorite zone.
Shirozu (1974) described intensely altered volcanic rocks
below the massive sulflde ore and surrounding the siliceous
Cu-stockwork ore, as strongly silicified with abundant
sericite and very little chlorite (Fig. 7.4B).
Deposits in the Noranda district and the Iberian pyrite belt
commonly have intense chlorite core zones surrounded by
sericite zones (Fig. 7.4C), but without siliceous core zones
(Franklin et al., 1983; Carvalho et al, 1999; Sanchez-
Espana et al., 2000).
In the Iberian pyrite belt, Relvas et al. (1997) recognised
a Na-bearing sericite (paragonite) zone extending beyond
and above the main sericite zone at both the Neves Corvo
and Aljustrel deposits.
In the Mattagami district in Canada, the intensely altered
central core of the footwall alteration pipe is talc-rich (Fig.
7.4D) and surrounded by chlorite and sericite zones (e.g.
Large, 1977; Roberts and Reardon, 1978).
In several eastern Australian deposits, Mg- and/or Fe-
bearing carbonates are common in the alteration mineral
assemblages (Large et al., 2001c). At Hellyer a chlorite
+ dolomite zone occurs below the massive sulflde near
the top of the chlorite zone (Fig. 7.4A, data sheet HE5,
Gemmell and Large, 1992). A more massive dolomite
zone is developed at the western margin of the altered
stringer pipe at Mount Chalmers (Large and Both, 1980).
Carbonate-rich zones are common in the footwall of many
Iberian pyrite belt deposits, either marginal to the massive
sulflde or distributed throughout the footwall alteration
systems (e.g. Rio Tinto, Williams et al., 1975, Solomon et
al., 1980; La Zarza, Strauss et al., 1981; Tharsis, Tornos et
al., 1998).
Lydon (1988, Fig. 8) included an Fe-oxide zone at the top of
the pipe below the massive sulflde in his footwall alteration
pipe model, compiled from a number of deposits, but there
are few examples of this facies.
Stratabound altered footwall zones
Many massive sulfide deposits, possibly half, do not have
footwall alteration pipes, but are underlain by stratabound or
semi-conformable altered zones, which extend laterally for up
to several kilometres away from the deposits (Figs 3.16B, C
and 7.5). These stratabound zones may extend for between 30
and several hundred metres below the massive sulfide. They
are typically developed around sheet-like deposits, and are
mainly associated with Zn-Pb-rich deposits (e.g. Rosebery,
Scuddles and Bathurst mining camp deposits). Sangster
(1972) and Goodfellow and McCutcheon (2003) consider
stratabound altered footwall zones to originally have been
pipes that were sheared and transposed parallel to stratigraphy
during tectonic deformation. However, this does not explain
stratabound footwall zones in weakly to moderately deformed
volcanic successions, such as the Mount Read province and
Iberian pyrite belt. In the Iberian pyrite belt, Tornos (in press)
notes that although many previous studies have claimed pipe-
like morphologies to the footwall stockworks and altered
zones (e.g. Costa et al., 1995; Carvalho et al., 1999; Saez et
al., 1999), recent studies have defined irregular to stratabound
morphologies for the footwall alteration associated with most
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 1 6 7
deposits (e.g. Tharsis, Tornos et al., 1998; Neves Corvo,
Relvas et al., 2000).
Stratabound altered footwall zones have similar alteration
mineral assemblages to footwall alteration pipes, but the zones
are distributed parallel to stratigraphy, rather than at right
angles. In some cases, the siliceous core and chlorite zones are
confined to the immediate footwall of the thickest Cu-rich
part of the massive sulfide lens (e.g. Rosebery and Thalanga).
In contrast to the pipes, the sericite zones of stratabound
altered footwall zones are the volumetrically dominant
zones being both laterally and vertically extensive. Massive
carbonate zones are more common in stratabound than
pipe-like alteration systems (Large et al., 2001c), typically
occurring immediately along strike from the massive sulfide
ore lenses (e.g. Rosebery, Thalanga and Mount Chalmers).
From an exploration perspective stratabound altered footwall
zones typically present broader targets than footwall alteration
pipes; however, they tend to be more diffuse and thus
create challenges when searching for the associated VHMS
deposits.
Altered hanging wall zones
Compared to footwall alteration, hanging wall alteration is
typically less intense and therefore has not received much
attention in the ore-deposit literature. Visible hanging wall
alteration mineral assemblages are commonly sericite-rich
and restricted to a few metres above the massive sulfide ore.
However, detailed petrographic and geochemical studies have
extended some altered zones to several tens of metres into
the hanging wall. There are a number of exceptions to the
generally limited altered hanging wall zones.
Copper-Au-rich VHMS deposits that form by replacement
below the seafloor may exhibit extensive sericite-rich altered
hanging wall zones (Fig. 7.6A). Mount Lyell in the Mount
Read province and Highway-Reward in the Mount Windsor
Subprovince are good examples of subseafloor sericite zones
and are described in detail in Sections 7.7 and 7.10.
In stacked ore systems, such as Millenbach and Amulet
deposits in the Noranda district, and Que River in the Mount
Read province, the lower ore body in the stack has an intense
altered hanging wall zone, similar to that in the footwall (Fig.
7.6B). This is due to hydrothermal fluids moving through the
lower ore lens and hanging wall on their way to depositing
the upper ore lenses.
Significant altered hanging wall zones may have developed
in situations where the hanging wall volcanic or sedimentary
units were deposited while the massive sulfide was still forming
on the seafloor. This is the case for the Hellyer deposit, which
has a hanging wall alteration plume that comprises a core
of fuchsite + carbonate (e.g. data sheet HE10), surrounded
by successive halos of chlorite + carbonate, quartz + albite,
and finally patchy sericite (Fig. 7.6C, Gemmell and Fulton,
2001). The interpretation that the hanging wall alteration
zones formed after the seafloor massive sulfide is not in doubt
because of sulfide and barite clasts in the directly overlying
volcaniclastic debris-flow unit (McArthur and Dronseika,
1990;Sharpe, 1991).
Although primary plagioclase destruction is a key process
in the hydrothermal alteration associated with VHMS
FIGURE 7.5 | Examples of stratabound altered footwall zones (modified after
Ashley et al., 1988; Large, 1992; Large et al., 2001c). (A) Scuddles, Western
Australia, in plan view. (B) Teutonic Bore, Western Australia, in cross-section.
(C) K lens at Rosebery, western Tasmania, in cross-section. (D) Thalanga,
Queensland, in cross-section. Abbreviations are: FW = footwall and HW =
hanging wall.
deposits, there are a few examples where albite zones exist in
the hanging wall of the deposit. At the Henty gold deposit
albite + quartz is a common hanging wall alteration mineral
assemblage (see Section 7.8, data sheet HN3), and at Hellyer
albite forms one of the altered zones in the hanging wall
alteration plume (Fig. 7.6C, data sheet HE8).
In addition to these examples of obvious altered
hanging wall zones, some recent detailed lithogeochemical
1 6 8 | CHAPTER 7
FIGURE 7.6 | Examples of cross-sections through altered hanging wall zones
associated with VHMS deposits. (A) Cu-Au deposit (modified after Large et al.,
2001c). (B) A stacked ore system (after the Millenbach deposit, Knuckey etal.,
1982). (C) Hellyer hanging wall alteration plume (modified after Gemmell and
Fulton, 2001).
studies have defined subtle geochemical halos in otherwise
least-altered hanging wall volcanic rocks, which may extend
several hundreds of metres above the massive sulfide ore
(Large et al., 2001b). For example, at Rosebery a hanging wall
alteration halo can be defined using three geochemical
parameters that may be applied during exploration: (1) whole-
rock Ba/Sr ratio, which outlines a halo extending about 100 m
into the hanging wall; (2) Mn content of carbonate, which
is anomalous over the same interval; and (3) whole-rock Tl
content, which forms a halo that extends over 200 m into
the hanging wall volcanic rocks. More detail of the Rosebery
alteration system is provided in Section 7.6.
Chemical reactions and mass changes
Footwall alteration results from the reaction of hydrothermal
fluid (principally composed of heated seawater) with volcanic
rocks. The temperature of the fluids that form VHMS deposits
are estimated to vary from about 200 to 350C based on
fluid inclusion evidence (Pisutha-Arnond and Ohmoto,
1983; Khin Zaw et al., 1996), the study of present day black
smoker systems (Goldfarb et al., 1983) and thermodynamic
calculations of mineral stabilities (e.g. Sato, 1973; Large,
1977; Ohmoto et al., 1983). Fluid salinities approximate that
of seawater, although values of up to four times seawater have
been recorded (de Ronde, 1995; Solomon et al., 2002). The
pH varies from about 3 to 7 based on mineral assemblage,
thermodynamic considerations and measurements at black
smoker vents (Huston and Large, 1989; Scott, 1997).
The principal result of interaction of this hot, mildly acidic
to neutral fluid with volcanic rocks as it ascends towards the
seafloor is the breakdown of feldspars and volcanic glass, and
their replacement by sericite, quartz, chlorite and carbonate.
Petrographic evidence of these reactions is shown in Figure
2.5 which depicts a series of progressively hydrothermally
altered pumice-rich rocks from the footwall to the Hercules
deposit in the Mount Read province.
Reactions that describe these footwall alteration processes
may include reaction R7.1 (from Sanchez-Espana et al., 2000)
in the sericite zone:
3NaAlSi
3
O
8
+ K
+
+ 2H
+
albite
- KAl
3
Si
3
O
10
(0H)
2
+ 6SiO
2
+ 3Na
+
sericite quartz
(R7.1)
This reaction is typical of sericite replacing albite in the
outer part of the alteration system. The reaction involves a
gain in K from hydrothermal fluid and loss of Na in the rock
as albite is replaced. Silica is conserved by the deposition of
quartz. Overall, the reaction leads to an increase in fluid pH
due to the consumption of H
+
. Reactions in the sericite zone
also involved sericitisation of K-feldspar and plagioclase in
addition to albite.
In the chlorite zone, there are two potential reactions
(Pisutha-Arnond and Ohmoto, 1983):
Note, these reactions can be written with H
4
SiO
4
(aq) or
SiO
2
(quartz) + 2H
2
O.
Both of these reactions involve the addition of Mg and
Fe
2+
from the fluid and the loss of alkalies (Na or K) and H
from the rock. However, the replacement of albite by chlorite
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS I 1 6 9
involves Si loss as H
4
SiO
4
, compared to the replacement of
sericite by chlorite, which involves Si gain. Mass transfer
calculations in chlorite zones invariably indicate significant
loss of Si (e.g. Gemmell and Large, 1992; Barrett and
MacLean, 1994b) suggesting that reaction R7.2 is the key
reaction. Also both of these chlorite replacement reactions
(R7.2 and R7.3) involve a release of H
+
to the fluid and will
cause an increase in fluid acidity. This means that continued
chloritisation of volcanic rocks over an extensive area might
produce moderately acidic fluids, which may subsequently
cause silicic (quartz + sericite), or in the extreme case, argillic
(pyrophyllite kaolinite sericite) alteration assemblages up-
flow from the chlorite zone. Examples of this occur at the
Neves Corvo and Lagoa Salgada deposits in the Iberian pyrite
belt where Relvas et al. (1994, 1997) have reported dombassite
and pyrophyllite in the chlorite-rich central stockwork zones
below the massive sulflde ores. Other reactions in the chlorite
zone, not listed here, include chloritisation of mafic minerals
such as biotite and amphibole.
In the siliceous core zone at the centre of the hydro thermal
system, three reactions are proposed:
Quartz is the main alteration mineral in this zone and
is associated with minor sericite and chlorite. Mass balance
calculations indicate that considerable SiO, gains. Aluminium
is commonly immobile (Gemmell and Large, 1992; Barrett
and MacLean, 1994b), but is significantly diluted by large
Si gains. Quartz is probably deposited in the siliceous core
zone according to reaction R7.4, as this involves mass gain of
Si without loss of Al. Enrichment of Si in the hydrothermal
fluid may be due to leaching of Si from the footwall volcanic
rocks during chloritisation (reaction R7.2). Leached Si is then
deposited in the hydrothermal vent immediately below the
seafloor. Silica deposition may be caused by rapid conductive
cooling (Fournier and Potter, 1982), mixing with seawater,
or intense fluid-rock interaction at high-fluid-rock ratios.
Replacement of both chlorite and albite by quartz (reactions
R7.5 and R7.6) requires a strongly acidic fluid and results
in loss of Al as Al(OH)
3
(aq). Aluminium mobility of this
type is rare in VHMS systems, but may occur in intensely
silicified zones associated with acid alteration. Mass balance
calculations suggest this was the case in the siliceous core zone
of the Henty volcanogenic gold deposit (Callaghan, 2001, see
also Section 7.8).
In summary, by writing simple chemical reactions to
describe replacements in the major altered zones, it is possible
to gain some idea of the chemical processes, elemental gains
and losses, and variations in pH of the hydrothermal fluid.
Most reactions in the altered footwall zone involve the
breakdown of feldspar and loss of Na (and usually Ca) to the
fluid. Major gains include K in the sericite zone, Mgand Fe in
the chlorite zone and Si in the siliceous core zone. In addition
to Na and Ca, other losses include Si in the chlorite zone,
and very rarely Al in the siliceous core. The fluid pH does not
show any systematic unidirectional change. Initially mildly
acidic fluids will become less acidic during sericitisation, but
more acidic during chloritisation. Intense siliceous core zones
may be related to rapid cooling, fluid mixing or intense fluid-
rock interaction of Si-saturated fluids during the peak of the
hydrothermal activity.
Alteration box plot trends in altered footwall
zones
The AI-CCPI Alteration box plot (Section 2.5) is a simple
way of tracking whole-rock compositional changes and
relating them to alteration mineralogy and position in the
altered footwall zones. In Figure 7.7, the fields of the major
altered zones are shown on the Alteration box plot in relation
to two alteration indices, Al and CCPI. Line AD represents
an array of altered felsic volcanic rock samples passing from
the outer edge of the altered footwall zone into the core zone
proximal to massive sulfide ore. Line ED represents a similar
sample array for mafic volcanic rocks.
Let us first consider line AD. Point A represents a rhyolitic
rock outside the altered zone. Sericite and weakly chlorite
altered rocks in the margins of the altered zone increase the
Al due to Na and Ca depletion, whereas the CCPI remains
relatively constant. Altered samples plot progressively along
the AB segment of the trend toward the plotted position of
sericite (phengite) on the perimeter of the Alteration box plot.
In the inner part of the sericite zone, the Al is commonly
greater than 90 and the trend becomes vertical due to a strong
increase in the CCPI caused by gains in Fe and Mg related to
FIGURE 7.7 | The AI-CCPI Alteration box plot showing trends for altered
footwall zones. These data are based on case studies presented in Large et al.
(2001a).
1 7 0 | CHAPTER 7
increasing pyrite and chlorite in the rock (segment BC). The
final segment CD represents the chlorite zone, where CCPI
reaches its maximum (80-100), due to the abundance of
chlorite and pyrite proximal to massive sulfide ore. Although
quartz is not plotted on the Alteration box plot, due to the
absence of SiO
2
from the two alteration indices, samples from
the siliceous core zone commonly plot in the CD segment
due to the presence of minor chlorite and pyrite. If carbonate
is present in the chlorite zone, as at Hellyer and Thalanga, the
samples typically plot between F and D.
In cases where the footwall comprises mafic volcanic
rocks, ED is the common trend from the edge to centre of
the footwall alteration pipe. This difference is due to the fact
that mafic rocks generally have higher Fe and Mg contents,
and thus greater initial values of CCPI compared with felsic
rocks (Fig. 2.10). Consequently, chlorite is generally more
abundant in the outer sericite zone and the combination of
increasing chlorite and pyrite gives a trend along ED toward
the core of the footwall alteration pipe (e.g. Hellyer, Section
7.5).
The genesis of footwall alteration pipes
The presence of tightly constrained altered zones, in a
circular or more commonly elongate pipe, suggests that
hydrothermal fluids were focussed along synvolcanic faults or
fault intersections, and massive sulfide deposition occurred
where the faults intersected the seafloor. The zonation in the
footwall alteration pipe is commonly interpreted to reflect a
decreasing thermal gradient away from the fluid conduit (e.g.
Large, 1977; Riverin and Hodgson, 1980).
up, and becomes more acidic due to fluid-rock interaction,
metals are leached from the volcanic succession (e.g. Kajiwara,
1973; Spooner and Fyfe, 1973; Solomon, 1976; Large, 1977;
Ohmoto, 1996). Alternatively, metal-rich magmatic fluid may
be derived from the crystallisation of a magma, which is also
the source of volcanism (e.g. Urabe and Sato, 1978; Henley
and Thornley, 1979; Sawkins and Kowalik, 1981; Stanton,
1985, 1990). For supporting evidence and relative merits of
these two models the reader is referred to recent discussions
by Lydon (1996) and Ohmoto (1996).
Recent research suggests that both fluids and metals are
probably derived from magmatic and seawater sources (e.g.
Fig. 7.8C). Distinguishing criteria for the source includes
the deposit style, proximity to volcanic centres, alteration
mineralogy, metal ratios of the deposits and the salinity
and composition of primary fluid inclusions. For example,
Large (1992) suggested that relatively soluble chloride-metal
complexes, such as Zn, Pb and Ag, are probably derived
principally from seawater leaching of the volcanic succession,
whereas the less soluble metals, such as Cu, Bi and Sn, may
be sourced directly from the magma chamber. Gold could
be derived either by seawater leaching of volcanic rocks as a
bisulfide-Au complex, or directly from magma as a chloride
complex or in a volatile phase (Fig. 7.8C). Goodfellow and
McCutcheon (2003) proposed a similar dual metal source
for the massive sulfide deposits of the Bathurst mining camp,
with the largest deposits having a major magmatic-metal
component. Recently Solomon et al. (2004) have compared
the salinity and composition of fluid inclusions in the stringer
zones of the Hellyer VHMS deposit with those of porphyry
Cu deposits, and concluded that the metals at Hellyer had a
magmatic source.
Source considerations Fluid-rock interaction in the alteration pipe
There are two competing models for the source of fluids The concept that the alteration zonation in footwall alteration
and metals that form VHMS deposits (Fig. 7.8A and B): (1) pipes is a function of decreasing temperatures and fluid-rock
evolved seawater, and (2) magmatic fluid. The first model ratios has recently been tested by Schardt et al. (2001) using
involves seawater convecting through a volcanic succession a thermodynamic model of fluid-rock interaction between
above an intrusion or magma chamber. As the seawater heats heated evolved seawater and an andesitic precursor (Fig.
FIGURE 7.8 | Models for fluid flow and metal source in VHMS hydrothermal systems (after Large, 1992). (A) Metals are derived from deep seawater leaching of
the volcanic succession and basement. (B) Metals are sourced directly from the magmatic vapour plume, with no significant leaching of volcanic rocks. (C) A mixture
of volcanic and magmatic sources, with low-solubility metals (i.e. Cu and Au) provided from magma and high-solubility metals (i.e. Zn, Pb and Ag) from seawater
leaching of volcanic rocks.
7.9). This modelling was based on geochemical data from
the Hellyer deposit. The classical sequence of altered footwall
zones observed in many VHMS deposits (from the core to
the margin of the pipe: quartz * chlorite > sericite) was
reproduced by simulating the reaction between a 250350
c
C
fluid, with a pH of 4.55.0, and andesite under conditions
of decreasing fluid-rock ratio and temperature. Simulated
cooling from 350 to 100C reproduced the full range of
footwall alteration mineral assemblages. The pH of the fluid
showed little variation, from 4.5 to 4.0 (Schardt et al., 2001).
Mg-rich chlorite formed in the inner chlorite zone, and Fe-
rich chlorite developed in the outermost part of the sericite
zone, similar to the pattern observed at many massive sulfide
deposits. This modelling was carried out using a Mg-bearing
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 1 7 1
fluid, with the assumption that a component of seawater-
derived Mg was incorporated into the fluid at depth.
The modelling has shown that sericite zones form at
temperatures below 250C from the reaction of andesitic
rocks with mildly acidic solutions (pH = 4.04.5). Extensive
Mg-chlorite zones are favoured by higher temperatures (250
300C) and less acidic fluids (pH = 4.5-5.5). At lower pH,
kaolinite and pyrophyllite are likely to develop in the sericite
zone. At higher pH and lower temperatures (<200C), K-
feldspar is developed at the outer margin of the sericite zone
and in least-altered andesitic rocks (Schardt et al., 2001).
Although carbonate alteration was not taken into account by
this modelling, it is likely that chlorite + carbonate assemblages,
such as those developed adjacent to massive sulfides or at the
FIGURE 7.9 | Thermodynamic model of fluid-rock interaction between heated evolved seawater and Hellyer andesite
(modified after Schardt et al., 2001). (A) Modelled mineralogical variations resulting from fluid-rock interaction with
decreasing temperatures. (B) Schematic representation of simulated water-rock interaction as a function of temperature.
1 7 2 | CHAPTER 7
periphery of many footwall alteration systems, are indicative
of more alkaline conditions. These may develop where hot,
near-neutral hydrothermal fluids have mixed with and heated
seawater, leading to saturation of carbonate at the margins of
the hydrothermal up-flow zones (Large et al., 2001c).
A model for the development of footwall alteration pipes
Using the results of the thermodynamic modelling of Schardt
et al. (2001), numerical fluid-flow modelling by Yang and
Large (2001), and previous thermodynamic modelling
of metal-sulfide growth by Huston and Large (1989), it is
possible to speculate on the progressive development of
subseafioor zoned alteration pipes (Fig. 7.10).
Stage 1 (Fig. 7.10): Initial hydrothermal fluid flow is
upwards along a sub-vertical permeable fault zone towards
the seafloor. As the rising hydrothermal plume approaches
within 1 km of the seafloor, secondary near-surface seawater
convection above the plume head may enhance normal
diagenetic reactions in volcanic rocks adjacent to the fault,
causing increased formation of zeolites, smectites and Mg-rich
chlorite, and albite replacement of primary feldspars. During
ongoing diagenesis and subsequent metamorphism this will
produce an outer albite zone (albite + sericite + chlorite),
which is commonly difficult to distinguish from regional
diagenetic and metamorphic mineral assemblages.
Stage 2 (Fig. 7.10): As low-temperature and mildly acidic
hydrothermal fluids, (T<250C, pH = 4.0-4.5) continue
to move upwards to the seafloor, sericite-rich alteration
overprints the early albite zone and expands out from the
fluid conduit to form a pipe-like sericite zone. During this
stage sphalerite + galena + pyrite massive sulfides deposit on
the seafloor above the sericite zone. Minor pyrite + sphalerite
+ galena stringer mineralisation may also occur in the core of
the sericite zone at these temperatures (Eldridge et al., 1983).
Convective near surface reflux of seawater leads to an albite +
chlorite zone extending laterally into the least-altered volcanic
rocks surrounding the sericite zone.
Stage 3 (Fig. 7.10): As the hydrothermal system intensifies
and the temperature of the discharging fluid rises above
250C, Mg-chlorite is stabilised adjacent to the main conduit
and a chlorite zone develops in the core of the footwall
alteration pipe. Under these high-temperature conditions,
the fluid is capable of carrying significantly more Cu (e.g.
Huston and Large, 1989), which is deposited in stringer veins
in the central chlorite zone and in the base of the massive
sulfide. Within the massive sulfide mound, Cu progressively
displaces Zn upwards. Eldridge et al. (1983), Huston and
Large (1989) and Hannington and Scott (1989) described
this zone refining process. As the alteration system evolves,
the pH of the fluid initially increases during formation of
the sericite zone (from 4 to 5.5) due to consumption of H
+
(reaction R7.1), and subsequently falls during the formation
of the chlorite zone (reactions R7.2 and R7.3).
FIGURE 7.10 | Model for the evolution of the footwall alteration pipe in a mound-style massive sulfide deposit. Stage 1: an initial low-
temperature hydrothermal system produces an albite zone. Stage 2: increasing temperature results in the development of the sericite and
Zn + Pb-rich sulfide zones. Stage 3: higher temperatures produce the chlorite and Cu + Zn +Pb-rich sulfide zones. Stage 4: maximum
temperatures and low pH result in the siliceous core and Cu + Pb + Zn-rich sulfide zones. This model does not apply to all VHMS
deposits, some of which may form in brine pools (Solomon and Groves, 1994; Solomon and Quesada, 2003).
Stage 4 (Fig. 7.10): With increasing temperature (300 to
35O
C
C) during chloritisation, Si is continually leached from
the volcanic rocks (reaction R7.2) and the fluid becomes
supersaturated in Si. Consequently, quartz is precipitated
(reaction R7.4) in the upper-central part of the alteration
pipe, forming a siliceous core zone. Maximum metal
precipitation in both the stringer zone and massive sulfide
is commonly associated with this stage. Subsequently, the
hydrothermal system wanes and collapses with an influx of
heated near-surface seawater that leads to overprinting by
lower temperature mineral assemblages, which are commonly
dominated by carbonates or barite, depending on the
oxidation level of the overlying water column.
Some previous workers have suggested that the zonation
in alteration pipes, from Mg-Fe chlorite in the core to sericite
at the margins, relates to the entrainment and mixing of
Mg-bearing seawater with a Mg-poor hydrothermal fluid
below the massive sulfide (e.g. Roberts and Reardon, 1978;
Lydon and Galley, 1986). However, Riverin and Hodgson
(1980) suggested that the presence of Mg-rich chlorite in the
central and most intensely altered zone of the alteration pipe,
and the abundance of Mg-chlorite in pyrite + chalcopyrite
veins in the stringer zone, is consistent with Mg derived
from hydrothermal fluid rather than seawater. In the model
outlined in Figure 7.10, we have assumed the hydrothermal
fluid is Mg-bearing, possibly either due to entrainment of
seawater at considerable depth (>1 km) below the seafloor or
due to the leaching of Mg from mafic volcanic rocks deep in
the volcanic succession.
Although near-surface entrainment of seawater is
considered to be important in stages 1 and 2 of our model,
it is likely to result in an increase in the rate and consequent
grade of diagenetic alteration. This would lead to Na-Mg
metasomatism and albite + chlorite formation at the margins
of the alteration pipe, rather than Mg metasomatism and
chlorite development within the core of the pipe, as previously
proposed (cf. Franklin et al., 1981).
In the thermodynamic modelling of Schardt et al. (2001),
temperature and pH were shown to be the principal factors
controlling the balance between chlorite and sericite zones in
the footwall alteration pipe. However, two other factors also
need consideration: (1) the composition of the immediate
footwall volcanic rocks, and (2) the initial chemistry of the
modified seawater as it rises up the conduit (e.g. Large, 1977).
In the first case, particularly at low fluid-rock ratios, chlorite
alteration is favoured in mafic host rocks and sericite alteration
in felsic host rocks. However, at high fluid-rock ratios typical
of the central parts of the hydrothermal pipe, the alteration
mineral assemblage is controlled by fluid chemistry rather than
rock chemistry. In the second case, seawater-rock interactions
in deep rhyolite-dominated footwall volcanic successions,
similar to those in the Hokuroku district and southern Mount
Read province, will generate modified seawater hydrothermal
solutions enriched in K and Si, but generally depleted in
Mg, Fe and Ca. These fluids will result in sericite zones as
they approach the seafloor. In contrast, footwall volcanic
successions dominated by andesite and basalt, similar to those
in the northern Mount Read province and the Abitibi belt,
will generate evolved seawater fluids enriched in Mg, Fe and
Ca, with lesser K and Si, and are more likely to develop zoned
chlorite-sericite alteration pipes.
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 1 7 3
In summary, the footwall alteration mineral assemblages
in VHMS systems are probably controlled by three factors:
(1) the initial composition of the convective seawater-
dominated ore fluid, which is constrained by the relative
abundance of mafic versus felsic volcanic rocks deep in the
succession; (2) the temperature and pH regime during fluid-
rock interactions in the footwall discharge zone (i.e. lower
temperature, acidic conditions favour sericite development,
whereas higher temperature and/or more neutral pH
conditions favour chlorite formation); and (3) the composition
of the immediate footwall host rocks.
Genesis of stratabound altered footwall zones
Stratabound altered footwall zones (e.g. Rosebery, Scuddles
and Teutonic Bore; Fig. 7.5) are interpreted to result from
hydrothermal-fluid flow parallel to volcanic strata (Fig. 7.11),
rather than at right angles to the stratigraphy as in the case
for footwall alteration pipes. Alteration pipes are commonly
developed in relatively impermeable footwall volcanic rocks
(e.g. the coherent or clastic facies of lavas and synvolcanic
intrusions) where fluids are focussed along sub-vertical syn-
volcanic faults (Fig. 7.10). In contrast, stratabound altered
footwall zones are more commonly developed in volcanic
rocks with moderate- to high-stratal permeability (e.g.
volcaniclastic facies such as pumice breccia and volcanic
FIGURE 7.11 | Genetic models for the formation of stratabound altered
footwall zones related to VHMS mineralisation. Fluid flow below and parallel
to the seafloor and stratigraphy is controlled by the distribution of permeable
volcanic facies (e.g. volcaniclastic units), or impermeable cap-rocks (e.g. sills
or lavas). (A) Stratabound subseafloor replacement mineralised and altered
zones (e.g. Mount Lyell deposit, Mount Read province and TAG deep Cu zone,
Middle Valley, Juan de Fuca Ridge). (B) Stratabound ore lens and altered
zones confined below an impermeable volcanic unit such as a sill (e.g. K lens at
Rosebery, Mount Read province).
174 | CHAPTER 7
sandstone). In high-permeability rocks, hydrothermal fluids
move laterally along the strata, sub-parallel to the seafloor,
and metals are deposited due to the mixing of hydrothermal
fluids with seawater and/or cooling. In these cases, fluids are
poorly focussed and alteration tends to be of lower intensity
and greater in lateral extent (Fig. 7.11). Sericite-altered rocks
dominate stratabound altered footwall zones, and chloritic
and siliceous zones are restricted to the immediate proximity
of massive sulfides, where temperatures and fluid-rock ratios
were at a maximum. At K lens in the Rosebery deposit,
Allen (1994b) proposed that rising hydrothermal fluids were
constrained to flow laterally below an impermeable quartz
porphyry sill. As a result, stratabound massive sulfide and
associated stratabound altered zones developed beneath the
sill by replacement of more permeable and chemically reactive
felsic pumice breccias (Fig. 7.1 IB).
Significance of pyrophyllite and kaolinite in VHMS
systems
Pyrophyllite and kaolinite are generally rare in VHMS
altered zones; however, because they are only stable under
relatively acidic conditions their presence warrants some
discussion. Kaolinite has been reported from the sericite
and montmorillonite zones of some of the Japanese Kuroko
deposits (e.g. Iijima, 1974; Ohmoto, 1996). It has also
been reported in the altered footwall zone of the Mount
Chalmers Cu-Au VHMS deposit in Queensland (McLeod,
1987). Pyrophyllite exists in the sericite zone of the Western
Tharsis VHMS deposit in the Mount Lyell field, Mount
Read province (e.g. data sheets WT4 and WT5: Huston and
Kamprad, 2001), and is also reported in the stockwork zones
of several VHMS deposits in the Iberian pyrite belt (Relvas
etal., 1997).
Figure 7.12 shows that the stability relationship between
kaolinite and pyrophyllite is temperature dependent:
pyrophyllite being stable above 280C, whereas the stability
between muscovite and pyrophyllite is controlled by the
3.
K+
/3.
H+
ratio. For a fluid with 3.
K+
varying from 0.1 to 0.01,
which is considered the range for VHMS-related fluid, then
pyrophyllite is stable at pH values of less than 34 units.
FIGURE 7.12 | Stability relations among selected silicate minerals at 500 bars
pressure (modified after Beane and Titley, 1981).
There are two ways that acidic fluids may be generated
to stabilise kaolinite or pyrophyllite in VHMS systems. The
first is by acid-producing fluid-rock reactions, such as the
replacement of albite (and volcanic glass) by Mg-Fe chlorite
(reaction R7.2). This may result in a reduction in fluid pH
by about 1 unit (from 5.5 to 4.5); however, buffering by
sericite + chlorite assemblages in the rock will generally
prevent the pH dropping below 3.5, which is needed for
kaolinite and pyrophyllite formation. These acidic alteration
minerals can only form by this method in volcanic rocks that
contain negligible K (e.g. tholeiitic basalts) and thus contain
no sericite to buffer the pH. The second method is by the
introduction of magmatic volatiles at some stage during the
life of the hydrothermal system (Ohmoto, 1996). Cooling
of a magmatic gas containing SO
2
at temperatures below
400C will increase fluid acidity to levels below pH = 3.5 by a
reaction similar to R7.7 (Burnham and Ohmoto, 1980):
4SO
2
(g) + 4H
2
O(1) - H
2
S(aq) + 3H
+
+ 3HSO
4
~ (R7.7)
Researchers have recently argued that the presence of
pyrophyllite or kaolinite in VHMS altered zones supports
the theory that these deposits are not simply the products of
seawater convection, but that their genesis involves input of
a magmatic-derived, low-pH fluid (e.g. Sillitoe et al., 1996;
Huston and Kamprad, 2001).
Metamorphism of altered zones
Few detailed studies have been published on the effects of
contact and regional metamorphism of VHMS-related
altered zones. Medium- to high-grade metamorphism of
chlorite and sericite zones leads to assemblages containing
cordierite, anthophyllite, garnet, biotite, andalusite, staurolite,
gahnite, hornblende and plagioclase, depending on the bulk
composition of the altered zones (Franklin et al., 1981).
Cordierite + anthophyllite assemblages commonly result from
metamorphism of chlorite-rich altered zones, with a spotted
texture due to cordierite porphyroblasts, leading to the term
dalmatianite (Fig. 3.7A: Riverin, 1977). Some examples of
mineral assemblages from metamorphosed hydrothermally
altered zones are provided in Table 7.1.
7.3 | THE SPECTRUM OF VOLCANIC-
HOSTED DEPOSITS AND
ASSOCIATED ALTERATION
PATTERNS
Our research into the Palaeozoic VHMS deposits of eastern
Australia (Large, 1992; Gemmell et al., 1998; Gemmell
and Herrmann, 2001; Large et al., 2001c) has revealed
considerable variation in terms of volcanic environment, ore
body shape, metal ratios, metal zonation, alteration mineral
assembalges and zonation, and ratio of massive to stringer
and disseminated styles of ore. Large (1992) identified 10
major styles of deposits in Australia (Fig. 7.13), only one of
which was the classic mound style and associated footwall
alteration pipe depicted in Figure 7.3. A number of factors
including temperature and salinity of the hydrothermal fluid;
oxidation and H
2
S/SO
4
characteristics of the hydrothermal
fluid and seafloor environment; composition of volcanic
and sedimentary rocks deep in the succession; permeability
of the footwall volcanic rocks; and depth of seawater control
the metal carrying capacity of the fluid, and the chemical
reactions that occur beneath and at the seafloor, leading to
a broad range of deposit styles and associated local alteration
halos.
The spectrum of VHMS deposits found in submarine
volcanic successions indicates that there may be continuum
of deposit styles between the end members that form the
basis for the current deposit classification in volcanic arc
and rift environments: VHMS Cu-Zn-Pb, porphyry Cu-Au,
epithermal Au-Ag and SEDEX Zn-Pb-Ag deposits. Recent
workers have emphasised the possible continuum between
VHMS and epithermal deposits (e.g. Lydon, 1996; Sillitoe
et al., 1996). Large (2000, 2004) and Large et al. (2001c)
extended this approach to include porphyry and SEDEX
deposits. Sillitoe et al. (1996) introduced high-sulfidation
and low-sulfidation VHMS deposits based on mineralisation
style, hypogene alteration and sulfide mineralogy, seawater
depth, and volcano-magmatic setting. High-sulfidation
VHMS deposits develop in shallow-water environments,
proximal to volcanic centres, and tend to be associated with
zones containing argillic alteration minerals (e.g. pyrophyllite,
allunite, kaolinite, diaspore) and high-sulfidation sulfide
minerals (e.g. enargite, luzonite, bornite, tennantite). For most
geologists the high-sulfidationlow-sulfidation classification
scheme has proven difficult to embrace because these terms
were originally based on the chemistry of the ore fluid and
environment of mineral deposition, rather than a series of
geological criteria that could be measured and applied in
the field or in drill core. For this reason we do not endorse
adoption of the high-sulfidationlow-sulfidation terminology
for VHMS deposits.
An alternative approach, suggested by Large et al. (2001c)
and expanded here, is to place individual deposits within a
range of features defined for ores in volcanic arcs and rifts. A
diamond-shaped diagram (Fig. 7.14) shows deposits plotted
in terms of their attributes relative to end-member deposit
models for VHMS, epithermal Au, porphyry Cu and SEDEX
Zn-Pb deposits. The main eastern Australian deposits
described in this chapter, and the Bathurst 12 deposit from
the Bathurst mining camp, are plotted on the diagram.
Hellyer plots very close to the ideal VHMS deposit end
member. This is because the deposit exhibits most of the
features of the idealised VHMS alteration-mineralisation
system outlined in Figure 7.3.
Mount Lyell (Western Tharsis, Section 7.7) and Highway-
Reward (Section 7. 10) plot toward the porphyry end of the
spectrum with a significant magmatic component. This is
because these deposits are Cu-Au-rich subsurface replacement
ores that formed in proximal volcanic environments dominated
by synvolcanic rhyolitic intrusions. Mount Lyell also contains
alteration minerals typical of an acid fluid or high-sulfidation
epithermal environment (pyrophyllite and zunyite), which
may suggest that magmatic fluid was involved (Huston and
Kamprad, 2001). These Cu-Au-rich massive sulfide deposits
and others like them (e.g. Mount Morgan, Boliden, Bousquet)
are considered to be hybrid VHMS-epithermal-porphyry
deposits that are not easily classified as end members in the
VHMS-epithermal-porphyry spectrum (Large, 2004).
Henty is a gold-rich, base-metal-poor volcanic-hosted
deposit within the Mount Read province (Section 7.8). The
gold ore occurs in a stratabound subseafloor replacement zone
surrounded by concentric altered zones dominated by quartz,
sericite, carbonate and albite. The deposit is neither a typical
VHMS nor an epithermal deposit, but has some features of
both, and is best described as a hybrid VHMS-epithermal
deposit.
Rosebery is a Zn-Pb-Ag-Au massive sulfide deposit
(Section 7.6). It differs from the classic VHMS deposit in its
low Cu content, sheet-like stratiform nature with no stringer
zone, and lack of a well-defined footwall alteration pipe. It is
hosted in proximal and distal volcanic facies dominated by
pumice breccia, volcanic sandstone and siltstone, and black
shales. Rosebery and other sheet-like Zn-Pb-rich deposits,
such as those in the Bathurst mining camp, have many features
similar to SEDEX deposits even though they are in volcanic
Mattabi Chloritoid + siderite + andalusite Chlorite Franklin et al. (1977)
Mattabi Quartz + chloritoid + andalusite + kyanite Silica core Franklin et al. (1977)
Coronation Cordierite + anthophyllite Chlorite Whitmore (196 9)
Anderson Lake Mg-chlorite + biotite + kyanite Chlorite Franklin etal. (1981)
Amulet A Anthophyllite + cordierite Chlorite Beaty and Taylor (1979)
Geco
B i
J
t e
! I f ?
+
.
chl0
|
r
!
te + muscovite + almandine + cordierite +
Chlorite Stanton (1984)
anthophylnte + staurolite
x
'
Balcooma Chlorite + quartz + staurolite + biotite + cordierite + garnet Chlorite Huston etal. (1992)
Balcooma Quartz + muscovite + biotite Sericite Huston etal. (1992)
Dry River South Quartz + muscovite + biotite + staurolite + andalusite Sericite Huston et al. (1992)
Skellefte district Chlorite + cordierite + andalusite Chlorite Weihed et al. (2000)
Boliden Sericite + quartz + andalusite + corundum Central zone Nilsson (196 8)
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 1 7 5
TABLE 7.1 | Mineral assemblages recorded in medium- to high-grade metamorphosed hydrothermally altered zones local to V HMS deposits.
1 7 6 | CHAPTER 7
FIGURE 7.13 | Schematic representation of
the various shapes and alteration zonation
pattens associated with VHMS deposits
(modified after Large, 1992).
successions. Recognising their hybrid natures and range in
features we have plotted the Rosebery and Brunswick No. 12
deposits on the boundary between VHMS and SEDEX
deposits (Fig. 7.14).
Hydrothermal alteration related to the spectrum of
deposits
Deposits in the VHMS spectrum exhibit a continuum of
alteration zonation patterns that are depicted in Figure 7.15.
The shapes of the alteration halos, their mineral assemblages
and zonation, change progressively along the spectrum.
Porphyry Cu (-Au) deposits (Fig. 7.15A) exhibit a series
of very extensive roughly concentric altered zones. These
include potassic zones in the cores (K-feldspar and/or biotite),
enveloped by phyllic zones (quartz + sericite + pyrite), and
finally propylitic zones (carbonate + chlorite + epidote) at the
margins, which merge with the regional diagenetic or meta-
morphic facies.
Hybrid massive sulfide Cu-Au deposits (Fig. 7.15B) exhibit
similarly shaped, but less extensive, altered hanging wall
zones compared to porphyry Cu deposits. Although they
lack potassic zones, they comprise siliceous and/or chlorite
core zones, which contain Cu-Au ore, and are surrounded
by sericite zones and propylitic halos. Alteration minerals
characteristic of highly acidic alteration (e.g. pyrophyllite,
kaolinite, zunyite, topaz) may occur in the sericite zone in a
similar pattern to porphyry systems.
Classic mound or lens-shaped Cu + Zn Pb-rich VHMS
deposits have both massive sulfide and footwall stringer zones
with well-zoned footwall alteration pipes (Fig. 7.15Q and
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 1 7 7
subordinate hanging wall altered zones, unlike the previous
members of the spectrum where altered rocks dominate the
hanging wall. Mg-bearing altered zones (Mg-chlorite or talc)
are common features of these deposits dependent on the
chemistry and temperature of the fluids.
Sheet style VHMS deposits are Zn-rich, strata parallel
and have extensive stratabound altered zones (Fig. 7.15D).
Stringer zones are less common, but where present they
are stratabound rather than pipe-shaped. Carbonates are
common alteration minerals, particularly around the margins
of the massive sulfide ore.
SEDEX Zn-Pb-Ag deposits are at the distal end of the
massive sulflde spectrum in terms of proximity to volcanic
centres, ratio of sedimentary to volcanic host rocks, and
temperature of formation (Fig. 7.15E). Alteration halos are
KEY ALTERATION MINERALS
chlorite
sericite
carbonate
FIGURE 7.15 | Variations in alteration halos for the spectrum of deposits from
porphyry Cu-Au to SEDEX Zn-Pb-Ag. (A) Classical porphyry Cu-Au deposit (e.g.
El Salvador, after Gustafson and Hunt, 1975; McMillan and Panteleyev, 1998).
(B) Hybrid Cu-Au massive sulfide deposit (e.g. Mount Morgan, Mount Lyell, or
Highway-Reward, after Large et al., 2001c). (C) Classic mound-style VHMS
Cu-Zn or Cu-Zn-Pb deposit (e.g. Hellyer, after Gemmell and Fulton, 2001). (D)
Sheet-style VHMS deposit (e.g. Rosebery, after Large et al., 2001b). (E) Classic
SEDEX deposit (e.g. HYC, after Large et al., 2000).
1 7 8 | CHAPTER 7
commonly stratigraphically controlled and dominated by
carbonate minerals (ferroan-dolomite, ankerite and siderite,
Large et al., 2001c). Chlorite and siliceous zones are rare,
and a sericite zone is usually restricted or entirely absent.
Manganese and Tl are common alteration halo indicators in
SEDEX and sheet style VHMS deposits, but are less common
in the mound-style VHMS, hybrid Cu-Au deposits and
porphyry Cu deposits (Large et al., 2001c).
7.4 | COMPARISONS BETWEEN
ARCHAEAN, PALAEOZOIC AND
CAINOZOIC VHMS ALTERATION
SYSTEMS
Australian Palaeozoic VHMS alteration halos
The altered zones around eastern Australian Palaeozoic
VHMS deposits have diverse morphologies and mineral
assemblages related to variations in their volcanic settings and
modes of formation (Large, 1992; Large et al., 2001c). Well-
defined footwall alteration pipes are relatively uncommon, or
perhaps unrecognised because of subsequent deformation.
In Australian deposits, footwall alteration pipes are mainly
associated with synvolcanic faults (e.g. Mount Morgan, Taube,
1986) and with relatively impermeable footwall rocks (e.g.
Hellyer and Highway-Reward, Gemmell and Large, 1992;
Large, 1992; Doyle, 2001). Laterally extensive stratabound
altered footwall zones are more typical, especially beneath
sheet-like Zn-rich polymetallic deposits, such as Rosebery,
Hercules and Thalanga. Stratabound altered footwall zones
are attributed to non-focussed discharge and lateral migration
of hydrothermal fluids in permeable volcaniclastic units in
the footwall (e.g. Rosebery, Green et al., 1981).
Quartz + sericite + pyrite assemblages are volumetrically
dominant in all types of alteration halos. The proximal
altered footwall zones are typically quartz rich, containing
less than 20% phyllosilicates but greater than 5% pyrite in
disseminations and veins. Much broader footwall feldspar-
destructive altered zones, with mineral assemblages dominated
by sericite or sericite + chlorite, a lower proportion of quartz,
and a few percent of disseminated pyrite, envelop them.
Chlorite-rich assemblages tend to be restricted to small zones
immediately beneath ore lenses (e.g. Rosebery and Thalanga,
Green et al., 1981; Paulick et al., 2001), and in several cases
are associated with carbonate (e.g. Hellyer, Thalanga and
Woodlawn, Davis, 1990; Herrmann and Hill, 2001) . Where
there are footwall alteration pipes, chlorite exists in the medial
zones, usually between a quartz-rich core and a surrounding
sericite zone (e.g. Hellyer and Highway-Reward, Gemmell
and Large, 1992; Doyle, 2001). However, the chlorite-rich
footwall alteration pipes that are characteristic of many
Canadian Archaean deposits do not seem to be present in
the Australian Palaeozoic deposits. Several of the Tasmanian
examples are virtually devoid of chlorite, such as Henty,
Boco and Chester (Boda, 1991; Green and Taheri, 1992;
Callaghan, 2001). These are base-metal-poor systems; some
of them contain aluminous minerals such as pyrophyllite and
kaolinite. Recent interpretations suggest they are analogous to
seafloor acid-sulfate epithermal systems, and possibly involve
significant magmatic fluid in their formation (Large et al.,
2001c).
The sheet-like Zn-rich deposits do not have extensive
visually recognisable hydrothermally altered hanging wall
zones, although there may be subtle hanging wall geochemical
halos (e.g. Rosebery, Large et al., 2001b). However, some
deposits with vertical pipe-like footwall alteration and/or
mineralised zones exhibit altered hanging wall zones that
extend for several tens to hundreds of metres above the
deposit. For example, the 200 m thick basalt unit overlying
the Hellyer deposit contains an upward flaring plume of
distinctive green, Cr-bearing muscovite (fuchsite), which is
more or less concentrically enclosed by discontinuous chlorite
+ carbonate and quartz + albite and sericite zones (Gemmell
and Fulton, 2001). A stratabound quartz + albite ( chlorite)
altered zone up to 100 m thick exists above the pyritic zones at
Henty (Halley and Roberts, 1997). The altered hanging wall
zones at Hellyer and Henty contain only traces of pyrite,
which suggests that they formed from fluids of very different
composition to those in the footwall. In contrast, the quartz
+ sericite + pyrite zones extending above the Highway and
Reward massive sulfide bodies are essentially similar to the
footwall stringer zones. This supports the interpretation that
these deposits formed entirely below the seafloor (Doyle and
Huston, 1999).
Carbonate alteration facies are common features of the
Australian Palaeozoic polymetallic deposits. They are typically
thin stratabound zones that enclose, lie immediately above,
or are laterally equivalent to, the sulfide lenses (e.g. Rosebery,
Henty and Thalanga, respectively). Except at Henty, the
hydrothermal carbonates are generally not calcic; they have
various Ca-Mg-Fe-Mn compositions, which in some cases
vary systematically towards ore (e.g. Rosebery, Large et al.,
2001b).
FIGURE 7.16 | Idealised cross-section of the four altered zones around Kuroko-
type massive sulfide deposits, Japan (modified from Franklin et al., 1981).
Japanese Cainozoic VHMS alteration halos
The Miocene deposits of the Hokuroku district in northern
Japan generally have similar hydrothermal alteration facies to
the Australian deposits, although the transitions to diagenetic
facies in the host succession is clearer because they are not
deformed or metamorphosed. The idealised Kuroko deposit
model has four altered zones surrounding the sulfide deposit
(e.g. Fig. 7.16: Franklin et al., 1981). These include an inner
quartz + sericite zone immediately beneath the sulfide deposit.
The inner quartz + sericite altered zone typically encloses a
quartz + pyrite stockwork zone (Keiko ore) that underlies
the massive sulfide ore (Oko and Kuroko ores) (Ohmoto
and Skinner, 1983) and is analogous to pyritic stringer zones
beneath some Australian deposits. The inner quartz + sericite
zone is laterally surrounded by a sericite + Mg-chlorite +
montmorillonite zone that extends in a thin layer over the
top of the deposit. It is succeeded outward by mixed-layer
clay alteration facies (mineral assemblages of sericite + inter-
layered illite/smectite + chlorite + albite + K-feldspar), which
may extend for up to several kilometres laterally and 200 m
into the hanging wall. The outermost zeolite zone typically
contains relict plagioclase in mineral assemblages progressing
from analcime + montmorillonite + quartz calcite, through
mordenite + montmorillonite + quartz inter-layered
illite/smectite, to the background diagenetic clinoptilolite +
mordenite alteration assemblage.
There is some deposit-specific variation within that
idealised Kuroko alteration zonation pattern. For example, at
the Fukuzawa deposits the sericite + Mg-chlorite zone in the
hanging wall contains relict plagioclase, and analcime seems
to exist in more distal parts of the zeolite zone than mordenite
(Date et al., 1983). Alteration mineral assemblages in the
altered footwall zones around the Uwamuki deposits broadly
conform to the idealised zonation but include peripheral
zones of kaolinite in (presumably disequilibrium) mineral
assemblages of sericite + chlorite + quartz + albite + pyrite
(Urabe et al., 1983). These authors noted that kaolinite is
not otherwise common around Kuroko deposits and, where
present, usually occurs in the core zones with pyrophyllite
+ diaspore. Subsequent work around the Uwamuki deposit
by Shikazono et al. (1998) indicated that the kaolinite
zones are greater than 200 m from ore and extend into the
hanging wall and therefore may not have been directly related
ore deposition. Marumo (1989) also found kaolin minerals
in the hanging wall of the small Inarizawa sulfide deposits
and concluded that they formed during a low-temperature
waning phase of the ore-related hydrothermal system. The
existence of kaolinite pyrophyllite diaspore assemblages,
characteristic of low pH, acid-sulfate systems, suggests that
the Hokuroku district also contains a spectrum of volcanic-
hosted deposits similar to those recently recognised in the
early Palaeozoic belts of Tasmania and north Queensland
(Large et al., 2001c).
Carbonate-bearing assemblages have not been widely
described in the Hokuroku district. Nevertheless, Shikazono
et al. (1998) reported that magnesite, siderite, dolomite and
calcite were common and characteristic in the ore horizon
and hanging wall rocks. Based on isotopic and fluid inclusion
data from Uwamuki they concluded that carbonates
precipitated in a post-ore hydrothermal stage by interaction
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS |
of hydrothermal fluids with biogenic marine carbonates. The
erratic distribution and the post-ore formation of carbonates,
and indications of complex overprinting of different systems
(VHMS and acid-sulfate) leaves some doubt about the
genetic relationships between the massive sulfide deposits and
carbonate assemblages.
Canadian and Australian Archaean VHMS
alteration halos
There are two major classes of altered zones associated with
Canadian Precambrian massive sulfide deposits: (1) well-
defined narrow footwall alteration pipes, and (2) broad
irregular altered footwall zones that are transitional to deep
semi-conformable alteration facies, with or without localised
pipes (Morton and Franklin, 1987; Kerr and Gibson, 1993;
Gibson et al., 1999). The former are commonly associated
with small (<5 Mt) Cu-Zn deposits and are interpreted to have
formed in deep water in dominantly coherent mafic volcanic
rocks. The latter generally exist beneath larger deposits of
variable metal associations formed in relatively shallow water
(^500 m) and in dominantly felsic volcaniclastic rocks.
Pipe-like altered footwall zones are epitomised by the Cu-
Zn deposits of the Noranda district in the Abitibi belt. These
characteristically have upward flaring footwall alteration pipes
that are roughly circular in plan view, generally with slightly
smaller diameter but greater vertical extent than the overlying
massive sulfide lenses. They are commonly recognisable for
up to 1 km below the deposits (Franklin et al., 1981). The
upper part of the footwall alteration pipe (Fig. 7.17) encloses
a stringer zone or stockwork of pyrite chalcopyrite
pyrrhotite veins in a core dominated by Fe-chlorite passing
laterally and upward through Mg-chlorite to an outer zone
dominated by sericite chlorite quartz (Lydon, 1984; Kerr
and Gibson, 1993). Lydon (1996) noted the existence of talc-
bearing or aluminous assemblages in the upper parts of some
footwall alteration pipes.
Depletions of Si, Na, Ca and K and additions of Mg and
Fe generally characterise the alteration of the chloritic core,
1 7 9
FIGURE 7.17 | Idealised cross-section of a typical zoned footwall alteration
pipe beneath Noranda-type massive sulfide deposits, Abitibi belt, Canada
(modified from Lydon, 1984; Kerr and Gibson, 1993).
1 8 0 | CHAPTER 7
whereas small additions of K and possibly Si occur in the
sericitic shell (Barrett and MacLean, 1994b). This generally
results in significant net loss of mass, due to Si loss, from
the overall alteration pipe (Barrett and MacLean, 1991). In
a few unusual cases there may be net mass gains (e.g. Norbec
deposit: Barrett and MacLean, 1999). Footwall alteration
pipes of this type may represent zones of hydrothermal
discharge that were focussed by synvolcanic faults. Their
vertical extent suggests that the hydrothermal fluid sources
were very deep. Overprinting relationships indicate that the
footwall alteration pipes were initially zones of sericite
quartz altered rock. As the hydrothermal system intensified,
sericite was replaced by chlorite concurrent with metal zone
refining in the sulfide lenses (Kerr and Gibson, 1993).
Non pipe-like, broad altered footwall zones have more
variable morphologies and mineral assemblages as exemplified
by the differences in the Home and Mattabi deposits. The
Home deposit has a poorly defined altered footwall zone of
quartz + sericite chlorite that is many times wider than the
massive sulfide bodies (MacLean and Hoy, 1991; Kerr and
Gibson, 1993). Calculations by MacLean and Hoy (1991)
indicate that the Home footwall alteration was accompanied
by significant net mass gains; mainly gains of Si, Fe and
K, slightly offset by losses of Na, Ca and Mg. Beneath the
Mattabi deposit are siderite + chloritoid andalusite, kyanite
and pyrophyllite zones, which narrow with depth and are
transitional downward and laterally into an extensive semi-
conformable ankerite + chlorite + sericite + quartz zone
(Franklin et al., 1975; Morton and Franklin, 1987). Gibson
et al. (1999) suggested that the aluminous assemblages at
Mattabi (and several other deposits that are notably Au rich)
were analogous to the advanced argillic assemblages formed
by low pH fluids in acid-sulfate epithermal systems.
The Archaean massive sulfide deposits of the Panorama
district in the Pilbara of Western Australia occur near the top
of a 2 km thick basaltic to rhyolitic volcanic succession above
a large synvolcanic granite pluton (Brauhart et al., 2001).
Large, semi-conformable altered zones of feldspar-destructive
sericite + quartz and chlorite + quartz alteration assemblages
occupy the lower and middle parts of the volcanic succession
and extend almost the entire exposed strike length (20 km).
Locally transgressive chlorite + quartz altered zones, bounded
by synvolcanic faults, extend upwards from the semi-
conformable altered zones to beneath the massive sulfide
prospects. Mass changes in the feldspar-destructive altered
zones were modest: small gains of Si and losses Ca, Na, Fe
and K in the lower sericite + quartz zones, and small gains of
Mg, Fe, Si and losses of K, Na, Ca in the transgressive chlorite
+ quartz zones.
In the Golden Grove district of the Archaean Yilgarn
craton, Western Australia, the volcanic succession that hosts
the Scuddles and Gossan Hill massive sulfide deposits also
exhibits the effects of regional-scale, intense feldspar-destructive
alteration. The entire footwall succession of altered andesitic
to rhyolitic volcaniclastic units, although preserving primary
volcanic textures, is composed essentially of quartz + chlorite
( minor sericite). The alteration process, interpreted as a
syn-depositional or early hydrothermal regional metasomatic
event, virtually removed all Ca, Na and K from the rocks
and added substantial Si, Fe and Mg (Sharpe and Gemmell,
2001). At Gossan Hill the regional quartz + chlorite alteration
facies is overprinted by two alteration facies related to sulfide
mineralisation. A narrow stratabound chlorite ( siderite,
ankerite talc and andalusite) zone envelops the lower Cu-rich
massive magnetite + sulfide ore lens. An intense quartz zone
underlies the upper Zn-rich massive sulfide lens and encloses
a discordant zone of sulfide stringer veins that connects the
upper and lower lenses (Sharpe and Gemmell, 2001).
There is considerable diversity among alteration facies
aroundArchaean massive sulfide deposits. Features that appear
to be common to most Archaean districts are large semi-
conformable altered zones and localised discordant altered
footwall zones. Brauhart et al. (2001) highlighted some of the
differences in mineral assemblages and mass changes between
Panorama and the Canadian semi-conformable altered zones.
However, the well-defined discordant footwall alteration
pipes are typically chlorite rich (if not metamorphosed to
higher grades) and characterised by significant net mass loss,
which is attributable to major Si loss and only partly offset by
Mg and Fe gains.
Comparisons
Despite the many variations in mineral assemblage, morphology
and extent of alteration facies associated with VHMS deposits,
both within districts and across geologic time, there is one
feature that is common to all: proximal altered footwall zones
do not contain feldspar. Feldspar destruction is usually manifest
in the presence of sericite, chlorite or smectite clays, or their
higher grade metamorphic equivalents. One or more of these
Al-bearing phyllosilicates is almost invariably present because,
although VHMS-type hydrothermal fluids readily transport
silica, alkalis and other cations, Al is generally immobile in
these moderately acidic systems. One suspects that pyrite is an
equally ubiquitous component of proximal altered footwall
zones but it is frequently not mentioned in alteration mineral
assemblages, due to the unnatural distinctions that many
authors make between alteration and mineralisation. It is also
becoming increasingly apparent that alumino-silicates such
as kaolinite, pyrophyllite, andalusite and others exist locally
in altered zones across the entire age spectrum of VHMS
deposits, from Archaean to Cainozoic, and that there may
always have been continua between moderate and low pH
submarine hydrothermal systems. Because of the consistent
feldspar destruction, alteration indices such as Al (Ishikawa
et al., 1976), CCPI and S/Na
2
O (Large et al., 2001a) should
be effective indicators of alteration intensity in all kinds of
VHMS districts, at least at prospect scales.
Footwall alteration facies of the Australian Palaeozoic and
Japanese Cainozoic deposits are typically more sericitic than
chloritic, and their proximal zones tend to be quartz rich.
The limited mass change data available indicate they have
typically undergone moderate to large net mass gains, which
are dominantly attributable to gains in Si that significantly
outweighed losses and gains in other components. This
predominance of sericite is probably not due to low availability
of Fe and Mg. Iron may be significantly added as pyrite and
Mg commonly appears in carbonates or chlorite in upper or
upper-peripheral altered footwall zones.
In contrast, chlorite dominates the well-defined footwall
alteration pipes that underlie many Canadian Archaean
deposits. These zones are characterised by significant net mass
loss, in which the large loss of Si outweighed addition of Mg
and Fe. Although there are some exceptions (e.g. Panorama),
major Si and net mass losses are indicated wherever chlorite
is dominant in an alteration mineral assemblage. This
generalisation also applies to Australian Palaeozoic systems;
for instance the small chlorite-rich zones in the Hellyer
alteration pipe (Gemmell and Large, 1992) and Thalanga
footwall (Herrmann and Hill, 2001).
In terms of mass change, the major difference between
Archaean deposits with chloride footwall alteration pipes,
and Palaeozoic to Cainozoic deposits with quartz + sericite-
dominated altered footwall zones, is that the former lost mass
and the latter gained mass. In addition, in all cases, the major
contributor to net mass change was Si.
This difference in the behaviour of Si is probably related
to the evolution of the hydrothermal systems and particularly
the compositions of hydrothermal fluids, which originated
as seawater in both cases. Evidence of chlorite overprinting
sericite quartz assemblages in the Canadian footwall
alteration pipes suggests that fluid compositions changed as
the hydrothermal system intensified. The initial fluid was
probably over-saturated in Si and deposited quartz along
the discharge zone to the seafloor as it cooled, whereas the
later fluid, possibly of higher temperature and associated with
Cu enrichment of the sulfide deposit, was undersaturated
and leached Si from the core of the discharge zone. This
change may be explained in terms of the regional deep semi-
conformable altered zones associated with Archaean deposits.
The lower semi-conformable altered zone is typically a zone
of silicification at temperatures greater than about 400C
(Kennedy, 1950; MacGeehan, 1978; Fournier, 1985; Galley,
1993; Skirrow and Franklin, 1994), attributed to down-going
modified seawater being heated up to the range of retrograde
Si solubility at 400-600C (Fournier, 1985). If fluid deposited
Si in the deep semi-conformable altered zone, it would then
be undersaturated in Si as it ascended the discharge zone,
even if it cooled as it ascended.
The link between Archaean deposits and regional semi-
conformable altered zones, which are generally not recognised
in the younger VHMS districts, suggests that Archaean
crustal conditions (thin crust and large high-level plutonic
intrusions) favoured large, intense and presumably long-lived
systems. Palaeozoic and younger VHMS districts are not
typically associated with large high-level plutons analogous
to the Flavrian pluton of the Noranda district (Kerr and
Gibson, 1993) or the Strelley granite at Panorama (Brauhart
et al., 2001). Their absence may account for the less extensive,
perhaps less evolved, altered footwall zones associated with
Si and net mass gains that are most common beneath the
Palaeozoic and younger massive sulfide deposits.
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 1 8 1
7.5 | HELLYER: A MASSIVE ELONGATE
POLYMETALLIC LENS
The Hellyer deposit is located in the northern part of the
Cambrian Mount Read province, western Tasmania (Fig.
1.5). The pre-mining resource was 16.2 Mt of 13.9% Zn,
7.1% Pb, 0.4% Cu, 168 g/t Ag and 2.5 g/t Au (Gemmell and
Large, 1992; McArthur, 1996). The deposit is a single elongate
lens of massive sulfide about 800 m in length (northsouth)
by 200 m in width (eastwest) and with an average vertical
thickness of 45 m (Fig. 7.18). It occurs in the mainly calc-
alkaline, intermediate to mafic Que-Hellyer Volcanics at
the base of the Mount Charter Group, which is equivalent
to the western volcano-sedimentary sequences (Corbett and
Sericite + quartz zone
Sericite zone
Chlorite zone
Quartz zone
FIGURE 7.18 | Hellyer plan showing the altered zones immediately below the
massive sulfide ore (approximately 400 RL). The black line is the outline of the
base of the massive sulfide. Modified after Gemmell and Large (1992).
1 8 2 | CHAPTER 7
Komyshan, 1989). The massive sulflde lens is bisected and
offset by a major north-south trending fault, the Jack fault
(Figs 7.18 and 7.19). Beneath the massive sulflde lens is
an elongate, carrot-shaped, zoned, footwall alteration pipe
(described in detail by Gemmell and Large, 1992). Above the
deposit is a moderately well developed altered hanging wall
zone (Gemmell and Fulton, 2001).
The massive sulflde lens exhibits classical metal zoning,
with minor Cu concentrated in a pyritic core, followed by
low grade Zn + Pb, and then high grade Zn + Pb + Ag in
the upper parts of the massive sulflde lens (McArthur and
Dronseika, 1990; Large, 1992; McArthur, 1996). The centre
of the massive sulflde deposit is capped by a quartz + pyrite
zone, which is flanked by thin irregular lenses of barite that
FIGURE 7.19 | Cross-section of the Hellyer deposit showing the distribution
of rock types, mineralised zones, and altered footwall and hanging wall zones.
(A) The ore lens and altered zones are offset along the Jack fault (modified after
Gemmell and Large, 1993). (B) Reconstructed 10740 N/10870 N cross-section
showing the massive sulfide and footwall alteration pipe prior to folding and
faulting (modified after Downs, 1993).
are directly over the high-grade massive sulflde (Sharpe,
1991). Barite and massive sulfide clasts occur in volcaniclastic
mass-flow units flanking the deposit.
Geological setting
The Hellyer ore body occurs above a footwall comprising
feldspar-phyric andesitic and basaltic lavas and sills that
consist of coherent and autoclastic facies, which are mainly
hyaloclastite and peperite (Fig. 7.20: Waters and Wallace,
1992). Basalt (Hellyer basalt) and black mudstone (Que
River Shale) dominate the hanging wall (Komyshan, 1986).
The abundance of basalt-mudstone peperite indicates that
most of the basalt units are sills that intruded the black
mudstone (McPhie and Allen, 1992; Waters and Wallace,
1992). Very thick, graded quartz-bearing rhyolitic pumiceous
and volcanic lithic breccias interbedded with turbidites and
mudstones of the Southwell Subgroup occur in the upper
parts of the hanging wall (Corbett, 1992; McPhie and Allen,
1992). The ore lens position is marked by a 040 m thick
volcaniclastic unit, which mainly consists of coarse polymictic
volcanic breccia, sandstone and laminated volcanic siltstone
(Waters and Wallace, 1992).
Trilobites in the Que River Shale, very thick sections of
black mudstone and the abundance of graded mass-flow units
collectively indicate that the Hellyer massive sulfide formed in
a moderate to deep (>1000 m) submarine setting (Large et al.,
2001a). The volcanic facies association indicates proximity to
intrabasinal vents for effusive, basaltic and andesitic eruptions
and synvolcanic intrusions.
Alteration facies and zonation
Gemmell and Large (1992), and Gemmell and Fulton
(2001) provided detailed description of both the footwall and
hanging wall alteration facies, and zonation at Hellyer. The
following section summarises their work.
Footwall alteration facies and zonation
A zoned carrot-shaped footwall alteration pipe extends for
at least 500 m beneath the Hellyer deposit (Figs 7.6C and
7.20). At the centre of the alteration pipe, immediately below
the massive sulfide lens, is a siliceous core zone dominated by
intense, pervasive quartz + sericite + pyrite alteration facies.
This zone is progressively enclosed in chlorite, sericite and
stringer envelope (or sericite + quartz) altered zones (Fig.
7.6C).
The moderate, selective sericite + quartz alteration facies
(e.g. data sheet HE2 in the stringer envelope zone) is the
10-50 m wide outermost part of the alteration pipe and
grades outward into weak, selective-pervasive albite + chlorite
alteration facies (least-altered footwall, data sheet HE1).
Primary volcanic textures are preserved (although modified),
lithic fragments exhibit sericite-altered margins, and feldspar
phenocrysts are partly altered to sericite. The AI shows an
increase from background values of around 3055 to values
of 60-70 (Fig. 2.8B).
intense footwall
v
" alteration plume
V V / V
V ,i I V n V
L OCAL HYDROTHE RMAL AL TE RATI ON RELATED TO V HMS DE POSI TS | 1 8 3
Southwell Subgroup: crystal-rich
volcaniclastic shale, greywacke and minor
felsic lava
Que River Shale: black shale and minor
sandstone
Hellyer basalt: massive to pillowed basalt,
pillow breccia, hyaloclastite and andesitic
lava
Mixed sequence: polymict volcanic breccia,
massive and auto-brecciated dacite and
massive sulfide ore
Lower andesites and basalts: andesitic,
dacitic and basaltic lavas, hyaloclastites
and minor volcani-clastic facies
F I G URE 7.20 | Schematic
stratigraphic section through the
Que-Hellyer V olcanics showing the
Lower basalt: massive to pillowed basalts,
ext ent of a
|
t er ed z ones at t he He
| |
yer
hyaloclastites and pillow breccias , ~
n
. , . . . ..,. .
' and Que River deposits. Modified
after Waters and Wallace (1 992 ).
Sericite chlorite dominates the strong, selective-
pervasive alteration facies in a 1015 m wide zone marking
the outer extent of an intense hydro thermal alteration system,
recognised by obliteration of volcanic textures, presence of
minor sulfides (mainly pyrite) and complete replacement of
feldspar phenocrysts and feldspathic groundmass by sericite
chlorite (e.g. data sheet HE3). The AI is typically between 70
and 85. On the Alteration box plot samples from this facies
plot along a line from the least-altered footwall field toward
the chlorite corner.
In the intense, pervasive chlorite alteration facies, all
primary minerals and glass in the footwall andesitic rocks
are completely replaced by fine-grained chlorite with minor
pyrite, sericite, quartz and carbonate (e.g. data sheet HE4).
The AI is between 90 and 100 and the CCPI between 80 and
90. In the upper parts, immediately below the massive sulfide,
this zone includes an intense, spheroidal chlorite + carbonate
alteration facies (Fig. 7.6C), which has up to 50% dolomite
in a fine-grained matrix of chlorite (e.g. data sheet HE5).
This alteration facies has a lower AI (5080) than the intense
chlorite alteration facies due to the elevated whole-rock CaO
related to the dolomite component in the rock.
In the siliceous core zone, all volcanic textures are
completely destroyed and the rock is composed of a fine
intergrowth of quartz + sericite + pyrite + chlorite (intense,
pervasive quartz + sericite + pyrite alteration facies, data sheet
HE6). This zone also contains a series of sub-vertical pyrite +
quartz + sphalerite + galena chalcopyrite carbonate barite
veins, interpreted as hydrothermal feeders below the ore body.
Alteration indices in the siliceous core zone are extremely high
with values of both AI and CCPI exceeding 90.
Hanging wall alteration facies and zonation \
Hanging wall alteration facies at Hellyer are less well
developed than the footwall alteration facies; however,
recent detailed studies by Gemmell and Fulton (2001) have
recognised an upward flaring zoned alteration system that is
centred above the thickest part of the massive sulfide lens. The
altered hanging wall zone extends through the hanging wall
pillow basalts up to the contact with the overlying Que River
Shale (Fig. 7.19). Data sheet HE7 is an example of the least-
altered hanging wall andesite. The distribution and intensity
of alteration facies in the altered hanging wall zone is patchy,
with pillow margins more intensely and pervasively altered
(e.g. data sheet HE9) than the pillow interiors. The outer
margin of the altered hanging wall zone is defined by weak
sericite alteration facies grading inwards to a pink-white,
strong, pervasive albite alteration facies (e.g. data sheet HE8),
moderate, pervasive chlorite + carbonate alteration facies (e.g.
data sheet HE9), and in the centre of the system, a distinctive
green, strong, pervasive fuchsite alteration facies (e.g. data
sheet HE 10). There is no systematic trend in the alteration
indices in the altered hanging wall zones. AI and CCPI values
are commonly low in the albite alteration facies due to high
Na
2
O, and low MgO and FeO values (e.g. data sheet HE9).
Ore genesis
Based on geological, textural, and metal zonation studies,
McArthur (1989, 1996), Large (1992) and Gemmell and
Large (1992) concluded that the Hellyer massive sulfide
deposit grew as a mound in a seafloor depression. The metal
zonation from Fe - Cu Pb-Zn > Ba was interpreted to
be an expression of hydrothermal zone refining (Large, 1992),
which developed similarly to that described by Eldridge et
al. (1983) for the Kuroko deposits. Solomon and Khin Zaw
(1999), however, presented fluid inclusion data (indicating
high ore fluid salinities: averaging 11 wt%) to propose that
sulfide deposition occurred in a seafloor depression brine-
pool, directly above the footwall alteration pipe. Solomon
and Gaspar (2001) provide textural evidence in support of
sulfide accumulation in a brine pool. Solomon and Groves
(1994) and Solomon et al. (2004) suggest that the abnormally
high salinity and other chemical characteristics of the Hellyer
fluid inclusions, are strongly suggestive of involvement of
magmatic fluids in the hydrothermal system.
1 8 4 [ CHAPTER 7
Weak, selective-pervasive albite + chlorite alteration facies
Least-altered footwall
H E1
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict mineralogy
Relict texture
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
Hfw-LAA(FPS-I)
weak, selective-pervasive albite + chlorite
footwall
Que-HellyerVolcanics
Mount Read Volcanics
monomictic mafic breccia
feldspar
feldspar phenocrysts, perlitic fractures and
curvi-angular cm-scale clasts, areas of
jigsaw-fit clasts
andesite
massive, matrix supported and poorly
sorted
andesitic hyaloclastite
albite + chlorite + sericite + calcite
selective-pervasive chlorite + calcite +
sericite in matrix and chlorite + sericite in
clasts, chlorite infill in perlitic fractures,
sericite + calcite-altered plagioclase
regional
good
weak
syn volcanic?
diagenetic
Hand specimen photograph Photomicrograph (ppl)
Geochemistry
SiO
2
54.6 9 K
2
O 1.55 Rb 6 8 Zr 125
TiO
2
0.6 4 P
2
O
5
0.12 Sr 299 Nb 7.0
AI
2
O
3
17.92 S 1.92 Ba 500 Y 19
Fe
2
O
3
7.6 5 CO
2
Cu 0
MnO 0.09 Total 97.96 Pb 0 Al 36
MgO 3.79 LOI 3.96 Zn 0 CCPI 57
CaO 3.14 Sb Ti/Zr 30.9
Na
2
O 6 .45 Tl
Moderate, selective sericite + quartz alteration facies
Footwall
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 1 8 5
HE 2
I
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
Hfw-SEZ
moderate, selective sericite + quartz
footwall
Que-Hellyer Volcanics
Mount Read Volcanics
polymictic mafic breccia
feldspar
porphyritic and perlitic clasts, areas of
jigsaw-fit clasts, clasts with curviplanar
margins
andesite-basalt
massive, clast supported, poorly sorted
resedimented polymictic hyaloclastite
sericite + chlorite + quartz + albite + calcite
+ pyrite
selective domainal, calcite vein infill,
disseminated pyrite, albite + sericite +
calcite-altered feldspars
alteration zone around pipe
moderate
moderate
synmineralisation
peripheral hydrothermal
Geochemistry
SiO
2
57.38 K
2
O 3.34 Rb 127 Zr 123
TiO
2
0.58 P
2
O
5
0.13 Sr 97 Nb 7.0
AI
2
O
3
15.33 S 4.15 Ba 6 700 Y 26
Fe
2
O
3
7.76 CO
2
Cu 300
MnO 0.10 Total 96 .27 Pb 2700 Al 6 3
MgO 3.51 LOI 7.89 Zn 4500 CCPI 6 9
CaO 2.71 Sb Ti/Zr 28.3
Na
2
O 1.29 Tl
Hand specimen photograph Photomicrograph (xn)
1 8 6 | CHAPTER 7
Strong, selective-pervasive sericite + chlorite alteration facies
Footwall
HE 3
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
Hfw-SZ
strong, selective-pervasive sericite +
chlorite
footwall
Que-Hellyer V olcanics
Mount Read V olcanics
massive polymictic breccia
rare feldspar
deformed feldspar phenocrysts and
andesite clasts
dacite-basalt
massive, matrix supported, poorly sorted
resedimented polymictic hyaloclastite
sericite + chlorite + quartz + pyrite +
ankerite + (albite)
selective pervasive, vein-halo (pyrite etc.),
disseminated pyrite, and infill (carbonate)
pipe
poor
strong
synmineralisation
hydrothermal
Geochemistry
SiO
2
54.50 P
2
O
5
0.10 Cu 500 Al 79
TiO
2
0.52 S 6 .02 Pb 5100 CCPI 77
AI
2
O
3
14.51 CO
2
Zn 7800 Ti/Zr 33.4
Fe
2
O
3
10.93 Total 97.92 Sb
MnO 0.21 LOI 8.11 Tl
MgO 5.16 Zr 94
CaO 1.47 Nb 6 .0
Na
2
O 0.90 ]]' Y 18
K
?
3
'
5 8
Ba 4100
Hand specimen photograph Photomicrograph (xn)
Intense, pervasive chlorite alteration facies
Footwall
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 1 8 7
HE 4
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
Hfw-CLZ
intense, pervasive chlorite
footwall
Que-Hellyer Volcanics
Mount Read Volcanics
coherent feldspar-phyric andesite
nil
porphyritic, perlitic fractures
andesite
massive
indeterminate
chlorite + pyrite + (quartz + sericite +
calcite + galena)
pervasive, chlorite infill in perlitic fractures,
chlorite + quartz-altered feldspars
pipe
moderate
intense
synmineralisation
hydrothermal
Geochemistry
SiO
2
37.6 9 K
2
O 1.82 Rb 79 Zr 140
TiO
2
0.59 P
2
O
5
0.12 Sr 39 Nb 8.0
AI
2
O
3
16 .08 S 8.6 5 Ba 2000 Y 25
Fe
2
O
3
18.86 CO
2
Cu 400
MnO 0.41 Total 96 .35 Pb 5200 Al 95
MgO 11.38 LOI 12.29 Zn 9300 CCPI 94
CaO 0.6 4 Sb Ti/Zr 25.2
Na
2
O 0.12 Tl
Hand specimen photograph Photomicrograph (xn)
CHAPTER 7
Intense, spheroidal chlorite + carbonate alteration facies
Footwall
HE 5
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
135756
intense, spheroidal chlorite +carbonate
footwall
Que-Hellyer Volcanics
Mount Read Volcanics
coherent, feldspar-phyric andesite
nil
porphyritic
andesite
massive
indeterminate
chlorite + dolomite + (quartz + sericite)
nodules-spheroids
local
poor
intense
synmineralisation
hydrothermal
Geochemistry
SiO
2
34.88 P
2
O
5
0.13 Cu 300 Al 75
TiO
2
0.6 0 S 4.13 Pb 7200 CCPI 93
AI
2
O
3
15.6 3 CO
2
Zn 11500 Ti/Zr 29.0
Fe2O
3
12.89 Total 88.72 Sb
MnO 0.80 LOI 15.40 Tl
MgO 13.07 Zr 124
CaO 4.86 Rb 6 6
N b 9 0
Na
2
O 0.01
Sr
58 1 28
K
2
O 1.73
Ba
1200
Hand specimen photograph Photomicrograph (xn)
Intense, pervasive quartz + sericite + pyrite alteration facies
Footwall
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS I 1 8 9
HE 6
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
Hfw-SCZ
intense, pervasive quartz + sericite + pyrite
footwall
Que-Hellyer Volcanics
Mount Read Volcanics
feldspar-phyric andesite breccia
nil
porphyritic, periitic fractures, jigsaw-fit
clasts
andesite
massive
in situ hyaloclastite
quartz + sericite + pyrite + (chlorite)
pervasive, pseudomorphs of feldspar
pyroxene?, quartz and sericite veins
disseminated pyrite, feldspar overgrowths
on phenocrysts
core of pipe
moderate
intense
synmineralisation
hydrothermal
Geochemistry Geochemistry
SiO
2
6 7.42 K
2
O 2.17 Sr 15 Nb 4.0
TiO
2
0.30 P
2
O
5
0.05 Ba 14800 Y 28
AI
2
O
3
8.13 S 8.87 Cu 2200
Fe
2
O
3
12.34 CO
2
Pb 8000 Al 91
MnO 0.07 Total 101.12 Zn 9800 CCPI 85
MgO 1.42 LOI 7.74 Sb Ti/Zr 24.7
CaO 0.32 Tl
Na
2
O 0.03 Rb 79 Zr 74
Hand specimen photograph Photomicrograph (xn)
1 9 0 | CHAPTER 7
Subtle, pervasive aibite + chlorite + calcite alteration facies
Ha n g i n g w a ll
HE 7
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
142562
subtle, pervasive aibite + chlorite + calcite
hanging wall
Que-HellyerV olcanics
Mount Read V olcanics
massive, feldspar-phyric amygdaloidal
andesite
plagioclase
massive, porphyritic, weakly amygdaloidal
andesite
massive
lava
chlorite + aibite + calcite + quartz +
(chalcopyrite)
selective-pervasive, chlorite or quartz infill
in amygdales, quartz + calcite veins
regional
good
weak
synmineralisation
diagenetic to metamorphic
Geochemistry
SiO
2
51.14 K
2
O 0.42 Rb 21 Zr 151
TiO
2
0.55 P
2
O
5
0.6 0 Sr 125 Nb 7.1
AI
2
O
3
14.77 S 0.06 Ba 226 Y 21
Fe
2
O
3
8.82 CO
2
4.76 Cu 745
MnO 0.21 Total 99.91 Pb 6 Al 38
MgO 5.56 LOI 8.17 Zn 147 CCPI 75
CaO 5.44 Sb 1.1 Ti/Zr 21.8
N a, 0 4.19 TI <0.5
Hand specimen photograph Photomicrograph (xn)
Strong, pervasive albite alteration facies
Hanging wall
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 1 91
HE 8
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
1426 22
strong, pervasive albite
hanging wall
Que-Hellyer Volcanics
Mount Read Volcanics
feldspar-phyric basalt breccia
altered plagioclase
porphyritic
basalt
massive and jigsaw-fit breccia
lava or sill
albite + chlorite + calcite + (pyrite)
pervasive, albite + chlorite, massive
chlorite + pyrite veins, patchy
calcite domains possibly irregular
pseudomorphs
local, plume?
poor
strong
post mineralisation
diagenetic-hydrothermal?
Hand specimen photograph Photomicrograph (xn)
1 92 I CHAPTER 7
Moderate, pervasive chlorite + carbonate alteration facies
Hanging wall
HE 9
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
142593
moderate, pervasive chlorite + carbonate
hanging wall
Que-Hellyer Volcanics
Mount Read Volcanics
monomictic basaltic andesite breccia
altered feldspars
porphyritic, jigsaw-fit clasts
basaltic andesite
massive
pillow lava
chlorite + caicite + sericite + (albite +
quartz)
pervasive, spheroidal and rhombic caicite,
caicite veins with chlorite vein-halo
alteration
local, plume?
good
moderate
post mineralisation
hydrothermal
Geochemistry Geochemistry
SiO
2
42.34 K
2
O 1.92 Rb 55 Zr 140
TiO
2
0.6 1 P
2
O
5
0.35 Sr 281 Nb 8.8
AI
2
O
3
11.78 S 0.07 Ba 737 Y 21
Fe
2
O
3
5.41 CO
2
13.00 Cu 109
MnO 0.12 Total 99.46 Pb 4 Al 28
MgO 4.34 LOI 15.99 Zn 51 CCPI 76
CaO 15.44 Sb 0.6 Ti/Zr 26 .2
Na
2
O 1.06 Tl 0.7
Hand specimen photograph Photomicrograph (xn)
Strong, pervasive fuchsite alteration facies
Hanging wall
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 1 9 3
HE 10
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
142643
strong, pervasive fuchsite
hanging wall
Que-Hellyer V olcanics
Mount Read V olcanics
massive feldspar-phyric basalt
feldspar
porphyritic, minor amygdales
basaltic andesite
massive
lava or sill
sericite (fuchsite) + calcite + ankerite +
chlorite + pyrite
pervasive, disseminated pyrite, sericite
cleavage, calcite sericite pseudomorphs
after feldspar
plume
poor
strong
post mineralisation
hydrothermal
Geochemistry
SiO
2
TiO
2
AIA
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
26 .32
0.54
17.35
3.38
0.25
1.88
22.6 7
0.00
K
2
O
S
CO
2
Total
LOI
5.00 Rb
0.21 Sr
0.08 Ba
19.48 Cu
99.15 Pb
21.13 Zn
Sb
Tl
143 Zr
176 Nb
4473 Y
16 2
2 Al
43 CCPI
8.1 Ti/Zr
18.3
74
4.9
18
23
50
43.9
Hand specimen photograph Photomicrograph (ppl)
1 9 4 | CHAPTER 7
7.6 | ROSEBERY: A POLYMETALLIC
SHEET-STYLE DEPOSIT
The Rosebery massive sulfide deposit is a sheet-style
polymetallic Zn-Pb-Cu-Ag-Au VHMS deposit in the northern
Central Volcanic Complex of the Mount Read Volcanics,
western Tasmania (Fig. 1.5: Green et al., 1981; Large, 1992).
The mining resource is 32 Mt at 14.7% Zn, 4.5% Pb, 0.6%
Cu, 146 g/t Ag and 2.3 g/t Au (data from Pasminco Mining
and Exploration). Compared with the Hellyer deposit, which
comprises a single ore lens (described in the previous section),
Rosebery is composed of at least 16 separate ore lenses (Fig.
7.21). These vary in size from 0.1 to 5 Mt. Unlike the carrot-
shaped footwall alteration pipe at Hellyer, the Rosebery
ore lenses are enclosed in strata-parallel altered zones. The
ore lenses are principally composed of massive and banded
sulfides of sphalerite, galena, barite, pyrite and chalcopyrite,
with minor tetrahedrite-tennantite, arsenopyrite, pyrrhotite,
hematite and magnetite. In some sections of the mine (e.g.
A and B lenses, Huston and Large, 1987) barite-rich lenses
overlie the Zn-Pb-Cu ore lenses.
Geological setting
The Rosebery, Hercules and South Hercules polymetallic
ore bodies are hosted by the same stratigraphic sequence in
the upper part of the Central Volcanic Complex, west of
the Henty fault (Solomon, 1964; Green et al., 1981). The
footwall comprises a thick (up to 500 m), poorly stratified
rhyolitic-dacitic succession of weakly graded, feldspar-phyric
pumice breccia, which is interpreted to be the product of
large volume submarine caldera-forming eruptions, and
rhyolitic and dacitic sills (Hercules Pumice Formation: Lees,
1987; Allen, 1994b; Large et al., 2001b). The ore lenses occur
in the 5 to 10 m thick, finely stratified pumiceous siltstone,
sandstone, crystal-rich sandstone and claystone top (host
rocks) of the footwall pumice breccias (Lees, 1987; Corbett
and Solomon, 1989; Allen, 1994b; Large et al., 2001b). The
host rocks are overlain by black mudstone, which represents
a hiatus in volcanism marked by non-volcanic sedimentation.
The hanging wall comprises a 5 to 400 m thick succession
of massive to stratified, feldspar + quartz-phyric rhyodacitic
volcaniclastic units of the White Spur Formation, interbedded
with black mudstone (Lees, 1987; Allen, 1994b). The footwall
rhyolitic pumice breccias haveTi/Zr of 7-9 (Fig. 4.5), whereas
the host interval porphyry sill has Ti/Zr of 12-14, and the
volcaniclastic facies 10-30 (Large et al., 2001b).
Bedforms and textures within the footwall pumice
breccias and host rocks are consistent with deposition from
volcaniclastic turbidity currents, debris flows and suspension
in a below-wave-base environment (McPhie and Allen, 1992).
The footwall pumice breccias are interpreted to represent the
submarine deposits from a large, felsic explosive eruption
(Allen, 1994a). The host rocks may have been derived
from water-settled suspension sedimentation or the influx
of volcaniclastic turbidites from distal rhyolitic volcanic
centres (Large et al., 2001b). The hanging wall volcaniclastic
units probably comprise the medial to distal facies from an
extrabasinal felsic volcanic centre (Allen, 1994a).
A below-wave-base submarine setting for the Rosebery-
Hercules succession is indicated by the presence of sedimentary
structures in the footwall pumice breccias consistent with
deposition from cold water-supported gravity flows and
water-settled fall, rare intercalated black pyritic mudstone and
the associated VHMS deposits (Gifkins and Allen, 2002).
FIGURE 7.21 | Long-section of the Rosebery mine, western Tasmania, showing the drives and main ore lenses, labelled alphabetically (provided by
Zinifex Rosebery mine, 2004). K lens is at depth at the north end of the mine.
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS I 1 9 5
Alteration facies and zonation
Four strata-parallel altered zones enclose the Rosebery ore
lenses. From the periphery to the core of the alteration system
these are: sericite zone, chlorite zone, Mn-carbonate zone,
and quartz + sericite zone (Fig. 7.5C: Large et al., 2001b).
Outside the altered footwall zone, least-altered rhyolitic
volcanic rocks contain plagioclase crystals in a pumice- and
shard-rich matrix (e.g. data sheet RBI). This matrix is weakly
sericite + chlorite + quartz altered, commonly enhancing the
shard and pumice textures. Plagioclase crystals are weakly
altered with disseminated fine-grained sericite and albite
overgrowths. These rocks have AI = 3060 and CCPI = 15
40 and plot in the least-altered box of the Alteration box plot.
Data sheet RB2 is an example of the least-altered hanging wall
rocks.
The outer part of the hydrothermal alteration system is a
broad sericite zone with scattered Mn-carbonate blebs, which
extends up to 300 m into the footwall, but less than 25 m
into the hanging wall rocks (Fig. 7.5C). It extends along the
upper contact of footwall pumice breccia for at least 1000 m
beyond the ore lenses (Large et al., 2001b). One textural-
compositional variant of this enveloping zone is represented
in alteration facies data sheet RB3. The white mica contents of
the facies varies from about 20 to 60% as plagioclase crystals
are increasingly replaced by carbonate and sericite, and the
glass shard-rich matrix by fine sericite, with proximity to
massive sulfide. The AI increases from 60 to 95 as the sericite
proportions increase.
Intense, schistose chlorite alteration facies (e.g. data sheet
RB5) is concentrated in the immediate footwall of the ore
lenses forming a thin (typically less than 5 m thick) chloritic
zone, which is commonly thickest (5-10 m) beneath the Cu-
rich sulfide lenses at the south end of the mine. This alteration
facies has variable chlorite (1550 wt%) and sulfide (10
30 wt%) contents. The AI is between 95 and 100, and the
CCPI between 70 and 90.
Commonly overlying the ore lenses is a zone of intense,
proximal Mn-carbonate alteration facies up to 10 m thick
(e.g. data sheet RB6), which is closely associated with massive
sulfide, but locally extends several tens of metres beyond the
limits of known sulfide lenses. The intense, proximal Mn-
carbonate alteration facies typically has a spotty texture, with
2560% Mn-carbonate spots in a sericitic, or locally chloritic,
matrix with low sulfide content. Carbonate composition in this
facies varies from rhodochrosite (MnCO
3
), to manganosiderite
((Mn,Fe)CO
3
) and kutnahorite (CaMn(CO
3
)
2
) (Braithwaite,
1974; Large etak, 2001b).
Typically the massive and semi-massive sulfides occur
in strata-parallel zones of intense quartz + sericite alteration
facies. This alteration facies has a bleached appearance with
textures that vary from massive-pervasive to spotty and augen-
schist textured (e.g. data sheet RB4). The latter comprises an
anastomosing fabric of strongly foliated sericite dominated
domains wrapping around siliceous knots of quartz with
minor sericite. The intense quartz + sericite alteration facies
continues laterally beyond the margins of the ore lenses, where
it contains 1-10% disseminated sulfides (pyrite, sphalerite,
galena).
Genesis of the ore lenses and alteration system
Most previous workers have interpreted the Rosebery deposit
to be synvolcanic exhalative in origin (e.g. Braithwaite, 1974;
Green et al., 1981; Huston and Large, 1987; Green and Iliff,
1989; Large, 1992; Khin Zaw et al., 1999). Solomon and
Groves (1994) consider that the sheet-like form, stratiform
sulfide banding, large size and high Zn-Pb metal content
indicate that Rosebery formed within a brine pool from
relatively high-salinity fluids, similar to the genesis of many
SEDEX deposits. However, the ore lenses are not associated
with well-developed stringer sulfide zones or alteration pipes
typical of seafloor systems. Instead, there are footwall zones
of disseminated sulfides, with altered zones that are aligned
parallel to the ore lenses and the volcanic strata. These
features, combined with textures in the massive sulfides
indicative of replacement, suggest that the ore lenses did
not form immediately above hydrothermal vents, but may
have formed from lateral fluid flow, either on the seafloor,
or below the seafloor by replacement of the fine-grained tops
of permeable pumice breccia units (Fig. 7.11: Allen, 1994a;
Doyle and Allen, 2003; Martin, 2004).
1 96 | CHAPTER 7
Weak, selective-pervasive albite + quartz * sericite alteration facies
Least-altered footwal!
R B 1
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
1396 02
weak, selective-pervasive albite + quartz
+ sericite
footwall
Kershaw Pumice Formation (CV C)
Mount Read V olcamics
feldspar-phyric pumice breccia
plagioclase
plagioclase crystals and cm-sized
plagioclase porphyritic tube pumice clasts
rhyolite
massive, clast supported, poorly sorted
subaqueous mass flow deposit
albite + sericite + quartz > chlorite + pyrite
+ hematite
selective-pervasive, disseminated, foliated,
sericite chlorite fiamme, sericite +
hematite stylolites, albite + sericite altered
plagioclase, feldspar overgrowths on
plagioclase
regional
good
weak
synvolcanic to burial
diagenetic
Geochemistry
SiO
2
TiO
2
AI
2
O
3
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
73.26
0.32
14.13
2.11
0.03
0.97
0.99
3.87
K
2
O
P
2
O
5
S
CO
2
Total
LOI
2.38
0.04
0.01
98.20
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
110
257
866
1
3
26
0.0
0.0
Zr
Nb
Y
Al
CCPI
Ti/Zr
257
14.2
33
41
31
7.5
Hand specimen photograph
Photomicrograph (ppl)
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 1 97
Subtle, selective albite * quartz + sericite alteration fades
Least-altered hanging wall
RB2
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
139586
subtle, selective albite + quartz + sericite
hanging wall
White Spur Formation
Mount Read Volcamics
feldspar > quartz crystal-rich pumiceous
sandstone
plagioclase and quartz
clastic (feldspar and quartz crystals, and
pumice shards)
Primary composition rhyolite-dacite
Lithofacies massive
Interpretation subaqueous mass-flow deposit
Alteration minerals albite + sericite + quartz + chlorite
selective clast alteration, disseminated
sericite, albite + sericite-altered plagioclase
regional
good
subtle
synvolcanic
diagenetic
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
Geochemistry
SiO
2
74.40 K
2
O 1.99 Sr 26 1 Nb 10.8
TiO
2
0.35 P
2
O
5
0.05 Ba 1472 Y 33
AI
2
O
3
14.34 S 0.03 Cu 3
:
Fe
2
O
3
2.06 CO
2
0.27 Pb 6 2 Al 34
MnO 0.04 Total 99.26 Zn 142 CCPI 27
MgO 0.57 LOI Sb 0.0 Ti/Zr 10.6
CaO 0.58 Tl 0.0
Na
2
O 4.41 Rb 75 Zr 197
Hand specimen photograph Photomicrograph (xn)
1 98 I CHAPTER 7
Moderate, foliated sericite alteration facies
Footwall
KB 3
Hand specimen photograph
Photomicrograph (xn)
Sample no. 139747
Alteration facies moderate, foliated sericite
Location footwall
Formation Kershaw Pumice Formation
Succession Mount Read Volcanics
Volcanic facies feldspar-phyric pumice breccia
Relict minerals plagioclase
Relict textures porphyritic, fibrous tube pumice clasts
Primary composition rhyolite
Lithofacies massive, normally graded
Interpretation syneruptive, mass-flow-emplaced deposit
Alteration minerals sericite + albite + quartz + carbonate
Alteration textures foliated, schistose, stylolites, fiamme?,
fractured and albite-altered plagioclase,
quartz veinlets
Distribution local
Preservation poor
Alteration intensity moderate
Timing synmineralisation
Alteration style hydrothermal and metamorphic
Geochemistry
SiO
2
71.46 K
2
O 4.16 Rb 182 Zr 229.9
TiO
2
0.32 P
2
O
5
0.06 Sr 56 Nb 12.6
AI
2
O
3
12.93 S 0.06 Ba 1042.5 Y 32
Fe
2
O
3
2.44 CO
2
1.87 Cu 3
MnO 0.11 Total 98.02 Pb 5 Al 6 7
MgO 1.6 5 LOI 3.89 Zn 29 CCPI 41
CaO 1.49 Sb 2.7 Ti/Zr 8.3
Na
2
O 1.36 Tl 0.9
Intense, augen
Footwal!
schistose quartz + seriate alteration facies
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 1 99
RB4
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
139778
intense, augen schistose quartz + sericite
footwall
Kershaw Pumice Formation (CV C)
Mount Read V olcamics
feldspar-phyric pumice breccia
nil
foliated clasts (pumice)
rhyolite
massive
indeterminate
sericite + quartz + sulfides
augen schistose, sericite + sulfide
cleavage
local
poor
intense
synmineralisation
hydrothermal and metamorphic
Geochemistry
SiO
2
TiO
2
Al
2
O
3
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
72.95
0.27
12.86
2.58
0.17
1.14
0.27
0.01
K
2
O
P
2
O
5
S
co
2
Total
LOI
4.24
0.04
1.94
0.40
98.19
3.57
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
202
18
176 7
239
4300
6 900
7.5
4.5
Zr
Nb
Y
Al
CCPI
Ti/Zr
224
13.2
35
95
45
7.2
Hand specimen photograph Photomicrograph (xn)
2 00 I CHAPTER 7
Intense, schistose chlorite alteration facies
Footwall
RB5
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
139743
intense, schistose chlorite
footwall
Kershaw Pumice Formation (CVC)
Mount Read V olcanics
feldspar-phyric breccia
nil
porphyritic?
rhyolite
massive
indeterminate
chlorite + pyrite + sphalerite + quartz +
sericite
foliated, schistose
local
poor
intense
synmineralisation
hydrothermal and metamorphic
Geochemistry
SiO
2
TiO
2
MA
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
44.22
0.33
14.48
16 .58
1.49
2.22
0.10
0.05
K
2
O
P
2
O
5
S
co
2
Total
LOI
3.37
0.06
7.18
0.95
91.03
7.55
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
174
14
948
16 78
604
6 8200
3.9
7.1
Zr
Nb
Y
Al
CCPI
Ti/Zr
247
12.0
35
97
83
8.0
Hand specimen photograph
Photomicrograph (ppl)
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS I 2 0 1
Intense, proximal In-carbonate alteration facies
Hanging waff
RB6
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
139740
intense, proximal Mn-carbonate
hanging wall
White Spur Formation
Mount Read V olcamics
feldspar-phyric pumice breccia?
nil
rare feldspar crystals
dacite
massive
indeterminate
rhodochrosite + sericite + pyrite
nodular-spheroidal rhodochrosite,
disseminated pyrite, sericite-altered
feldspar
local, immediate hanging wall ofsulfide
lens
poor
intense
synmineralisation
hydrothermal
Geochemistry
SiO
2
TiO
2
AI
2
O
3
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
16 .79
0.39
10.73
7.50
29.36
1.97
2.04
0.05
K
2
O
P
2
O
5
S
co
2
Total
LOI
Rb
3.6 4
0.13
3.24
22.20
98.6 4
23.35
183
Sr
Ba
Cu
Pb
Zn
Sb
TI
Zr
27
4297
20
674
995
8.8
24.6
142
Nb
Y
Al
CCPI
Ti/Zr
7.2
20
73
70
16 .5
Hand specimen photograph Photomicrograph (ppl)
I CHAPTER 7
7.7 | WESTERN THARSIS: A HYBRID Cu-
Au VHMS DEPOSIT
Western Tharsis is a stratabound disseminated pyrite + chalco-
pyrite deposit in the northwestern part of the Mount Lyell
mining field, western Tasmania (Fig. 1.5). It contains around
12.4 Mt at 1.3% Cu and 0.3 g/t Au (Huston and Kamprad,
2001). Although discovered in 1897, it is the only non-exploited
deposit of at least 22 deposits in the Mount Lyell field, which
together produced a total of 113 Mt of ore at average grades of
1.36% Cu, 6.8 g/t Ag and 0.4 g/t Au (Corbett, 2001).
The mineralised zone is a sub-vertical stratabound lens,
up to 150 m thick and narrowing towards the surface, with
a down dip extent of greater than 1000 m and strike extent
of about 300 m. Most of the deposit consists of disseminated
pyrite + chalcopyrite in a gangue of quartz + sericite chlorite
and, locally, magnetite. Smaller bornite-rich mineralised
zones similar to the North Lyell ores exist in the upper parts,
particularly associated with quartz and quartz + pyrophyllite
alteration assemblages. The bornite zones also contain minor
chalcocite, chalcopyrite, mawsonite, digenite, enargite, molyb-
denite, woodhouseite and barite.
Geological setting
The Lyell ore bodies and their altered halos are focussed along
the Great Lyell fault and occur at a variety of stratigraphic
intervals in the Central Volcanic Complex and in the
overlying Mount Julia Member of the Comstock Formation,
Tyndall Group (Corbett, 2001). Western Tharsis is situated
in a steeply west-dipping, overturned, east-facing succession
of altered intermediate to felsic volcanic rocks assigned to the
Central Volcanic Complex of the Middle Cambrian Mount
Read Volcanics. In the Mount Lyell area, these volcanic rocks
are reverse-faulted to the east against the late Cambrian to
Early Ordovician siliciclastic conglomerate and sandstone of
the Owen Group. All the Mount Lyell field deposits lie within
a 6 x 1 km pyritic altered zone adjacent to the complex fault
contact (Corbett, 2001).
Two units of rhyolitic volcaniclastic rocks with subordinate
interbedded volcanogenic sandstone and siltstone comprise
the immediate stratigraphic footwall and host rocks at
Western Tharsis (Huston and Kamprad, 2001). These units,
each several hundred metres thick, contain some ash- to
lapilli-sized clasts but primary volcanic textures are typically
obscured by alteration and their volcaniclastic origin is largely
interpretative. They are separated by a 10-50 m thick group
of andesitic volcaniclastic rocks and locally amygdaloidal
coherent lavas or sills. The stratigraphic hanging wall consists
of a 200300 m thick complex of intercalated felsic and
intermediate volcaniclastic rocks. A thin unit of felsic quartz
porphyry, possibly a correlate of the lower Tyndall Group,
occurs between the altered Central Volcanic Complex and the
faulted contact with the Owen Group. In this part of the field,
the (North Lyell) fault contact dips at 70 to the southwest.
Deep drilling indicates that the Western Tharsis mineralised
zone may intersect the fault at around 1500 m below surface
(Corbett, 2001).
The presence of thick, graded beds in the Central Volcanic
Complex at Lyell indicates a subaqueous environment of
deposition for the host succession. The occurrence of exhala-
tive massive sulfide bodies, limestone with shallow marine
fauna, and welded ignimbrite in the Tyndall Group at
Comstock suggest a shallow marine setting for mineralisation
(Jago et al., 1972; Corbett et al., 1974; White and McPhie,
1997; Corbett, 2001).
Alteration facies and zonation
Corbett (2001) showed that the deposit is enclosed by a
400-500 m wide zone of quartz + sericite + pyrite schist
adjacent to the North Lyell fault. This is part of a pyritic core
zone extending 4 km from the Lyell Highway to the Lyell
Comstock mine. This proximal, strong to intense, feldspar-
destructive altered zone grades outwards to less intense sericite
+ chlorite alteration facies in felsic and intermediate volcanic
rocks. At surface above Western Tharsis and around the upper
mineralised zone, the proximal quartz + sericite + pyrite
alteration facies includes numerous bodies up to 20 m across,
of microcrystalline quartz pyrite, termed silica heads.
Huston and Kamprad (2001) subdivided the Western
Tharsis system into five main alteration facies. An intense,
pervasive, proximal quartz + chlorite + pyrite sericite alter-
ation facies (e.g. data sheet WT8) exists in the chalcopyrite +
pyrite mineralised zone at depths greater than 350 m below
surface.
An intense, proximal quartz + pyrophyllite + pyrite
alteration facies (e.g. data sheets WT4 and 5) occurs in a
150 m wide zone associated with the bornite + chalcopyrite
mineralised zone between 100 and 400 m below surface
and in a 50 m thick zone extending along the stratigraphic
footwall to 750 m below surface. This facies includes narrow
zones of quartz + topaz assemblages (particularly in the upper
parts, which may correspond to Corbett's silica heads) and
locally minor phases including fluorite, barite, zunyite and
woodhouseite.
A strong, pervasive, medial quartz + sericite + pyrite alter-
ation facies (e.g. data sheets WT3, 6 and 7), which encloses
the two proximal facies above, and extends up to 150 m
outwards into the stratigraphic footwall and through most of
the hanging wall. The outer margins, adjacent to the weak,
medial chlorite + sericite carbonate alteration facies and
in upper part of hanging wall succession, contain minor
disseminated carbonate.
A weak, pervasive, medial chlorite + sericite carbonate
alteration facies (e.g. data sheet WT2) occurs in the peripheral
zones. It exists in both footwall and hanging wall, about 150-
200 m outwards from the mineralised zone, particularly in
andesitic volcanic rocks with minor pyrite or hematite.
A weak, selective quartz + chlorite + sericite + carbonate
alteration facies (e.g. data sheet WTl ) grades westward into
least-altered rocks composed of quartz + albite + chlorite (
sericite and carbonate).
Sericite compositions in the outer zones are slightly
phengitic (up to 0.5 Fe + Mg atoms per formula unit) grading
to essentially non-phengitic and slightly sodic (molecular Na/
Na+K <0.15) in the proximal to medial zones. This variation
has potential as a deposit-scale exploration vector, which can
be effectively measured by short wavelength infrared (SWIR)
spectral analysis (Herrmann et al., 2001). SWIRspectrometry
is similarly effective in delineating pyrophyllite, topaz and
zunyite-bearing zones (see Section 8.2 for more detail).
Chlorite compositions are moderately Fe-rich (molecular
Mg/Mg + Fe ratios = 27-48) with a subtle trend to Fe enrichment
towards the mineralised zone. Carbonates in the distal to
medial alteration facies are ankeritic to sideritic in composition.
They show a distinct trend of Fe enrichment from the footwall
towards the mineralised zone (Huston and Kamprad, 2001).
Carbonates in the hanging wall are moderately manganiferous
(up to 0.2 Mn atoms per formula unit).
Ore genesis
Metallogenic interpretations of the Mount Lyell deposits
have fuelled geological debate for over a century and remain
controversial today (Corbett, 2001; Huston and Kamprad,
2001). Early models that related mineralisation to Devonian
or Cambrian intrusions were succeeded, during the 1960s,
by acceptance of Cambrian synvolcanic origins. In the 1980s
and early 1990s, the North Lyell type bornite ores were
popularly attributed to re-mobilisation during Devonian
deformation. Large et al. (1996) revived the magmatic
connection, interpreting Cambrian granites to be the source
of hydrothermal fluids and metals. In recent years, a magmatic
connection has been further supported by wider recognition
of advanced argillic type assemblages, which are consistent
with the involvement of magmatic volatiles and acidic fluids.
Nevertheless, there is still disagreement over the timing
of mineralisation. Huston and Kamprad (2001) pointed
to an apparent (Pb-isotopic) 40 Ma age difference between
stratiform synvolcanic Pb + Zn + Cu sulfide lenses at Lyell
Comstock and the disseminated Cu + Au deposit at Prince
Lyell. They suggested a two event history: Middle Cambrian
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 03
syngenetic stratiform Pb + Zn + Cu mineralisation followed
by a 40 Ma period of tectonism that culminated in high-
sulfidation type Cu + Au mineralised zones derived from
deep Ordovician granites. However, the more extensive
field evidence gathered by Corbett (2001) indicates that all
the alteration and mineralisation was restricted to Middle
Cambrian, and ceased during deposition of the lower part of
the Tyndall Group.
Rather than a temporal overprinting of different styles,
Corbett (2001) envisaged a single, vertically extensive,
submarine volcanic, hybrid magmatic-seawater hydrothermal
system. It produced disseminated chalcopyrite + pyrite (and
locally magnetite + apatite) mineralised zones in the deeper
parts, high-sulfidation type bornite mineralised zones and
intense siliceous altered zones in the upper subseafloor zones,
and deposited exhalative Pb + Zn + Cu massive sulfide lenses
at the seafloor. The Western Tharsis zone encompasses the
transition between deep chalcopyrite + pyrite and upper high-
sulfidation types of mineralisation (Fig. 7.22).
Corbett's diagrammatic representation shows the system
as sub-vertical, cutting through sub-horizontal volcanic
strata and focussed along or adjacent to the Great Lyell fault.
However, the Western Tharsis deposit appears to be sub-
vertical and stratabound. This is possibly a misinterpretation;
primary volcanic textures and facies associations are largely
obscured in the intensely altered zones. Furthermore, Corbett's
(2001) model suggests diapir-like upward movement of the
phyllosilicate-rich altered volcanic rocks on the hanging wall
side of the fault zone, which may have disrupted the volcanic
sequence.
The arguments about Mount Lyell are not yet settled.
Nevertheless, the emerging recognition of high-sulfidation
ore deposits may renew interest in exploration in western
Tasmania.
pyntic core zone
(senate + chlorite + pyrite
silica schists)
disseminated
chalcopyrite-pyrite bodie
FIGURE 7.22 | Cross-section model of the Mount Lyell,
vertically extensive, submarine, hybrid magmatic-seawater
hydrothermal, alteration and mineralisation system, western
Tasmania (modified after Corbett, 2001).
2 04 | CHAPTER 7
Weak, selective chlorite + sericite + quartz + carbonate alteration facies
Least-altered footwall
WT 1
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
113084
weak, selective chlorite + sericite + quartz
+ carbonate
>250 m stratigraphicaliy below mineralised
zone
Central V olcanic Complex
Mount Read V olcanics
massive plagioclase + quartz-phyric
rhyolite
phenocrysts of albitised plagioclase >
quartz
porphyritic
rhyolite
massive
rhyolite lava
sericite + chlotite + carbonate > quartz
selective-pervasive, matrix altered to 2 0-
40 | jm chlorite, aligned sericite cleavage,
albite + sericite-altered plagioclase
regional?
moderate
weak
synvolcanic and syndeformation
diagenetic and tectonic-metamorphic
Geochemistry
SiO
2
75.6 9 K
2
0 3.33 Sr 37 Nb 13
TiO2 0.26 P
2
O
5
0.04 Ba 243 Y 40
AI
2
O
3
13.00 S 0.37 Cu 3
Fe
2
O
3
2.45 CO
2
1.17 Pb 14 Al 6 5
MnO 0.04 Total 99.32 Zn 28 CCPI 40
MgO 0.78 LOI 3.12 Sb 0.7 Ti/Zr 5.4
CaO 1.04 Tl 0.9
Na
2
O 1.15 Rb 112 Zr 291
Hand specimen photograph Photomicrograph (xn)
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 05
Weak, pervasive, medial chlorite + sericite carbonate alteration facies y\/T 2
Sample no. 113086
Alteration facies weak, pervasive, medial chlorite + sericite
carbonate
Location -2 00 m stratigraphically below
mineralised zone
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
Central V olcanic Complex
Mount Read V olcanics
altered andesite
nil
nil
andesite
massive
indeterminate
chlorite + carbonate > sericite
pervasive, foliated, carbonate-altered
plagioclase, carbonate veins
poor
weak
synvolcanic and subsequent
syndeformation
diagenetic and tectonic-metamorphic
Geochemistry
SiO
2
47.78 K
2
0 1.47 Rb 46 Zr 72
TiO
2
0.51 P
2
O
5
0.09 Sr 83 Nb 3
AI
2
O
3
14.56 S 0.04 Ba 36 7 Y 17
Fe
2
O
3
12.59 CO
2
7.84 Cu 10
MnO 0.19 Total 97.30 Pb 6 Al 45
MgO 4.6 3 LOI 10.71 Zn 199 CCPI 82
CaO 5.53 Sb 0.8 Ti/Zr 42.5
Na
2
O 2.07 Tl 0.5
Hand specimen photograph Photomicrograph (xn)
2 06 | CHAPTER 7
Strong, pervasive, medial quartz + sericite + pyrite alteration facies
WT 3
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
113092
strong, pervasive, medial quartz + sericite
+ pyrite
-1 00 m stratigraphically below mineralised
zone
Central V olcanic Complex
Mount Read V olcanics
volcaniclastic rhyolite
nil
relict granular texture, possibly clastic
rhyolite
indeterminate
quartz + sericite + siderite
pervasive, mosaic of sutured 2 0-6 00 | j
fractured quartz grains with interstitial
sericite and patches coarse siderite
poor
strong
synmineralisation
hydrothermal
Geochemistry
SiO
2
70.58 K
2
O 3.6 3 Rb
TiO
2
0.23 P
2
O
5
0.04 Sr
AI
2
O
3
12.12 S 0.02 Ba
Fe
2
O
3
8.03 CO
2
3.11 Cu
MnO 0.11 Total 98.92 Pb
MgO 0.6 8 LOI 4.44 Zn
CaO 0.16 Sb
N a,0 0.21 TI
126 Zr 246
19 Nb 13
758 Y 42
18
5 Al 92
6 3 CCPI 6 7
2.5 Ti/Zr 5.6
0.9
Hand specimen photograph Photomicrograph (xn)
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 07
Intense, proximal quartz + pyrophyllite + pyrite alteration facies yyi 4
Hand specimen photograph Photomicrograph (xn)
Sample no. 113102
Alteration fades intense, proximal quartz + pyrophyllite +
pyrite
Location proximal to upper mineralised zone
Formation Central V olcanic Complex
Succession Mount Read V olcanics
V olcanic fades altered rhyolite
Relict minerals nil
Relict textures nil
Primary composition rhyolite
Lithofacies
Interpretation indeterminate
Alteration minerals quartz + pyrophyllite > sericite + pyrite
Alteration textures pervasive, mosaic of sutured 4 0-4 00 prn
, quartz grains with ragged patches of semi-
aligned pyrophyllite, minor sericite and
disseminated, fractured pyrite
Distribution
Preservation poor
Alteration intensity intense
Timing
Alteration style hydrothermal
Geochemistry
SiO
2
80.78 P
2
O
5
0.05 Cu 72 Al 80
TiO
2
0.24 S 1.6 1 Pb 7 CCPI 54
AI
2
O
3
12.25 CO
2
0.40 Zn 6 Ti/Zr 5.8
Fe
2
O
3
2.21 Total 99.49 Sb 0.6
MnO 0.01 LOI 3.08 Tl <0.5
MgO 0.12 Zr 249
CaO 0.03 Rb 31 Nb 13
Na
2
O 0.36 Sr 117 Y 5
K
2
O 1.43 Ba 1170
2 0 8 | CHAPTER 7
Intense, proximal quartz + pyrophyllite + pyrite alteration facies
WT 5
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
113105
intense, proximal quartz + pyrophyllite +
pyrite
-2 00 m straigraphicaily above mineralised
zone
Central Volcanic Complex
Mount Read Volcanics
altered rhyolite
rhyolite
indeterminate
quartz + topaz + pyrite > carbonate
pervasive, domainal 50-100 pm
microcrystalline quartz and granular topaz,
minor disseminated pyrite, irregular mm-
scale patches > carbonate veinlets
poor
intense
hydrothermal
Hand specimen photograph
Photomicrograph (xn)
Geochemistry
SiO
2
6 8.25 P
2
O
5
0.07 Cu 10 Al 36
TiO
2
0.28 S 0.37 Pb 21 CCPI 96
AI
2
O
3
18.13 CO
2
3.11 Zn 14 Ti/Zr 4.6
Fe
2
O
3
1.35 Total 95.14 Sb 0.1
MnO 0.30 LOl 8.19 Tl <0.5
MgO 1.12 Zr 36 4
CaO 2.05 Rb 1 Nb 14
Na
2
O 0.05 Sr 33 Y 10
K
2
O 0.06 Ba 171
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 2 0 9
Strong, pervasive, medial quartz + sericite + pyrite alteration facies yyy g
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
113110
strong, pervasive, medial quartz + sericite
+ pyrite
-1 00 m stratigraphically above mineralised
zone
Central V olcanic Complex
Mount Read V olcanics
altered dacite
dacite
indeterminate
quartz + sericite + pyrite > chlorite and
carbonate
pervasive, 5 0-1 00 | jm microcrystalline
quartz with interstitial shreds and seams
of aligned sericite, disseminated euhedral
pyrite, some highly deformed and re-
crystallised quartz + carbonate > chlorite
veins/patches
poor
strong
hydrothermal
Geochemistry
SiO
2
TiO
2
AI , 0
'2^3
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
6 0.29
0.35
12.84
10.77
0.52
1.08
1.26
0.13
K
2
O
S
co
2
Total
LOI
3.97
0.08
6 .28
2.6 4
100.21
8.09
Rb
Sr
Ba
Cu
Pb
Zn
Sb
120 Tl
25 Zr
1737 Nb
88 Y
214
138 Al
1.0 CCPI
Ti/Zr
1.0
182
9
23
78
72
11.5
Hand specimen photograph Photomicrograph (xn)
2 1 0 I CHAPTER 7
Strong, pervasive, medial quartz + sericite + pyrite alteration facies
WT 7
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
113284
strong, pervasive, medial quartz + sericite
+ pyrite
deep mineralised zone
Central V olcanic Complex
Mount Read V olcanics
altered rhyolite
nil
nil
rhyolite
indeterminate
quartz + sericite + pyrite + chalcopyrite
pervasive: 5 0-1 00 pm microcrystalline
quartz, interstitial shreds and seams of
sericite, disseminated euhedral pyrite,
> chalcopyrite
poor
strong
hydrothermal
Geochemistry
SiO
2
TiO
2
AI
2
O
3
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
72.75
0.22
10.11
6 .11
0.01
0.36
0.33
0.05
K
2
O
S
CO
2
Total
LOI
2.80
0.35
4.17
0.02
97.28
4.22
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
6 4 Zr
111 Nb
1330 Y
13800
21 Al
32 CCPI
99.2 Ti/Zr
<0.5
200
11
12
6 7
6 .6
Hand specimen photograph Photomicrograph (xn)
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 1 1
Intense, pervasive, proximal quartz + chlorite + pyrite sericite alteration facies WT 8
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
11326 4
intense, pervasive, proximal quartz +
chlorite + pyrite sericite
deep mineralised zone
Central V olcanic Complex
Mount Read V olcanics
altered dacite
dacite
indeterminate
quartz + sericite + chlorite + pyrite chal-
copyrite
pervasive: mosaic of 4 0-2 00 pm dusty
quartz, interstitial and anastomosing
sericite cleavage, irregular domains of
chlorite + sulfides
Distribution
Preservation
Alteration intensity
Timing
Alteration style
poor
intense
hydrothermal
Geochemistry
SiO
2
TiO
2
AI
2
O
3
F e
2
O
3
MnO
MgO
CaO
58.74
0.49
13.6 5
12.6 6
0.23
1.75
0.26
K
2
O
P
2
O
5
S
co
2
Total
LOI
3.45
0.27
5.38
0.17
97.08
6.73
Na
2
O 0.03
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
92
36
2097
12000
15
99
0.9
<0.5
Zr 16 7
Nb 13
Y 25
Al 95
CCPI 79
Ti/Zr 17.6
Hand specimen photograph Photomicrograph (xn)
2 1 2 | CHAPTER 7
7.8 | HENTY: A V O L C A N O G E N I C G O L D
DEP OSIT
The Henty volcanogenic gold mine and nearby Mount Julia
gold prospect are hosted by Cambrian Mount Read Volcanics,
near the junctions of the North and South Henty faults
and the Great Lyell fault (Fig. 1.5). The deposit comprises
at least six steeply dipping, thin, stratabound, disconnected
siliceous lenses of up to a few hundred metres vertical extent,
distributed over about 2.5 km of strike length (Callaghan,
2001). The estimated total geological resource in December
2001 was 2,154,000 tonnes at 12.1 g/t Au (838,800 oz.).
This included production of 820,000 tonnes at 17.5 g/t Au
(462,000 oz.).
Geological setting
The ore lenses lie in a laterally extensive but narrow stratabound
altered zone (A-zone of Callaghan, 2001) at the stratigraphic
boundary between the Central Volcanic Complex and the
base of the Tyndall Group. In the Henty area these units
trend NNW to NNE and face east, with steep easterly to
slightly overturned steep westerly dips. The Henty fault zone,
trending about 015 and dipping at 70 to the west, obliquely
truncates the volcanic succession. The stratabound altered and
mineralised zones occur in the immediate footwall of the fault
zone, extending about 200 m down-dip from the fault (Halley
and Roberts, 1997; Callaghan, 2001). The intersection of the
fault and the favourable stratigraphic horizon plunges at a
low angle to the south. This is a consequence of the gradual
change in trend and slight overturning of the host sequence,
from NNW with steep easterly dip in the south, to NNE
and steep westerly dip in the northern part of the mine area
(Halley and Roberts, 1997). As in many Au deposits, grade
cut-offs rather than lithological differences define the ore
zones (Callaghan, 1998). Most of the high-grade ore exists in
thin lenses or sheet-like bodies up to 7 m thick in the intense,
massive quartz (MQ) alteration facies (Halley and Roberts,
1997) but this facies is not uniformly auriferous (Callaghan,
2001). The stratigraphic upper part of the mineralised A-zone
typically has a high disseminated base-metal-sulfide content or
is spatially associated with lenses of massive pyrite or massive
to banded sphalerite + galena (Penney, 1998). Discontinuous
massive pyrite lenses up to 2 m thick exist at this stratigraphic
level for 600 m of strike but extend less than 150 m down-dip
from the Henty fault.
Feldspar-phyric to aphyric dacitic lavas and rare basaltic
lavas intercalated with dacitic to basaltic hyaloclastite and
polymictic volcanic breccias of the Central Volcanic Complex
dominate the stratigraphic footwall between the Henty
fault and A-zone (e.g. data sheet HN1: Callaghan, 1998).
The footwall succession includes discontinuous calcareous
volcaniclastic units and hematitic fossiliferous limestone.
Immobile element ratios indicate that the protoliths of the
altered and mineralised zone were compositionally uniform
dacitic volcanic units.
The stratigraphic hanging wall, immediately east of the
A-zone, comprises massive, andesitic, feldspar crystal-rich
volcanic sandstone (e.g. data sheet HN2), dacitic volcanic
breccia, lavas and polymictic volcano-sedimentary breccia,
intercalated with calcareous volcanic sandstone, hematitic
fossiliferous limestone and minor mudstone. This lithologically
diverse part of the hanging wall succession, up to 200 m
thick, is recognised as the Lynchford Member: the lowermost
unit of the Tyndall Group (Callaghan, 2001). It is succeeded
eastwards by the Mount Julia Member consisting of graded
rhyolitic breccia, quartz + feldspar-rich volcanic sandstone and
minor siltstone intruded by southward thickening quartz +
feldspar porphyritic rhyolite sills, cryptodomes and associated
hyaloclastites. Overlying this is a thick succession of quartz-
rich epiclastic sandstone and volcanolithic conglomerate (Zig
Zag Hill Formation), which passes conformably eastwards
into siliciclastic and micaceous sandstone and conglomerate.
White and McPhie (1996) interpreted the massive
crystal-rich volcanic sandstone of the Lynchford and Mount
Julia Members as originating from large subaerial or shallow
marine explosive eruptions that produced pyroclastic flows,
which transgressed into a shallow marine environment. The
fossil assemblage in the limestone units (Jago et al., 1972)
and local welded ignimbrite units in the Mount Julia Member
(White and McPhie, 1996) also indicate that the Henty host
rocks, or at least those immediately overlying the mineralised
zone, were deposited in a near-shore, shallow-marine setting.
Alteration facies and zonation
The distribution of alteration facies in this elongate and
stratabound alteration system reflects decreasing alteration
intensity down-dip away from its intersection with the Henty
fault zone, and differing thermo-chemical conditions from
footwall to hanging wall. Table 7.2 summarises the features of
the HentyMount Julia alteration facies.
A moderate, footwall sericite + quartz carbonate alter-
ation facies (MA) occupies the wedge of stratigraphic footwall
between the Au-bearing A-zone and the Henty fault.
The A-zone has a discontinuous sheet-like inner zone
of intense, massive quartz alteration facies (MQ, e.g. data
sheet 7), which is composed of microcrystalline quartz with
multiple generations of fine veinlets of quartz + calcite
sulfides (pyrite, chalcopyrite and galena, which contain most
of the Au). It grades outwards (stratigraphically up and down
as well as down-dip away from the fault) through intense,
proximal, domainal quartz + sericite alteration facies (MV,
e.g. data sheet HN6) to intense, foliated sericite + quartz +
chlorite + pyrite alteration facies (MZ, e.g. data sheet HN5).
These enveloping alteration facies have progressively less
quartz and greater phyllosilicate contents and exist in variable
proportions in different parts of the deposit. They typically
have sericitic groundmasses with foliated to schistose fabrics
anastomosing around small siliceous domains that possibly
represent silicified lithic clasts (Callaghan, 2001). Massive
pyrite lenses at the stratigraphic top of the A-zone grade
laterally and down-dip to massive carbonate alteration facies
(CB, e.g. data sheet HN4) in peripheral areas. This facies
includes massive rocks composed of carbonate + chlorite,
and thin bands and lenses of carbonate that are difficult
to distinguish from fossiliferous limestone. Some of the
carbonate bands contain fragments of red jasper.
A zone of strong, selective, hanging wall albite + quartz
alteration facies (e.g. data sheet HN3) lies either immediately
above (Callaghan, 2001) or 20-40 m stratigraphically above
the mineralised zone (Halley and Roberts, 1997). In proximal
areas, the hanging wall albite + quartz alteration fades is up to
100 m thick, possibly extending up to the base of the Zig Zag
Hill Formation (Callaghan, 1998). It is considered to be of
hydrothermal origin, distinct from regional-scale diagenetic
albite alteration facies.
Ore genesis
Halley and Roberts (1997) interpreted Henty as a Au-rich
volcanogenic massive sulfide deposit because of its association
with conformable pyrite and carbonate lenses, colloform
textures in pyrite, the presence of red jasper clasts that resemble
siliceous exhalites, and C-, O- and Pb-isotopic data that
indicate a Cambrian synvolcanic origin for the stratabound
alteration system. They suggested that its unusual high Au/Ag
ratios, extent of footwall silicification and high proportion of
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 2 1 3
base-metal sulfides disseminated in the upper altered footwall
zone were due to the shallow-marine, near-shore setting where
input of meteoric water produced low-salinity hydrothermal
fluids, which boiled and cooled at some depth below the
seafloor. Although the Au in the intense, massive quartz
alteration facies exists largely in late-stage veinlets related to
Devonian brittle deformation and remobilisation, there is no
evidence for addition of metals during this event.
Callaghan (2001) proposed synvolcanic inputs of
magmatic volatiles, fluids and metals to account for the Au
+ Cu + Bi + Ag + Te metal association and Al mobility in the
intense, massive quartz alteration facies, which are atypical of
seawater dominated VHMS systems. He envisaged a low pH,
high salinity, submarine, subseafloor type of pulsed magmatic
plus seawater, high-sulfidation epithermal system. This model
invokes the proto-Henty fault as a magmatic volatile and fluid
conduit that reactivated during Devonian deformation to
dislocate the eastern and western halves of the hydrothermal
system.
Table 7.2 | The Henty-Mount Julia alteration facies and their defining characteristics (Callaghan, 1998).
Alteration facies
Intense, massive
quartz
Intense, proximal
quartz + sericite
Intense, peripheral
sericite + quartz +
chlorite + pyrite
Massive carbonate
Moderate, footwall
sericite + quartz
carbonate
Strong, hanging wall
albite + quartz
Code
MQ
MV
MZ
CB
MA
AS
Mineral assemblage
Quartz (carbonate, sericite,
pyrite, chalcopyrite, galena,
gold)
Quartz + sericite
(carbonate, pyrite,
chalcopyrite, galena,
sphalerite)
Sericite + quartz + pyrite
+ chlorite (carbonate,
chalcopyrite, galena)
Calcite chlorite
Sericite + quartz
(carbonate, pyrite)
Albite + quartz (chlorite)
Sulfides
(%)
~2
0.1 to 5
2-10
<10
<2
0
Gold
(g/t)
Variable;
average 36
0.1 to 1
0.5 to 2
?
?
0
Distribution
Thin lenses in core of A-zone.
Enclosing and gradational to intense massive
quartz alteration facies.
Peripheral, enveloping the intense proximal
quartz + sericite alteration facies.
Discontinuous stratiform lenses at stratigraphic
top of A-zone in peripheral parts of the system;
laterally equivalent to massive pyrite lenses.
Stratigraphic footwall, in felsic volcanic rocks
between Henty fault and A-zone.
Directly adjacent to A-zone and extending up to
100 m into hanging wall succession.
2 1 4 | CHAPTER 7
Moderate albite + chiorite + calcite alteration facies
Least-altered host rock
HN 1
Sample no.
Alteration Facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
I nterpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
2 5 5 005
moderate albite + chlorite + calcite
down di pofA-zone
Central V olcanic Complex
Mount Read V olcanics
massive plagioclase-phyric dacite
albitised plagioclase phenocrysts
porphyritic
dacite
massive to brecciated
dacite lava
albite + chlorite + calcite + quartz
selective-pervasive in irregular chlorite and
calcite veinlets and blebs, albite sericite-
altered plagioclase
regional
moderate
moderate
synvoicanic plus subsequent fault-related
deformation
diagenetic and tectonic deformation
Geochemistry
SiO
2
TiO
2
AI
2
O
3
F e
2
O
3
MnO
MgO
CaO
N a
2
O
58 .40
0.46
12.50
4 .4 2
0.13
1.51
8.71
5.05
K
2
O
S
co
2
Total
LOI
0.97
0.17
0.01
6.82
99.15
7.91
Au 0.005
Sr
Ba
Cu
Pb
Zn
Sb
Tl
Zr
114
327
7
8
139
1.2
0.5
188
Nb
Y
Al
CCPI
Ti/Zr
4 0.9
28
15
48
14.7
Hand specimen photograph Photomicrograph (xn)
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 2 1 5
Weak, selective aibite + chlorite + caicite alteration facies
Least-aitered hanging wall
HN2
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
I nterpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
2 5 5 02 4
weak, selective aibite + chlorite + caicite
lower hanging wall unit + down dip of A-
zone
Lynchford Member (Tyndall Group)
Mount Read V olcanics
massive, feldspar crystal-rich volcanicastic
sandstone
albitised plagioclase crystals
abundant sand-sized crystals and sparse
lithic clasts, subangular
andesite
massive to crudely banded, moderately
well sorted, matrix supported
volcaniclastic mass-flow deposit
aibite + chlorite + caicite > sericite
selective-pervasive aibite + chlorite-altered
matrix, irregular discontinuous caicite
veinlets, domainal microcrystalline quartz +
chlorite + pyrite
regional
moderate
weak
synvolcanic
diagenetic and tectonic deformation
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
54.90 K
2
O 0.29 Au 0.005 Zr 130
TiO
2
0.6 4 P
2
O
5
0.09 Sr 134 Nb 29.6
AI 2O
3
14.40 S 0.08 Ba 127 Y 22
Fe
2
O
3
3.73 CO
2
6 .97 Cu 7
MnO 0.19 Total 98.89 Pb 5 Al 12
MgO 1.81 LOI 8.72 Zn 132 CCPI 42
CaO 9.05 Sb 0.8 Ti/Zr 29.5
Na
2
O 6 .74 Tl 0.5
2 1 6 | CHAPTER 7
Strong, selective, hanging-wall albite + quartz alteration facies
Hanging waff
HN3
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
255038
strong, selective, hanging wall albite +
quartz (AS)
hanging wall, + 70 m stratigraphically
above A-zone
Mount Julia Member (Tyndall Group)
Mount Read V olcanics
massive volcanicastic sandstone
quartz and albitised plagioclase crystals
sparse crystals and few lithic clasts in
1 0-2 0 urn matrix
rhyolite
massive to crudely banded
volcaniclastic mass-flow deposit
albite + quartz > sericite > chlorite + calcite
selective-pervasive altered matrix,
microcrystalline albite + quartz matrix with
minor interstitial chlorite, quartz + calcite
infill irregular veinlets, sericite in later
parallel veinlets
local and stratabound in hanging wall
sequence
moderate
strong
albite + quartz probably syn volcanic,
sericite veinlets syn deformation
hydrothermal?
Geochemistry
SiO
2
TiO
2
AI
2
O
3
F e
2
O
3
MnO
MgO
CaO
Na
2
O
77.70
0.16
12.10
1.13
0.02
0.36
0.75
6 .57
K
2
O
S
co
2
Total
LOI
0.29
0.01
0.00
0.99
100.09
0.90
Au 0.005
Sr
Ba
Cu
Pb
Zn
Sb
Tl
Zr
47 Nb
100 Y
4
2 Al
100 CCPI
0.9 Ti/Zr
0.5
182
14.6
26
17
5.3
Hand specimen photograph Photomicrograph (xn)
Massive carbonate alteration facies
Host-rock equivalent?
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
255050
massive carbonate
upper A-zone, laterally equivalent to
massive pyrite
Lynchford Member (Tyndall Group) or
Central Volcanic Complex
Mount Read Volcanics
marine limestone with minor volcaniclastic
component
calcite + plagioclase + quartz
plagioclase and quartz crystal fragments
massive to thinly bedded
impure marine carbonate
calcite?
10-50 |jm microcrystalline calcite, calcite
veins, stylolites
local and stratabound in peripheral upper
part of A-zone
moderate
weak
synvolcanic diagenesis, syndeformational
dynamic recrystallisation
diagenetic and tectonic-metamorphic,
doubtful hydrothermal carbonate
component
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 1 7
HN4
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
22.8 0 K
2
O 0.58 Sr 26 3 Nb 1.0
TiO
2
0.18 P
2
O
5
0.08 Ba 476 Y 16
AI
2
O
3
4.16 S 0.72 Cu 21
Fe
2
O
3
2.33 CO
2
24.70 Pb 76 A! 4
MnO 0.26 Total 94.44 Zn 100 CCPI 6 0
MgO 0.85 LOI 29.11 Sb 10.0 Ti/Zr 22.0
CaO 36 .40 Tl 0.5
Na
2
O 1.38 Au 0.017 Zr 49
2 1 8 | CHAPTER 7
Intense, foliated sericite + quartz + chlorite + pyrite alteration facies
Footwali?
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
255030
intense, foliated sericite + quartz + chlorite
+ pyrite (MZ)
peripheral altered zone
Central Volcanic Complex
Mount Read Volcanics
volcaniclastic breccia
minor quartz
blocky clasts
dacite
indeterminate
indeterminate
sericite + quartz + pyrite + chlorite
foliated, semi-mylonitic, disseminated
pyrite, non-foliated domains of quartz +
calcite > sericite
local, enclosing mineralised lens
poor
intense
synmineralisation
hydrothermal and tectonic-metamorphic
HNS
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
6 5.30 K
2
O 5.35 Sr 15 Nb 29.4
TiO
2
0.53 P
2
O
5
0.12 Ba 475 Y 25
AI
2
O
3
15.00 S 3.31 Cu 22
Fe
2
O
3
6 .18 CO
2
0.88 Pb 30 Al 83
MnO 0.03 Total 99.32 Zn 100 CCPI 54
MgO 1.28 LOI 4.85 Sb 3.1 Ti/Zr 18.5
CaO 0.97 Tl 2.0
Na
2
O 0.37 Au 0.424 Zr 172
I
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 1 9
Intense, proximal, domainal quartz + sericite alteration facies
Footwali?
Hi 6
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
255053
intense, proximal, domainal quartz +
sericite (MV)
proximal altered zone enclosing and
transitional to mineralised MQ
Central V olcanic Complex
Mount Read V olcanics
indeterminate
dacite
indeterminate
indeterminate
quartz + sericite + pyrite
10-40 | jm microcrystalline sutured mosaic
of quartz, selective domainal sericite with
cleavage, disseminated pyrite
local, enclosing mineralised lens
poor
intense
synmineralisation
hydrothermal
Geochemistry
SiO
2
TiO
2
MA
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
S
co
2
88.10 K
2
O
0.44 P A
6 .29
1.21
0.01 Total
0.31 LOI
0.31
0.08 Au
1.94
0.03
0.54
0.37
99.6 3
1.44
0.318
Sr
Ba
Cu
Pb
Zn
Sb
Tl
Zr
9
100
913
24
100
1.6
0.9
178
Nb
Y
Al
CCPI
Ti/Zr
12.2
14
85
41
14.8
Hand specimen photograph Photomicrograph (xn)
2 2 0 | CHAPTER 7
intense, massive quartz alteration facies
Footwall?
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
255044
intense, massive quartz (MQ)
central siliceous core of A-zone
Central V olcanic Complex
Mount Read V olcanics
indeterminate
sparse polycrystalline quartz crystals?,
pseudomorphs of feldspar?
porphyritic?
dacite
indeterminate
indeterminate
quartz + calcite + pyrite + chalcopyrite
2 CM0 (jm microcrystalline in coarse
and fine equigranuiar domains of quartz,
disseminated calcite blebs, possible calcite
pseudomorphs after feldspar crystals,
pyrite + chlorite calcite veinlets
local, ore lens
poor
intense
synmineralisation
hydrothermal
Geochemistry
SiO2
TiO2
AI 2 O3
F e2 O3
MnO
MgO
CaO
N a2 O
87.70
0.39
0.80
1.17
0.07
0.12
5.29
0.05
K
2
O
P
2
O
5
S
co
2
Total
LOI
0.24
0.04
0.22
4.25
100.35
4.33
Au
Sr
Ba
Cu
Pb
Zn
Sb
Tl
0.048 Zr
23 Nb
100 Y
16 90
47 Al
100 CCPI
1.6 Ti/Zr
0.5
163
2.3
12
6
80
14.3
Hand specimen photograph Photomicrograph (xn)
7.9 | THALANGA: A POLYMETALLIC
SHEET-STYLE DEPOSIT
The Thalanga deposit is located near the western end of the
Early Ordovician Mount Windsor Subprovince (Fig. 1.7). It
was the most economically significant deposit in the Mount
Windsor Subprovince. The sulfide lenses were up to 25 m
thick and distributed over about 3000 m strike and 400 m
vertical extent. The pre-mining resource estimate was 6.6 Mt
grading 1.8% Cu, 2.6% Pb, 8.4% Zn, 69 g/t Ag and 0.4 g/t
Au.
Geological setting
The Thalanga deposit consisted of several semi-connected,
thin, stratabound and stratiform massive sulfide lenses hosted
in a distinctive quartz crystal-rich volcanic unit, which
is sandwiched between the underlying rhyolitic Mount
Windsor Formation and the overlying mixed andesitic-dacitic
Trooper Creek Formation (Fig. 7.23). The host unit (known
as the Thalanga horizon or favourable unit) is composed of
quartz + feldspar crystal-rich volcanic breccia, sandstone
and siltstone, and co-magmatic, peperitic quartz + feldspar
intrusions (Paulick and McPhie, 1999).
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 2 1
The ores are massive to semi-massive lenses dominated
by pyrite and sphalerite with variable proportions of galena,
chalcopyrite, pyrrhotite, magnetite and barite (Gregory et
al., 1990). Barite-rich zones exist in the up-dip and western
peripheries of the West and East Thalanga lenses. Chlorite +
tremolite + carbonate rocks, interpreted as metamorphosed
chlorite + carbonate alteration assemblages (Herrmann and
Hill, 2001), are closely associated with the West Thalanga
ore lenses. Magnetite-bearing quartzite bodies in the
peripheral or upper parts of some of the sulfide lenses, and
also intercalated with volcanic siltstone of the host unit to
the west, are interpreted to be metamorphosed exhalative
siliceous ironstones (Duhig et al., 1992).
The stratigraphic footwall is a laterally extensive,
1200 m thick, submarine rhyolitic succession. It is domin-
ated by sparsely quartz + feldspar-phyric coherent rhyolitic
lavas (e.g. data sheet TH1) and domes that may have
formed a low volcanic rise in the Thalanga area (Paulick and
McPhie, 1999). Rhyolitic hyaloclastite breccias and volcanic
sandstones are locally significant, particularly in the upper
part of the footwall beneath the western sulfide lenses. The
hanging wall succession is composed mainly of unaltered
to weakly altered coherent lavas and sills of feldspar-phyric
to aphyric dacite (e.g. data sheet TH2), and minor basaltic-
andesite. At Thalanga, it includes minor volcaniclastic rocks
FIGURE 7.23 | Schematic facies architecture of the submarine volcanic succession, from the Mount Windsor Volcanics,
through the Trooper Creek Formation, to the Rollston Range Formation, northwest of the Thalanga mine, Queensland.
Modified after Hill (1996 ).
2 2 2 | CHAPTER 7
of mixed dacitic-rhyolitic derivation, including lithic mass-
flow breccia, sandstone and massive to laminated cherty
siltstone, which increase in proportion westward. Paulick and
McPhie (1999) interpreted the volcanic facies assemblage
to indicate that the deposit formed in a below-storm-wave-
base environment on an elevated, lava-dominated, rhyolitic
centre. The compositions of the footwall and hanging wall
successions, respectively, indicate that they were rhyolitic
magmas derived from crustal melting, and mixed-mafic-felsic
magmas from subduction-modified mantle, in an extensional
back-arc-basin setting (Stolz, 1995).
Regional deformation and metamorphism, related to
Mid-Late Ordovician granitoid intrusions produced upper
greenschist facies metamorphic mineral assemblages and
a near-vertical foliation, particularly in phyllosilicate-rich
hydrothermally altered volcanic rocks.
Alteration facies and zonation
Underlying the Thalanga deposit is an extensive zone of strong,
pervasive quartz + sericite + pyrite chlorite alteration facies
(e.g. data sheet TH3) characterised by 14% disseminated
pyrite and an absence of primary feldspars. This alteration
style was pervasive in both clastic and coherent rhyolites,
and typically produced pseudoclastic breccia and mottled,
domainal alteration textures (Paulick and McPhie, 1999). The
zone extends beneath the entire strike length of the deposit
and is at least 200 m thick in the Central area, pinching out
to less than 50 m near the lateral and down-dip extremities
(Herrmann and Hill, 2001). It has a broad, upward flaring
shape and gradational boundaries with the surrounding least-
altered rhyolites.
Within the broad zone of feldspar destruction there are
semi-stratiform stringer zones of intense, pervasive quartz +
pyrite alteration facies (e.g. data sheet TH4) up to 50 m thick.
They extend obliquely up through the footwall at about 15
to the host unit and intersect it beneath the East, Central and
"eastern edge of the West Thalanga ore lenses, suggesting that
they were paths of maximum hydrothermal fluid flow. These
zones are composed essentially of quartz and up to 20% pyrite
in disseminated grains and anastomosing veins. They typically
contain less than 20% phyllosilicates (sericite chlorite).
Intense, macrocrystalline quartz + K-feldspar alteration facies
(e.g. data sheet TH5) exist in the immediate footwall, lateral
to the sulfide lens and stringer zone at East Thalanga, and also
stratigraphically lower in the footwall succession at Central
and West Thalanga.
Stratabound chlorite + dolomite altered zones, sub-
sequently metamorphosed to chlorite + tremolite carbonate
assemblages, formed in permeable volcaniclastic footwall
rocks close to the palaeo-seafloor and lateral to the West
Thalanga ore lenses (stratabound alteration facies; data sheets
TH6, 7 and 8). Local zones of non-pyritic, foliated altered
rhyolite with relict plagioclase (moderate, foliated sericite
+ chlorite alteration facies; data sheet TH9), exist within
the least-altered rhyolite, mainly around the peripheries of
the feldspar-destructive, strong quartz + sericite + pyrite +
chlorite alteration facies, and may represent low-temperature
hydrothermal recharge zones.
Ore genesis
There is consensus amongst researchers that Thalanga is
a sheet-like, synvolcanic, deformed and metamorphosed
VHMS deposit formed in a deep-marine back-arc rift.
Isotopic data suggests that the hydrothermal fluid and sulfur
were dominantly of seawater origin (Hill, 1996; Herrmann
and Hill, 2001). The hydrothermal system pervasively altered
a very broad zone in the mainly coherent rhyolitic footwall
succession to quartz + sericite + pyrite + chlorite assemblages.
The ore-forming fluids were focussed in low-angle quartz +
pyrite stringer zones. These pyritic stringer zones have a semi-
stratiform distribution, which suggests control by volcanic
facies related permeability contrasts. However, some appear
to cut through coherent rhyolite units and thus may represent
deformed synvolcanic fault zones (Paulick and McPhie,
1999).
A major proportion of the massive sulfide ore was
deposited in thin, extensive, stratiform and stratabound
lenses, either directly on the palaeo-seafloor or a few metres
below it. The distribution of massive pyrite and Cu-rich zones
suggests that the down-dip eastern parts of West and Central
Thalanga ore bodies, and central part of East Thalanga, were
sites of high-temperature hydrothermal discharge (Hill, 1996;
Paulick et al., 2001). Subordinate stratabound semi-massive
ore lenses were formed by subsurface replacement and/or
infilling of coarse volcaniclastic units of the host unit, which
were deposited by syneruptive, synhydrothermal mass flows
on top of the accumulating seafloor massive sulfide lenses.
Chlorite + carbonate (pre-metamorphic) alteration mineral
assemblages intimately associated with the West Thalanga
sulfide lenses, probably formed by mixing of hydrothermal
fluid and cold seawater in permeable volcaniclastic units
immediately below the palaeo-seafloor, in proximal to medial
parts of the hydrothermal discharge system. Apart from
disseminated and vein-type pyrite in the altered footwall
zones, all the sulfides were deposited at the top of the rhyolitic
Mount Windsor Formation, or in the quartz crystal-rich unit
that immediately overlies it. The massive, coherent dacite
lavas and sills of the hanging wall succession are essentially
unaltered and unmineralised; their emplacement appears to
have ended local hydrothermal circulation.
Weak, patchy quartz + sericite alteration facies
Least-altered footwal!
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 2 3
TH1
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
140802
weak, patchy quartz + sericite
Thalanga footwall
Mount Windsor Formation
Mount Windsor Subprovince
massive quartz + plagioclase-phyric
rhyolite
quartz and albitised plagioclase
phenocrysts
porphyritic
rhyolite
massive
rhyolite lava
quartz + sericite + biotite + chlorite
selective-pervasive (patchy)
microcrystalline quartz matrix, weakly
aligned sericite biotite in cleavage
regional, broadly stratabound, footwall
succession
moderate
weak
synvolcanic
diagenetic
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
76 .40 K
2
O 4.04 Rb 110 Tl 0.5
TiO2 0.11 P
2
O
5
0.02 Sr 82 Zr 146
AI
2
O
3
11.90 S 0.01 Ba 1056 Nb 13
Fe
2
O
3
1.6 4 CO
2
<0.1 Cu 5 Y 40
MnO 0.04 Total 98.52 Pb 20
MgO 0.6 7 LOI 0.6 0 Zn 48 Al 56
CaO 1.42 Sb 0.2 CCPI 25
N a, 0 2.27 Ti/Zr 4.5
2 2 4 | CHAPTER 7
Subtle, selective-pervasive quartz 4- albite alteration fa.cies
Least-altered hanging wall
TH2
Sample no.
Alteration fades
Location
Formation
Succession
Volcanic facies
Relict minerals
Relic textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
140799
subtle, selective-pervasive quartz +
albite
Thalanga hanging wall
Trooper Creek Formation
Mount Windsor Subprovince
massive sparsely plagioclase-phyric
dacite
plagioclase phenocrysts
weakly porphyritic, faintly flow banded
dacite
massive to flow banded
dacite lava or sill
quartz + albite (chlorite + actinolite
epidote)
selective-pervasive; mosaic of 20 pm
quartz + albite, dissemiated chlorite and
acicular prisms actinolite defining weak
relict flow banding, quartz calcite
veins
regional, broadly stratabound, hanging
wall succession
good
subtle
synvolcanic
diagenetic
Hand specimen photograph
Photomicrograph (xn)
Geochemistry
SiO
2
74.00 K
2
O 3.82 Rb 6 4 Tl 0.5
TiO
2
0.33 P
2
O
5
0.06 Sr 76 Zr 16 4
AI
2
O
3
12.70 S <0.01 Ba 1285 Nb 9
Fe
2
O
3
1.47 CO
2
Cu 3 Y 26
MnO 0.03 Total 98.50 Pb 11
MgO 0.52 LOI 0.47 Zn 29 Al 44
CaO 1.47 Sb 0.2 CCPI 19
Na
2
O 4.10 Ti/Zr 12.1
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 2 5
Strong, pervasive quartz + sericite + pyrite chlorite alteration fades
Footwall
TO 3
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
140808
strong, pervasive quartz + sericite +
pyrite chlorite
Thalanga footwall
Mount Windsor Formation
Mount Windsor Subprovince
massive quartz + plagioclase-phyric
rhyolite
quartz
porphyritic
rhyolite
massive
rhyolite lava
quartz + sericite + chlorite + pyrite
pervasive, microcrystalline quartz matrix
with strongly aligned sericite (cleavage),
scattered 1-2 cm elliptical chlorite-rich
domains
local, broadly stratabound in footwall
beneath entire Thalanga system,
>200 m thick, thinning laterally
poor
strong
synmineralisation
footwall hydrothermal
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
75.70 K
2
O 2.39 Rb 6 4 Zr 128
TiO
2
0.07 P
2
O
5
0.01 Sr 8 Nb 14
AI
2
O
3
11.40 S 0.6 5 Ba 326 Y 36
Fe
2
O
3
5.14 CO
2
<0.1 Cu 18
MnO 0.08 Total 98.03 Pb 11 Al 96
MgO 2.38 LOI 2.75 Zn 77 CCPI 73
CaO <0.01 Sb 0.2 Ti/Zr 3.3
Na
2
O 0.21 Tl <0.5
2 2 6 I CHAPTER 7
Intense, pervasive quartz + pyrite alteration facies
Footwall
TH4
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
140724
intense, pervasive quartz + pyrite
Thalanga footwall
Mount Windsor Formation
Mount Windsor Subprovince
massive quartz + plagioclase-phyric
rhyolite
quartz phenocrysts
porphyritic
rhyolite
massive
rhyolite lava
quartz + sericite + pyrite > biotite
chlorite
pervasive, mosaic of 100-200 pm quartz
and pyrite with interstitial shreds of semi-
aligned white mica > biotite
local, broadly stratabound in footwall,
discrete sheet-like zones at -15to host
unit, intersecting it beneath suifide lenses
poor
intense
synmineralisation
hydrothermal
Hand specimen photograph
Photomicrograph (xn)
Geochemistry
SiO
2
6 7.00 K
2
O 1.99 Sr 16 Nb 7
TiO
2
0.05 P
2
O
5
0.01 Ba 2300 Y 44
AI
2
O
3
6 .6 0 S 10.6 1 Cu 22
Fe
2
O
3
14.34 CO
2
<0.1 Pb 16 Al 97
MnO 0.02 Total 102.40 Zn 74 CCPI 88
MgO 1.6 7 LOI 8.24 Sb 1.1 Ti/Zr 3.8
CaO 0.05 Tl 1.9
Na
2
O 0.06 Rb 6 8 Zr 79
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 2 7
intense, microcrystalline quartz + K-feldspar alteration facies
Footwall
TH5
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
140902
intense, microcrystalline quartz + K-
feldspar
Thalanga footwall
Mount Windsor Formation
Mount Windsor Subprovince
massive quartz + plagioclase-phyric
rhyolite
quartz and albitised plagioclase
phenocrysts
porphyritic
rhyolite
massive
rhyolite lava
quartz + K-feldspar > trace pyrite
selective-pervasive; matrix of 10-50 pm
microcrystalline quartz + K-feldspar,
albitised plagioclase, quartz veins with
feldspar-alteration selvage
local, lozenge shaped zones, broadly
stratabound in footwall
poor
intense
synmineralisation
hydrothermal
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
84.20 K
2
O 5.87 Sr 42 Nb 9
TiO
2
0.06 P
2
O
5
0.01 Ba 1936 Y 20
AI
2
O
3
7.40 S 0.48 Cu 3
Fe
2
O
3
0.6 0 CO
2
Pb 70 Al 93
MnO <0.01 Total 99.15 Zn 118 CCPI 9
MgO 0.08 LOI 0.49 Sb 0.4 Ti/Zr 4.5
CaO 0.11 Tl 1.5
Na
2
O 0.34 Rb 94 Zr 80
2 2 8 | CHAPTER 7
Intense, pervasive, stratabound chlorite + tremolite alteration facies
Footwall
TH8
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
145401
intense, pervasive, stratabound chlorite
+ tremolite
Thalanga footwall
Mount Windsor Formation
Mount Windsor Subprovince
quartz + plagioclase-phyric rhyolitic
volcaniclastic breccia?
nil
nil
rhyolite
indeterminate
phlogopite + chlorite + tremolilte > minor
pyrite, sphalerite, chalcopyrite
pervasive, foliated phlogopite chlorite,
30% coarse prisms and bands of
randomly oriented tremolite
local, stratabound, typically immediately
below sulfide lenses
nil
intense
synmineralisation
footwall hydrothermal
Hand specimen photograph Photomicrograph (xn)
Geochemistry
SiO
2
44.57 K
2
O 3.48 Rb Zr 171
TiO
2
0.135 P
2
O
5
0.04 Sr Nb 10
AI
2
O
3
13.06 S 1.6 6 Ba 38300 Y 32
Fe
2
O
3
3.88 CO
2
0.30 Cu 1500
MnO 0.15 Total 91.89 Pb 200 Al 75
MgO 17.6 7 LOI 3.23 Zn 5900 CCPI 85
CaO 6 .76 Sb Ti/Zr 4.7
Na
2
O 0.18 Tl
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS I 2 2 9
Intense, stratabound. pervasive chlorite + tremolite + calcite alteration facies
Host rock
TH7
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
145417
intense, stratabound, pervasive chlorite +
tremolite + calcite
Thalanga host rock
Mount Windsor Formation
Mount Windsor Subprovince
quartz + plagioclase-phyric rhyolitic
volcaniclastic breccia?
nil
nil
rhyolite
indeterminate
?
chlorite + tremolite + calcite > pyrite +
chalcopyrite + sphalerite + galena
pervasive; coarse interlocking prisms of
tremolite with 10% interstitial sulfides and
ragged calcite patches
local, stratabound, proximal to medial,
closely associated with or lateral to West
Thalanga sulfide lenses
nil
intense
synmineralisation
hydrothermal
Geochemistry
SiO
2
33.10 K
2
O 0.58 Rb Zr 89
TiO
2
0.06 0 P
2
O
5
0.15 Sr Nb 6
AI
2
O
3
6 .48 S 4.82 Ba 8700 Y 13
Fe
2
O
3
4.82 CO
2
7.10 Cu 7500
MnO 0.57 Total 92.41 Pb 4200 Al 55
MgO 19.00 LOI 5.86 Zn 23800 CCPI 97
CaO 15.6 0 Sb Ti/Zr 4.0
Na
2
O 0.13 Tl
Hand specimen photograph Photomicrograph (xn)
2 3 0 I CHAPTER 7
Intense, strataboynd tremolite + dolomite + calcite alteration facies
Host rock
TH8
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relic textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
145418
intense, stratabound tremolite + dolomite
+ calcite
Thalanga host rock
Mount Windsor Formation
Mount Windsor Subprovince
quartz + plagioclase-phyric rhyolitic
volcaniciastic breccia?
nil
nil
rhyolite
indeterminate
?
dolomite + calcite > minor tremolite
+ chlorite + pyrite + chalcopyrite +
sphalerite + galena
pervasive; mosaic of 1 mm sutured
spheroidal dolomite and interstitial calcite
disseminated sulfides and sparse ragged
tremolite prisms
local, stratabound, proximal to medial,
closely associated with or lateral to West
Thalanga sulfide lenses
nil
intense
synmineralisation
footwall hydrothermal, seawater mixing?
Hand specimen photograph
Photomicrograph (xn)
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS I 2 3 1
Moderate, foliated sericite + chlorite alteration fades
FootwaSI
TH9
Hand specimen photograph Photomicrograph (xn)
Sample no. 140727
Alteration facies moderate, foliated sericite + chlorite
Location Thalanga footwall
Formation Mount Windsor Formation
Succession Mount Windsor Subprovince
V olcanic facies massive quartz + plagioclase-phyric
rhyolite
Relict minerals quartz and albitised plagioclase
Relict textures porphyritic
Primary composition rhyolite
Lithofacies massive, foliated
Interpretation rhyolite lava
Alteration minerals sericite + quartz + biotite + chlorite
Alteration textures selective-pervasive 5 0-1 00 pm
microcrystalline quartz matrix in
<1 mm lenses wrapped by foliated
sericite biotite (augen texture),
broken grains, cleavage, sericite
altered plagioclase
Distribution local, stratabound in upper part of
medial to distal footwall, particularly
down dip of ore zones
Preservation moderate
Alteration intensity moderate
Timing synmineralisation
Alteration style hydrothermal, tectonic-metamorphic
Geochemistry
SiO
2
70.90 Na
2
O 1.05 Rb 145 Zr 16 2
TiO
2
0.10 K
2
O 3.92 Sr 40 Nb 19
AI
2
O
3
14.50 P
2
O
5
0.01 Ba 952 Y 54
Fe
2
O
3
1.6 9 S <0.01 Cu 3
MnO 0.04 CO
2
0.10 Pb 5 Al 87
MgO 3.94 Total 96 .37 Zn 231 CCPI 52
CaO 0.12 LOI 2.6 7 Sb 0.2 Ti/Zr 3.7
Tl 2.6
2 3 2 | CHAPTER 7
7.10 | HIGHWAY-REWARD: A PIPE STYLE
Cu-Au VHMS DEPOSIT
The Highway-Reward Cu-Au deposit, in the central part of the
Mount Windsor Subprovince (Fig. 1.8), represents a contrast
in style of deposit and stratigraphic setting. It consists of two
discordant, vertical pipe-like bodies of massive pyrite about
200 m apart, hosted in the proximal facies of a non-explosive,
submarine felsic volcanic centre located near the top of the
Trooper Creek Formation (Fig. 7.23).
Geological setting
The lithofacies association at Highway-Reward represents
a deep marine intrusion-dominated felsic volcanic centre.
Doyle and McPhie (2000) recognised at least 13 coherent
feldspar- and quartz + feldspar-phyric dacitic to rhyolitic
synvolcanic sills, small cryptodomes (e.g. data sheets HR1
and 4) and lavas in the immediate area. The abundance and
complex overlapping relationships of coherent intrusive units
indicate a proximal volcanic setting. Thin volcanic sandstone
and siltstone units and thicker units of crystal- and pumice-
rich sandstone and breccia separate the intrusions. The crystal-
and pumice-rich facies were mainly derived from explosive
eruptions and deposited in the submarine basin from water-
supported gravity flows. The succession is upright and dips at
20-30 to the southeast.
Massive pyrite chalcopyrite exists in two vertical pipe-
like bodies about 150 m apart. Both are discordant to bedding,
parallel to a locally developed northeast trending sub-vertical
cleavage (S
4
) and have irregular-amoeboid outlines with plan
dimensions of about 200 x 75-150 m. The western pipe
(Highway) has a vertical extent of 250 m and the eastern pipe
(Reward) of at least 350 m (Doyle and Huston, 1999). They
are dominantly composed of fine-grained (<0.5 mm) pyrite
with interstitial chalcopyrite, minor tennantite, sphalerite,
quartz and sericite, and traces of chlorite, galena, barite,
hematite and aikinite (PbCuBiS
3
). The massive sulfide pipes
are intersected by chalcopyrite, barite, quartz + carbonate and
anhydrite veins, and contain inclusions of quartz + sericite +
pyrite altered volcanic rocks in their margins. The Highway
and Reward massive sulfide pipes contain approximately
2 Mt and 5 Mt of pyrite, respectively. They include hypogene
sulfide resources estimated at 1.2 Mt @ 5.5% Cu, 1.2 g/t Au
and 6.5 g/t Ag in the Highway pipe and 0.2 Mt @ 3.5% Cu,
1 g/t Au and 13 g/t Ag in the Reward pipe.
The pipes are enveloped by a broad 200 x 500 m halo of
vein and disseminated low-grade Zn + Pb + Ba sulfides. Within
that are several small zones of massive to laminated sphalerite
+ pyrite + galena + chalcopyrite + barite. A 20-30 m thick
stratabound lens of sphalerite-rich massive sulfide exists in
volcaniclastic rocks 50 m above and south of the Reward
pipe. It has a pyrite-rich base that thickens northwards into
a discordant lens of massive pyrite lying above the southern
edge of the Reward pipe. Sphalerite-rich sulfides also exist
locally in narrow discordant zones at the margins of the main
Highway and Reward massive sulfide pipes.
Alteration facies and zonation
A discordant zone of feldspar-destructive hydrothermal alter-
ation envelopes the massive sulfide pipes. It has an elliptical
area of 500 x 250 m in plan and extends from 60 m above
to at least 150 m below the massive sulfide bodies (Doyle
and Huston, 1999). Doyle and Huston's (1999) alteration
zonation is here simplified down to six alteration facies.
Intense, stringer quartz + sericite + pyrite alteration facies
(e.g. data sheet HR8), locally flanked by intense, pervasive
chlorite + pyrite alteration facies (e.g. data sheet HR7) occupy
feeder zones which extend vertically beneath both pipes and
possibly meet at depth. Zones of similar quartz + sericite +
pyrite altered rocks extend into the hanging wall above the
southern parts of both massive sulfide pipes (Doyle and
Huston, 1999). These intensely altered zones pass laterally
outwards to enveloping zones of intense sericite + quartz +
pyrite (e.g. data sheet HR6) and strong, pervasive chlorite
+ sericite + quartz + pyrite alteration facies (e.g. data sheet
HR5). These locally enclose, and in turn grade laterally and
upwards in to, non-pyritic zones of moderate, pervasive
chlorite alteration facies (data sheet HR3). The weak, regional,
selective albite hematite alteration facies (e.g. data sheet
HR2) exists at greater than 50200 m from the massive sulfide
pipes. It comprises two sub-facies with mineral assemblages
of feldspar + carbonate quartz chlorite + sericite and
hematite + quartz sericite chlorite albite, which are
regionally distributed in the Trooper Creek Formation and
are respectively attributed to alteration during diagenesis and
synvolcanic low-temperature fluid convection.
Ore genesis
The deposits were initially thought to have had a two-stage
origin (Beams et al., 1998). The stratiform Zn-rich zone was
interpreted as a syngenetic Cambro-Ordovician sulfide lens
and the pyrite + chalcopyrite pipes as Siluro-Devonian syn-
deformational deposits, because of their discordance to host
volcanic rocks, parallelism to the youngest cleavage (S
4
) and
the observation that anhydrite overprinted the dominant
S
3
cleavage. However, Doyle and Huston (1999) refuted
this microtextural relationship and argued for a syngenetic
volcanic-associated, subseafloor replacement origin for the
massive sulfide pipes. Lead isotopic ratios, the gradation from
stratiform Zn-rich sulfides into discordant Cu-rich massive
pyrite, relict framboidal sulfide textures, hydrothermal alter-
ation facies and their relationships to primary volcanic facies,
and the S
3
tectonic overprint are all consistent with Early
Ordovician synvolcanic formation of all the sulfide zones.
The Highway-Reward massive sulfide pipes have some
similarities with disseminated to massive Cu-Au deposits in
the Mount Lyell field (Large et al., 2001c). The similarities
include metal ratios, dominance of pyritic subseafloor
replacement style mineralisation and low 5
34
S values; mostly
in the range 5 to 7.5%o at Highway-Reward and 5 to 10%o at
Mount Lyell (Solomon etal., 1969; Doyle and Huston, 1999).
Given the emerging evidence for involvement of magmatic
fluids at Mount Lyell (Corbett, 2001; Huston and Kamprad,
2001) it is reasonable to similarly classify Highway-Reward as
a hybrid seawater-magmatic hydrothermal system.
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS I 2 33
Weak, selective-pervasive quartz + sericite +albite alteration facies
Least-altered rhyolite
H R 1
Sample No.
Alteration Facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
L ithofacies
I nterpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
137068
weak, selective-pervasive quartz + sericite
+ albite
upper medial, 50 m east of Highway pipe
(1 007 5 N )
Trooper Creek Formation
Mount Windsor Subprovince
massive quartz + plagioclase-phyric
rhyolite
quartz, plagioclase
porphyritic
rhyolite
massive
partly extrusive cryptodome
quartz + sericite + albite? > (calcite, pyrite)
selective-pervasive, microcrystalline
groundmass, albite + sericite or calcite-
altered plagioclase
regional
moderate
weak
diagenetic
Geochemistry
SiO2
TiO2
AI 2 O3
F e2 O3
MnO
MgO
CaO
N a2 O
75.08
0.30
12.75
1.57
0.08
2.02
0.6 1
2.16
K
2
O
S
co
2
Total
LOI
2.22
0.06
0.51
97.37
2.88
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
5 0 Zr
4 6 Nb
1382 Y
43
14 Al
104 CCPI
2.2 Ti/Zr
1.0
158
9
2 2
6 0
4 4
11.4
Hand specimen photograph Photomicrograph (xn)
2 3 4 | CHAPTER 7
Weak, regional, selective albite + hematite alteration facies
HR2
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
137105
weak, regional, selective albite + hematite
upper periphery, 250 m east of Reward
pipe
Trooper Creek Formation
Mount Windsor Subprovince
massive quartz + plagioclase-phyric
rhyolite
quartz, piagiociase
porphyritic, giomeroporphyritic piagiociase,
microcrystalline groundmass
rhyolite
massive
synvolcanic sill
quartz + albite + chlorite > (sericite +
calcite + hematite)
selective-pervasive in groundmass; 20-6 0
urn microcrystalline quartz + albite
calcite-altered piagiociase, disseminated
chlorite and hematite patches
regional
good
weak
synvolcanic
diagenetic
G eochemistry
SiO
2
TiO
2
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
70.58
0.33
13.75
2.41
0.05
2.26
1.78
6 .51
K
2
O
S
CO
2
Total
LOI
0.34
0.06
<0.01
98.07
2.42
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
8 Zr
142 Nb
92 Y
<2
5 Al
38 CCPI
0.8 Ti/Zr
<0.5
16 1
9
21
24
39
12.3
Hand specimen photograph Photomicrograph (xn)
Moderate, pervasive chlorite alteration facies
LOCAL HYDROTHERMAL ALTERATION RELATED TO VHMS DEPOSITS | 2 3 5
HR3
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
137079
moderate, pervasive chlorite
upper proximal zone, between Highway
and Reward sulfide pipes
Trooper Creek Formation
Mount Windsor Subprovince
massive quartz + plagioclase-phyric
rhyolite
quartz, mafic phenocrysts?
porphyritic, amygdaloidal?
rhyolite
massive
partly extrusive cryptodome
quartz + sericite + chlorite
pervasive, microcrystalline mosaic of
quartz + chlorite + sericite, sericite
pseudomorphs after plagioclase
phenocrysts, anastomosing wispy sericite
foliation, recrystallised overgrowths on
quartz
local; medial to proximal zones laterally
equivalent to upper parts of sulfide pipes
moderate to poor
moderate
synmineralisation
hydrothermal
Geochemistry
SiO
2
TiO
2
AI
2
O
3
F e
2
O
3
MnO
MgO
CaO
N a
2
O
71.21
0.30
14.09
3.40
0.15
3.59
0.39
0.16
K
2
O
S
CO
2
Total
LOI
3.27
0.05
0.01
96 .6 1
3.6 1
Rb
Sr
Ba
Cu
Pb
Zn
Sb
70 Tl
21 Zr
1791 Nb
6 Y
7
120 Al
0.8 CCPI
Ti/Zr
1.5
16 5
9
23
93
66
10.9
Hand specimen photograph Photomicrograph (xn)
2 36 I CHAPTER 7
Weak, pervasive albite + sericite alteration facies
Least-altered dacife
HR4
Sample no.
Alteration facies
Location
Formation
Succession
Volcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
136 919
weak, pervasive albite + sericite
Highway, medial footwall
Trooper Creek Formation
Mount Windsor Subprovince
massive, sparsely plagioclase-phyric
massive dacite
plagioclase <1mm
porphyritic and micropoikilitic
dacite
massive to weakly flow banded
cryptodome
albite + chlorite + sericite > (zeolite?,
quartz)
pervasive groundmass, microcrystalline
partly preserving micropoikilitic texture,
albite sericite-altered plagioclase,
chlorite veinlets
regional
good
weak
synvolcanic
diagenetic
Geochemistry
SiO
2
TiO
2
AI2O
3
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
6 6 .74
0.55
16 .6 3
4.30
0.18
2.48
0.19
4.40
K
2
O
P2O5
S
co
2
Total
LOI
1.81
0.10
0.00
97.38
2.42
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
44 Zr
6 8 Nb
515 Y
1
2 Al
16 2 CCPI
0.3 Ti/Zr
<0.5
16 1
9
25
48
51
20.5
Hand specimen photograph Photomicrograph (xn)
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 2 3 7
Strong, perfasiwe chlorite + sericite + quartz + pyrite alteration facies HR 5
Sample no.
Alteration facies
Location
Formation
Succession
V olcanic facies
Relict minerals
Relict textures
Primary composition
Lithofacies
Interpretation
Alteration minerals
Alteration textures
Distribution
Preservation
Alteration intensity
Timing
Alteration style
137127
strong, pervasive chlorite + sericite +
quartz + pyrite
Highway footwall, 100 m east of stringer
zone
Trooper Creek Formation
Mount Windsor Subprovince
massive quartz + plagioclase-phyric
rhyodacite
quartz + altered plagioclase
porphyritic
rhyodacite
massive
cryptodome
quartz + sericite + pyrite + chlorite rutile
pervasive in groundmass, millimetre
patches of microcrystalline quartz and
wispy domains of aligned sericite, some
broken quartz phenocrysts
local; medial to proximal zones laterally
equivalent to upper parts of sulfide pipes
poor
strong
synmineralisation
hydrothermal
Geochemistry
SiO
2
TiO
2
AI
2
O
3
Fe
2
O
3
MnO
MgO
CaO
Na
2
O
6 4.55
0.44
17.02
5.54
0.04
1.87
0.20
0.18
K
2
O
S
co
2
Total
LOI
4.53
0.08
2.97
97.43
4.93
Rb
Sr
Ba
Cu
Pb
Zn
Sb
Tl
92 Zr
28 Nb
36 83 Y
16
32 Al
6 8 CCPI
2.0 Ti/Zr
2.0
149
8
25
94
59
17.8
Hand specimen photograph Photomicrograph (xn)
2 3 8 | CHAPTER 7
Intense sericite + quartz + pyrite alteration facies
HR
Hand specimen photograph Photomicrograph (xn)
Sample no. 137080
Alteration fades intense sericite + quartz + pyrite
Location upper proximal, 20 m east of Highway pipe
Formation Trooper Creek Formation
Succession Mount Windsor Subprovince
V olcanic facies massive, sparsely plagioclase-phyric
dacite
Relict minerals altered plagioclase
Relict textures porphyritic
Primary composition dacite
Lithofacies massive
Interpretation cryptodome
Alteration minerals quartz + sericite + pyrite
Alteration textures pervasive, polycrystalline quartz
pseudomorphs after plagioclase,
microcrystalline matrix of quartz + sericite
> pyrite
Distribution local, proximal zone enveloping Highway
sulfide pipe
Preservation poor
Alteration intensity intense
Timing synmineralisation
Alteration style hydrothermal
Geochemistry
SiO
2
73.8 2 K
2
O 3.6 5 Rb 78 Zr 107
TiO
2
0.41 P
2
O
5
0.06 Sr 14 Nb 6
AI
2
O
3
12.23 S 2.89 Ba 2222 Y 13
Fe
2
O
3
4.55 CO
2
Cu 1385
MnO 0.02 Total 98.56 Pb 10 Al 96
MgO 0.76 LOI 4.09 Zn 90 CCPI 57
CaO 0.09 Sb 1.3 Ti/Zr 23.0
Na
2
O 0.08 Tl 5.3
Intense, pervasive chlorite + pyrite alteration facies
Footwall
LOCAL HYDROTHERMAL ALTERATION RELATED TO V HMS DEPOSITS | 2 3 9
HR7
Hand specimen photograph Photomicrograph (xn)
Sample no. 137083
Alteration fades intense, pervasive chlorite + pyrite
Location Highway footwall, adjacent to stringer
zone
Formation Trooper Creek Formation
Succession Mount Windsor Subprovince
Volcanic fades altered dacitic pumice breccia?
Relict minerals nil
Relict textures nil
Primary composition dacite
Lithofacies indeterminate
Interpretation indeterminate
Alteration minerals chlorite + pyrite > quartz
Alteration textures pervasive cryptocrystalline groundmass
or matrix of chlorite, cleavage, 5%
disseminated euhedral pyrite with quartz
pressure shadows
Distribution local; narrow zones enveloping footwall
quartz + pyrite stringer zone
Preservation nil
Alteration intensity intense
Timing synmineralisation
Alteration style hydrothermal
Geochemistry
SiO
2
24.90 K
2
O 0.01 Rb <1 Zr 16 9
TiO
2
0.6 1 P
2
O
5
0.12 Sr 9 Nb 9
AI
2
O
3
18.79 S 6 .37 Ba 21 Y 27
Fe
2
O
3
16 .73 CO
2
Cu 102
MnO 0.35 Total 90.83 Pb 12 Al 99
MgO 22.70 LOI 13.80 Zn 522 CCPI 100
CaO 0.21 Sb Ti/Zr 21.7
Na
2
O 0.04 Tl
2 4 0 | CHAPTER 7
Intense, stringer quartz + sericite + pyrite alteration fades
Footwali
HR8
Hand specimen photograph
Stacked SWIR spectra (hull quotient)
Photomicrograph (xn)
TiO
2
-Zr immobile element plot
Sample no. 137129
Alteration facies intense, stringer quartz + sericite + pyrite
Location Reward footwall stringer zone
Formation Trooper Creek Formation
Succession Mount Windsor Subprovince
Volcanic facies massive
Relict minerals nil
Relict textures nil
Primary composition rhyodacite
Lithofacies massive
Interpretation synvolcanic sill?
Alteration minerals quartz + sericite + pyrite
Alteration textures pervasive, irrregular sericite
pseudomorphs after feldspar in <10 pm
microcrystalline quartz, interstitial sericite
groundmass, disseminated 5-10%
euhedral pyrite
Distribution local; footwall stringer zones beneath
sulfide pipes
Preservation nil
Alteration intensity intense
Timing synmineralisation
Alteration style hydrothermal
Geochemistry
SiO
2
75.78 K
2
O 1.80 Rb 30 Zr 52
TiO
2
0.18 P
2
O
5
0.02 Sr 13 Nb 3
AI
2
O
3
5.87 S 7.16 Ba 1404 Y 8
Fe
2
O
3
9.89 CO
2
Cu 11
MnO 0.02 Total 101.21 Pb 3 Al 90
MgO 0.26 LOI 5.81 Zn 12 CCPI 83
CaO 0.20 Sb 1.2 Ti/Zr 20.8
Na
2
O 0.03 Tl 1.2
I 241
8 | FINDING ORE DEPOSITS IN ALTERED
VOLCANIC ROCKS
Recognising alteration facies that may be genetically related to
ore deposits is an important step in mineral exploration. Even
more helpful is the ability to identify alteration facies that
are likely to be associated with particular ore deposit types
and thus prioritise exploration targets. The characteristics of
alteration facies have the potential to be used as exploration
vectors, guiding explorers to the most prospective altered
zones in a system, and thereby enabling efficient and early
testing of the best targets, avoiding expensive, protracted
exploration programs, and improving the chance of success.
The processes that alter volcanic facies and the range of
textures and mineral assemblages they produce are complex
and challenging. As described in previous chapters, there
are a variety of alteration processes, which can produce a
broad range of alteration mineral assemblages and textures.
Ancient volcanic rocks commonly contain a complex
assemblage of overprinting alteration minerals and textures,
which reflect multiple episodes of alteration by a variety of
processes: diagenesis, hydrothermal alteration, deformation,
metamorphism or weathering.
In early Palaeozoic volcanic regions, like the Mount
Read Volcanics, western Tasmania and the Mount Windsor
Subprovince, north Queensland, patience and experience
are required to unravel the complexities of altered rocks
and recognise those altered zones that are 'red-herrings'
to mineral explorers. In fact, numerous geologists have
initially doubted that the foliated, weathered and moss-
covered rocks encountered in western Tasmania really were of
volcanic origin. Several intensive, protracted and ultimately
unsuccessful exploration programs have been conducted in
the Mount Read Volcanics on unfavourable altered zones.
On the other hand, there may be altered zones that remain
under-explored because favourable alteration facies were
not recognised. Recognising the occurrence of altered rocks
and identifying favourable or prospective alteration facies
and zones are important steps toward minimising risk and
expenditure during exploration in these environments.
This chapter draws together the descriptive and
geochemical techniques described in Chapters 2, 3 and 4,
and an understanding of the different alteration processes that
modify submarine volcanic successions. It proposes methods
for discriminating alteration facies associated with particular
processes, identifying favourable altered zones for mineral
exploration, and guiding exploration within those zones
toward potentially mineralised areas.
8.1 | PRINCIPLES OF DISCRIMINATING
BETWEEN DIAGENETIC,
HYDROTHERMAL AND
METAMORPHIC ALTERATION
FACIES
Diagenetic facies
As discussed in Chapter 5, the characteristics of diagenetic
facies are:
They are typically widespread with district or regional-scale
distribution.
At local scales, they display variable alteration intensity
and patchy distribution. This is mainly controlled by
distribution of coherent versus clastic volcanic facies,
and variations in the primary composition, permeability,
porosity and the proportion of glassy to crystalline facies.
They occur in vertically-stacked, extensive, sub-horizontal
altered zones, which have mineral assemblages that reflect
increasing temperature with depth.
They have undergone relatively minor (< 10 wt%) chemical
changes that are predominantly in response to hydration or
alkali-exchange reactions between the volcanic facies and
modified seawater. Mass transfers are generally small, an
order of magnitude less than those in intense hydrothermal
alteration facies. The scale of migration of elements is also
small (millimetres to tens of centimetres) and thus on a
larger scale (i.e. basin scale) the changes are essentially
isochemical.
Their mineralogical and textural changes vary from subtle
to strong. Quartz phenocrysts, for example, are relatively
stable and commonly well preserved, whereas mafic
phenocrysts and volcanic glass are relatively unstable and
typically completely altered.
These changes are commonly overprinted or obscured
2 4 2 I CHAPTER!
by subsequent alteration as diagenesis is often the earliest
preserved post-emplacement process.
Metamorphic facies
Metamorphic facies share some characteristics with diagenetic
facies, but also differ significantly in these ways:
Distribution varies in scale: contact metamorphic facies
associated with intrusions may be as narrow as a few
centimetres and up to several kilometres wide. Regional
metamorphic facies (either burial metamorphism or
metamorphism associated with deformation) can be tens
or hundreds of kilometres wide and several kilometres
thick.
Metamorphic facies are uniform and pervasive: they are
not typically patchy at a scale of metres to tens of metres.
Primary volcanic textures have virtually no influence on
high-grade metamorphic facies, which are principally
determined by whole-rock compositions and metamorphic
conditions.
Chemical changes are minor; metamorphism is generally
a process of phase-change in response to changing
temperature and pressure at low water-rock ratios, which
limits the redistribution of chemical components in and
out of the system. The most common metamorphic
reactions are dehydration and decarbonation reactions. The
composition and mineralogy of metamorphic facies are
generally strongly influenced by the primary composition
of volcanic facies.
Mineralogical and textural changes vary from subtle
to intense depending on the degree of metamorphism.
Typically, primary volcanic quartz phenocrysts are well
preserved up to about amphibolite grade, but fine-grained
or glassy facies and some mafic phenocrysts are unlikely
to survive even low grades of zeolite and greenschist facies
or contact metamorphism. Metamorphic re-crystallisation
produces a wide variety of distinctive textures, such as
granoblastic, porphyroblastic, decussate, schistose, and
gneissic, which are not easily confused with primary
volcanic or diagenetic textures.
Hydrothermal alteration facies
Hydrothermal alteration facies are unlike diagenetic and
metamorphic facies in their potential for major compositional
change. This is because hydrothermal alteration typically
involves large volumes of fluid, which facilitate large-scale
mass transfers into, out of, or around hydrothermal systems.
Depending on the intensity of alteration, this characteristic
determines or limits the other characteristics of hydrothermal
alteration facies.
Hydrothermal alteration facies generally have local
distribution, limited to tens or hundreds of metres and
rarely exceeding a few kilometres.
Hydrothermally altered zones commonly have high aspect
ratios (i.e. narrow lateral and great vertical extents) because
convecting, typically ascending, fluids produce them.
Locally, on small-scales, the distribution of hydrothermal
alteration facies is generally uniform, or pervasive. However,
the distribution is mainly dependent on permeability and
porosity; therefore hydrothermal alteration facies may be
restricted to fractures and vein selvedges in coherent or
otherwise impermeable rocks.
The degrees of mineralogical and textural preservation, and
chemical changes are extremely dependent on alteration
intensity and pre-hydrothermal alteration composition and
texture of the facies. Pre-existing textures and minerals are
less likely to be preserved in proximal zones of hydrothermal
systems, through which hot reactive fluids are flushed,
than in peripheral zones with lower temperature, partly
neutralised fluids and lower fluid-rock ratios. As in the other
types of alteration, quartz crystals in felsic volcanic facies
tend to survive intense alteration, except where major loss
of silica is involved (e.g. in chlorite zones). Other primary
crystal phases are commonly progressively replaced (e.g.
feldspars altered to sericite) and may be useful as indicators
of alteration intensity. Hydrothermal alteration facies rarely
preserve primary textures in originally glassy facies.
Hydrothermal mineral assemblages are largely controlled
by fluid composition and physicochemical conditions, and
are not noticeably influenced by primary compositions; at
least in the intensely altered zones, which had high fluid-
rock ratios. Thus, an intensely hydrothermally altered zone
may cut across volcanic lithofacies of different primary
compositions and textures (e.g. coherent andesite and
rhyolitic breccia) and comprise only one alteration facies
in which the protoliths are mineralogically and texturally
indistinguishable.
Hydrothermal alteration commonly involves significant
mass transfer of chemically mobile elements. Elements
may be gained through precipitation or lost through
dissolution. These mass transfers may produce large
positive or negative net mass changes within particular
alteration facies (generally with implications for volume
change) or balance each other out to produce negligible net
change. Significant mass changes are commonly evident
in composition data and derivative alteration indices. For
example, Na depletion typically accompanies hydrolysis
and sericitisation of plagioclase. However, substantial mass
changes in some major elements are commonly obscured
by the constant sum effect; this applies especially to Si.
Major chemical modifications are frequently reflected in
exotic mineral assemblages. For example, VHMS-related
alteration facies commonly contain disseminated pyrite
or base-metal sulfides, and several types of Zn deposits are
associated with Mn-rich mineral assemblages.
It is important for economic geologists to recognise
hydrothermal alteration facies, which may indicate the large-
scale transport and deposition of economically valuable
elements, and to discriminate these from alteration facies that
result from other alteration processes that are unrelated to
ore deposition. In some cases, examination of an individual
altered sample can reveal important facts that help to identify
the alteration process. For example, a rock with gneissic fabric
is metamorphic; a rock composed essentially of quartz and
pyrite is probably of hydrothermal origin. However, alteration
textures and mineral assemblages may not easily distinguish
some weak hydrothermal alteration facies, perhaps in
peripheral zones, from diagenetic or metamorphic facies.
One of the main criteria distinguishing hydrothermal
from other alteration facies is the distribution or extent of
the altered zone. This cannot be determined by observation
of an individual specimen or outcrop. It requires systematic
prospect-scale mapping and knowledge of the district-scale
geological context. Unfortunately, in the last decade of the
20
th
century there has been a significant decline in in-field
and on-ground geological data collection, particularly in the
mineral exploration and mining industries. There is a trend
towards using high technology remote sensing to rapidly
explore large areas at continuously improving resolutions.
However, to ensure meaningful interpretation of these data,
it is imperative that this virtual geology is not disconnected
from real rocks. The combination of a variety of criteria, and
high-quality mapping, will lead to the best interpretation of
alteration facies.
8.2 | EXPLORATION VECTORS AND
PROXIMITY INDICATORS
Mineral zonation
Mapping of sulfide distribution, particularly pyrite, is an
important exploration technique. Sulfide abundances are
easily estimated by eye, even in weathered samples, and should
be applied at an early stage of exploration wherever geological
exposure permits. VHMS deposits commonly have extensive
footwall zones of disseminated pyrite. For example, the Mount
Lyell Cu-Au deposits (western Tasmania) all lie in a zone of
greater than 1 % disseminated pyrite, which is 6 km long and
1 km wide at the surface (Corbett, 2001). Such pyritic zones
provide very large exploration targets for initial area selection.
They have the potential to be delineated into high-abundance
zones, in order to reduce the size of the targets for intensive
exploration and drill testing. Interpretation of sulfide vectors
is straightforward: more is better, and sulfide proportions
generally increase with proximity to sulfide deposits.
Other components of alteration mineral assemblages that
are easily recognisable in all sample types and may be spatially
zoned around mineral deposits include silicates, carbonates
and Fe-oxides. The ratios of quartz to phyllosilicates,
sericite to chlorite, and carbonate to silicates are commonly
systematically zoned around VHMS deposits, and recognition
of the zonation patterns can provide useful exploration
vectors, at least on a prospect scale.
Unfortunately, interpretation of the patterns is rather
complex. Australian VHMS deposits are typically associated
with siliceous proximal zones (Section 7.4) but there are many
variations even within mineral fields and districts. Therefore
it is unwise to be too strictly empirical or model-driven
in applying this approach. It is better to map out mineral
distributions and relate them to alteration intensities, rather
than rely on the recognition of specific zonation patterns,
which may relate to an ore deposit model. Carbonate +
chlorite assemblages, for example, are indicators of ore
proximity in some VHMS deposits, such as Rosebery, Hellyer
and Thalanga deposits, but only occur in the peripheral or
least-altered zones of the Western Tharsis deposit (Mount
Lyell field). Therefore, rigidly applying a carbonate + chlorite
FINDING ORE DEPOSITS IN ALTERED VOLCANIC ROCKS | 2 4 3
vector could be misleading and potentially guide exploration
away from some Cu-Au deposits in the Mount Lyell field.
Mapping of mineral zonation is effective where large
systematic datasets are available (i.e. where there are plenty of
outcrops or drill cores) and mineral assemblages are visually
distinctive or can be determined by simple field tests (e.g.
effervescence in acid for carbonate or sodium-cobaltinitrite
staining for K-feldspar). However some mineral assemblages
that are not readily identified visually, are discretely zoned
and may be diagnostic of a deposit style. New field-based
mineralogical tools, such as portable SWTR spectrometers
(Section 2.4), will facilitate major improvements, which will
not only aid exploration, but also contribute to understanding
these deposit systems (Thompson et al., 1999). SWIR spectral
studies have recently shown some spectacular examples of
mineral zonation, particularly in acid-sulfate type systems
(e.g. case studies in Thompson et al., 1999, and Huston and
Kamprad, 2001).
Major element lithogeochemistry
Although intense hydrothermal alteration frequently
produces simple alteration mineral assemblages, the minerals
are commonly fine grained. These minerals may be difficult to
recognise visually, and it can also be difficult to estimate their
abundances. In these cases, lithogeochemistry can frequently
help to identify minerals and quantify compositional changes
even in less intensely altered rocks that contain incipient,
overprinting or domainal alteration minerals. Analysis of
whole-rock samples to determine major element abundance
is a way of supporting and augmenting estimates of mineral
proportions and alteration intensity, which have been
determined visually or by other methods (e.g. Section 2.4).
Quantitative lithogeochemical data can be used in two
ways: (1) to indicate alteration intensity, and (2) to estimate
mineral proportions in mineral assemblages where the mineral
species and their individual compositions are known.
Interpreting exploration vectors based on compositional
data seems straightforward. The data can be plotted as contour
maps or cross-sections (e.g. Figs 2.7 and 2.12), down-hole
line graphs (e.g. Fig. 2.14) and so on, and the vectors inferred
based on expected variations in mineral abundance or
composition. Decreases in Na
2
O contents of volcanic rocks,
for instance, are usually related to increasing sericite or chlorite
at the expense of plagioclase. Na
2
O depletion is a popular and
reliable vector used in VHMS exploration (e.g. Na
2
O halo
maps of the Fukuzawa area in Date et al., 1983). Variations in
carbonate content, both increases and decreases, are typically
evident in CO
2
data and in CaO, MgO or Fe
2
O
3
, depending
on the carbonate species. Sulfide content can be quantified
by sulfur analyses. Weight percent sulfur is generally nearly
equivalent to volume percent of pyrite in felsic rocks, if pyrite
is the only sulfur-bearing phase. This is due to pyrite having
a density of just under twice the density of felsic rock, and
sulfur comprising just over half the mass in pyrite.
However, major element data are subject to distortion
by closure, otherwise known as the constant sum effect.
This phenomenon is more fully explained in Section 4.1.
It particularly affects the dominant chemical components
(e.g. SiO
2
, A1
2
O
3
and Fe
2
O
3
) and can be significant in
2 4 4 | CHAPTER!
hydrothermally altered rocks with large net mass gains.
Although additions of exotic hydrothermal components such
as sulfur and CO
2
, and depletions of Na
2
O, are relatively
immune to the effects of closure, it seriously compromises
the use of some other major components as exploration
vectors. For example, SiO
2
may not provide effective vectors
in hydrothermal systems where mineralisation was associated
with silicification. If closure in major element data is likely
to obscure the effects of alteration on the compositional data
and exploration vectors, then it is preferable to estimate the
individual component mass changes (by immobile element
techniques, Section 4.1) and use those as exploration vectors.
The alternative approach is to convert quantitative major
element lithogeochemical data to modal mineral proportions
using a method such as MINSQ (Herrmann and Berry,
2002) or GENMIX (Le Maitre, 1981). This does not remove
the effects of closure, but is a way of quantifying mineral
proportions, which can then be used as vectors in mineral
exploration. This approach was used by Large et al. (2001b,
Fig. 6) to demonstrate systematic variations in proportions
of alteration minerals around the Rosebery K-lens sulfide
deposit (Fig. 2.14).
Alteration indices
Alteration indices formulated from two or more components
of major element analyses (Section 2.4) enhance the
compositional contrast between variably altered samples and
thus are frequently more effective as exploration vectors than
single component lithogeochemical data.
For example, sulfur and Na
2
O proportions in the footwall
of the Rosebery K-lens deposit range from 0.01% (limit of
detection) up to about 7.2% and 5.6%, respectively (Large
and Allen, 1997). However, the ratio S/Na
2
O ranges from
0.002 to 194, because those components increase and decrease
respectively in response to increasing alteration intensity
(Large et al., 2001b). Both components vary over two to three
orders of magnitude, whereas S/Na
2
O varies across about five
orders of magnitude. Carefully formulated indices can in this
way amplify compositional changes and reflect variations in
more than one mineral composition or abundance.
Where systematic lithogeochemical data are available,
plotting and contouring of alteration indices on plans and cross-
sections provides numerical indications of alteration intensity
(e.g. Fig. 2.7). Datasets of alteration indices are of assistance
in guiding exploration towards potentially mineralised altered
zones, especially when used in combination with alteration
facies or mineral zonation maps. Mineral explorers have
increasingly applied these techniques to VHMS exploration
over the last two decades; however, few results or case studies
have been published.
Exploration data are commonly limited to a few
samples or drill holes and are not suitable for contouring.
Nevertheless, useful vectors can be inferred from sparse but
strategically or fortuitously located data. This is exemplified in
lithogeochemical data from a few drill holes near the northern
end of the Rosebery deposit (65R, 109R, 113R and 128R;
Table 8.1). If, in a VHMS exploration scenario, the first two
holes were drilled in sequence (65R followed by 109R), then
the lithogeochemical vectors would suggest that exploration
was heading away from the most favourable zone. The
intermediate third hole, 113R, would then be superfluous,
merely reinforcing interpretation of vectors in the first two
holes. The anomalous values in the near-miss hole (65R)
would encourage further persistence. If, on the other hand,
the first hole in a greenfields exploration program was 109R,
the major element or alteration indices data would not justify
continuing exploration in that vicinity, even if the favourable
stratigraphic setting was recognised. In this case, success
would depend on the explorer recognising other vectors or
indicators of proximity, such as the distal trace element Tl and
Sb halos identified by Large et al., (2001b).
In favourable geologic settings, limited lithogeochemical
data, even from a single drill hole, may yield useful vectors.
For example, samples from a single hole, such as HL6 or
TABLE 8.1 | Selected major element data and alteration indices for samples of pumice breccia from the footwall to the Rosebery K-lens
massive sulfide deposit, western Tasmania. The values tabulated are (A) averages of three samples from the top 30 m of the footwall unit
and (B) the uppermost sample of the footwall intersected in eachdriii hole. The alteration indices, S/Na
2
O and Al, generally show greater
increases with proximity to ore than the changes in Na
2
O, S and Zn. Averaging the uppermost three samples smoothes the gradients
towards ore, but diminishes the anomalies in the medial intersection, 113R. Data from Large and Allen (1997). Locations of the drill holes
are shown in Figure 2.7 of this volume and Figure 2 of Large et al. (2001b).
(A) Averages of three samples from the top 30 m of the footwall unit
109R 450 2.99 0.18 0.00 0 44
113R 250 1.66 0.37 0.13 2 61
65R 75 0.28 0.79 0.18 44 89
128R 0 0.01 0.49 0.49 49 89
(B) Uppermost sample of footwall unit
109R 450 1.53 0.29 0.01 0 51
113R 250 0.21 1.09 0.38 5 89
65R 75 0.08 1.02 0.02 13 95
128R 0 0.01 0.37 0.19 37 89
HL14 drilled through the footwall zones beneath the Hellyer
massive sulfide deposit, generally exhibit gradually increasing
alteration indices (Figs 9 and 20 of Gemmell and Large,
1992). Recognition of these variations, in combination with
the visible alteration facies, confirms that an altered zone exists
and indicates the direction of increasing alteration intensity,
guiding further exploration.
Drill hole NC4, which intersected the Tyndall Group-
Central Volcanic Complex contact south of Henty is another
interesting example. In this hole, an abrupt down-hole increase
in the alteration indices is associated with a change of lithotype
(Fig. 4.4). The lithogeochemical data support the recognition
of an extensive stratabound altered zone of which the upper
boundary is most favourable for VHMS exploration.
Bivariate (x-y) plots of two alteration indices, such as the
AI-CCPI Alteration box plot (Large et al., 2001a), are useful
in identifying compositional trends and different alteration
facies. This graphic approach simplifies the recognition of rock
compositions that lie outside the normal range of primary
volcanic compositions (i.e. those that have been modified by
chemical or depositional processes; Fig. 2.9). It also assists
classification of different alteration facies and identifying the
zones of greatest prospectivity (Fig. 2.11).
In recent CODES research projects, box plots of custom-
designed alteration indices have been effective in several
other types of hydrothermal systems, including low- and
high-sulfidation epithermal Au-Ag deposits (Williams, 2000)
and Broken Hill type Pb-Zn-Ag deposits (Large, 2004).
The Ishikawa et al. (1976) alteration index (AI) has been
successfully applied to many plagioclase-destructive and/
or K-feldspar-bearing alteration styles, but there is scope
for more experimentation with new indices. As outlined
in Section 4.1, the formulae for alteration indices typically
have chemical components that were increased by alteration
in their numerators, and components that were decreased
in the denominators. The gained or lost components can
often be inferred from the differences in alteration mineral
assemblages. However, immobile-element-based mass change
calculations provide a more rigorous method of selecting
components for formulating alteration indices. Section 4.1
summarises several techniques of estimating mass changes
by comparing compositions of alteration facies to their least-
altered precursor compositions and their potential application
to exploration vectors is discussed below.
Mass change v e c t o r s
Hydrothermal alteration commonly involves major changes
in chemical composition; in fact these changes are one of
the characteristic features of hydrothermally altered rocks.
Significant masses of mobile chemical components may have
been gained or removed from an altered zone. However,
closure in composition data will obscure or distort the
amounts of these changes, except in the special cases where
the mass gains and loses balance exactly, so that there is no net
mass change. It is unsound to assume zero net mass change in
an alteration facies, and in these cases the unquantified effect
of closure on raw major element data limits their usefulness,
or that of alteration indices based on them, as indicators of
alteration intensity.
FINDING ORE DEPOSITS IN ALTERED VOLCANIC ROCKS | 2 4 5
The solution to the closure problem is to estimate the
mass changes of all mobile major-element components, using
an immobile-element-based method of the type described in
Section 4.1. Spatially located mass change data can then be
used as direct indicators of alteration intensity or as multiple
component alteration indices, in the same way as major
element lithogeochemical data.
This approach has the potential to target favourable areas
during exploration. It provides closure-free quantification of
compositional changes, which help delineate hydrothermal
fluid pathways, zones of greatest alteration intensity and
prospective areas. Furthermore, the quantification of absolute
mass changes is a means of estimating the 'quality' of an
altered zone.
For example, let us consider a hypothetical program of
lithogeochemical sampling over two altered zones of similar
dimensions in a VHMS district. Mass change estimates might
show that the first altered zone involved negligible mass
transfers and the second had significant mass gains, of the
order of 2030 g/lOOg and equating to tens of millions of
tonnes of altered rock (cf. Thalanga footwall zone, Herrmann
and Hill, 2001). In this case, we would conclude that the
second altered zone has greater mineral potential. Substantial
mass changes demonstrate that a hydrothermal system had
the intensity, and perhaps duration, to move large amounts
of SiO
2
, CO
2
, S and other components into the alteration
facies. Therefore, it probably also had the capacity, if fluid
compositions were suitable, to move large amounts of base
and precious metals and potentially, if a favourable site and
process for deposition is available, form an ore deposit. The
first altered zone in our hypothetical example was produced by
near-isochemical alteration and resulted in negligible changes
to the whole-rock composition, suggesting that alteration
involved less reactive or smaller volumes of fluid, perhaps
over a short duration. The differences may be semi-evident in
alteration mineral assemblages and intensities, and possibly
in the composition data despite distortion by closure, but the
only way to quantify the difference for objective exploration
decisions is by mass transfer techniques.
The major difficulty in this method is in determining
precursor compositions to compare with the altered compo-
sitions. Poor exposures, limited lithogeochemical data, lateral
variation in the primary composition of volcanic facies or
structural complexity make the pairing of alteration facies
and unaltered (or least-altered) precursors problematic, and
frequently impossible, in practical application. There are no
published examples where mass change calculations have
led to a mineral discovery, probably because of the least-
altered precursor problem and the only recent development
of easy mass change calculation techniques. Nevertheless,
the mass change approach will contribute to a higher
level of lithogeochemical interpretation and exploration
targeting where host volcanic successions are compositionally
uniform and sufficiently understood to enable its confident
application.
Mineral chemistry vectors
As noted in Section 4.2, the main limitations to the wide use of
mineral chemistry in exploration have been that the analytical
2 4 6 | CHAPTER 8
tools electron microprobe and X-ray diffraction are
complex laboratory-based instruments requiring considerable
expertise in operation and data interpretation. That has made
mineral analysis slow and expensive relative to geochemical
analyses of rocks, soils and sediments, and consequently
explorers have largely ignored the mineral chemistry vector
possibilities. Researchers at CODES are currently developing
laser ablation ICP-MS techniques for micro-analysis of trace
elements in sulfides. These are likely to provide exploration
vectors but, for similar reasons, they may not ultimately be
widely applied by mineral explorers.
However, the advent during the last decade of portable
short wavelength infrared (SWIR) spectrometers, which
can indirectly measure compositional variations in micas,
clays and carbonates, could establish mineral composition
mapping as a viable exploration technique (Sections 3.1
and 4.2, and references therein). SWIR spectrometers such
as PIMA are relatively inexpensive at about US$21,000 to
purchase or US$70 per day for hire. They can analyse up
to a few hundred samples per day of all types of geological
materials, which require no preparation apart from drying.
SWIR spectrometers are simple to operate and the PC-based
spectral recognition software now available has simplified
spectral interpretation and data manipulation, so that an
operator can quickly become an expert interpreter.
White micas, chlorites and clays in altered zones around
mineral deposits frequently show spatial compositional varia-
tions that could be exploration vectors (Section 4.2). The ease
of SWIR spectral analysis now enables explorers to rapidly
test for the existence of mineral composition vectors in a large
enough set of orientation samples. If the results are promising,
the technique can be inexpensively applied on a routine basis
to assist exploration targeting. If, on the other hand, SWIR
spectral features are invariant or spatially erratic, then little
time and money will have been expended.
There are not yet many published mineral exploration
case studies involving portable SWIR spectral analysis because
it is a relatively new technique (e.g. Denniss et al., 1999;
Huston et al., 1999; Merry and Pontual, 1999; Herrmann
et al., 2001; Jones et al., in prep.). Nevertheless, recent and
current research at CODES shows great potential for SWIR-
determined white mica composition vectors, on scales of tens
to hundreds of metres, in a variety of volcanic-hosted gold
and base-metal deposits. Further work is required on spatial
SWIR spectral variations in chlorites and clay minerals. It is
likely that mineral explorers will rapidly adopt this technique
over the next few years.
Part of the stimulus comes from very recent developments
in airborne high-resolution visible-to-SWIR spectral scanning
systems, such as HyMap", which offer great promise for district-
scale mineral mapping in exploration of well-exposed bedrock
areas (Taranik, 2001). For example, mineral maps from a trial
HyMap* airborne spectral survey of the Panorama VHMS
district, Western Australia, apparently 'show the complete
hydro thermal convective system' (Cudahy et al., 2000). At
Panorama, these authors consider that spectrally interpreted
distributions of white mica, pyrophyllite and topaz define
altered zones that formed at the boundary between magmatic
fluid and seawater convection, in addition to seawater recharge
zones, and hydrothermal discharge zones. The discharge zones
are prospective for massive sulfides. A similar HyMap* survey
of the Mount Lyell area in western Tasmania has produced
mineral distribution and pyrophyllite abundance maps (e.g.
Fig. 8.1). These illustrate the high spatial resolution now
available from airborne spectral surveys, and their enormous
potential for alteration mapping and using vectors during
exploration in well-exposed, thinly vegetated areas.
Remote sensing spectral systems are also finding
applications in regolith mapping (Craig, 2001) and exploration
of partly covered areas. Bierwirth et al. (2002) used HyMap
data to map distributions of a range of minerals including
pyrophyllite, white mica, Mg- and Fe-chlorite, calcite,
dolomite, kaolinite, tourmaline, hematite and goethite in
altered zones associated with epithermal and lode Au deposits
in the poorly exposed, largely alluvium- and calcrete-covered
Indee District of the Central Pilbara.
These demonstrations of district-scale mineral and
mineral compositional mapping by remote sensing tools
should certainly encourage explorers to use spectral data in
prospect-scale investigations. In addition, high-output, multi-
purpose visible-SWIR and thermal infrared spectral, and
laser instruments such as CSIRO's HyLogging and HyChips
systems (Syddell, 2004) and the OARS prototype (CSIRO,
2002), are being developed for routine logging of drill core,
cuttings, soil and other geological sample materials.
Isotopic vectors
Section 4.3 introduces the potential for isotope geochemistry
to yield interpretations of hydrothermal fluid sources,
temperatures, water-rock ratios, and broad halos for
exploration targeting.
Oxygen isotopes are particularly useful in exploration
because oxygen is a major component of hydrothermal fluids,
and it readily exchanges isotopes with silicate minerals at
fractionation factors that are mineral specific and temperature
dependent. Furthermore, the 5
18
O-depletion halos observed
around several deposit types typically extend further from
ore than most other geochemical anomalies and may provide
direct vectors to ore zones. For example, the 5
18
O-depletion
zone around the Fukuzawa deposits in the Hokuroku district,
Japan, extends up to 1 km beyond the Na
2
O-depletion
anomaly (Green et al., 1983). Waring et al. (1998) found
6
18
O-depletion anomalies in dolomitic shale at Mount Isa
(Queensland), which extend up to 2 km beyond Cu ore zones,
with low and uniform isotopic gradients (<2%o per 100 m)
that allow estimates of the distance to ore. Most importantly,
the O-isotopic anomalies produced in hydrothermally
altered zones appear to survive subsequent deformation
and metamorphism. For instance, Cartwright (1999)
argued convincingly that a hydrothermally related regional-
scale 6
18
O depletion zone in Proterozoic metapelites in the
Broken Hill district, NSW, had survived high-grade regional
metamorphism up to granulite facies. The final section of this
chapter summarises several VHMS-related alteration studies
and exploration programs, which have applied whole-rock O-
isotope geochemistry.
Sulfur-isotope geochemistry has been widely applied
to interpretations of sulfur (and hence fluid) sources, and
hydrothermal temperatures, which have been used in the
development of VHMS genetic models. For example,
FINDING ORE DEPOSITS IN ALTERED VOLCANIC ROCKS I 2 4 7
Map projection: UTM zone 55, AGD6 6
FIGURE 8.1 | Mineral maps of the Mount Lyell mine area, western Tasmania, interpreted from HyMap airborne hyperspectrai data. Map A shows the zonal
distributions of eight important alteration minerals. Map B shows relative abundance of pyrophyllite (warm colours = high abundance), and discriminates the
pyrophyllite-rich facies at North Lyell, Western Tharsis and Glen Lyell from weaker responses in the Owen Group exposed on Mount Lyell. The spatial resolution (pixel
size) is about 5 m. Mineral spectral responses are partly restricted by vegetated areas, which appear as dark grey tones on the HyMap band (greyscale) background
airphoto images. These maps were created by K. Yang, M.A. Quigley and J.F. Huntington as part of the 2003 HyMap mineral mapping project for Copper Mines of
Tasmania and Mineral Resources Tasmania, carried out through the C-Vista strategic alliance between CSIRO and HyVista Corporation.
2 4 8 | CHAPTER 8
S-isotope compositions constrained some of the genetic
interpretations for formation of the Hellyer deposit (e.g.
Gemmell and Large, 1992; Solomon and Khin Zaw, 1997).
It also has exploration potential for discriminating different
types of deposits and hydrothermally altered zones, which
may have different economic potential. The regional study
of sulfide deposits in the Mount Read province by Solomon
et al. (1988) found considerable variation in S-isotope
compositions consistent with different geologic settings and
metal associations, and which contributed to interpretations
of hydro thermal geochemistry. Green and Taheri (1992)
took both a genetic and discriminatory approach to the
interpretation of low 6
34
S values of pyrite (-1.2 to +4.7%o)
at the Boco prospect, western Tasmania. They suggested that
altered zones at Boco formed in a seawater-hydrothermal
system, which leached sulfur from volcanic host rocks at
temperatures that were too low to inorganically reduce
seawater sulfate, and transported base metals to form an ore
deposit. Subsequent recognition of advanced argillic alteration
mineral assemblages at Boco and several other Tasmanian and
Victorian prospects indicate possible involvement of magmatic
fluids, and hence a magmatic source of sulfur (Herrmann et
al., 2004). Regardless of the genetic uncertainties, sulfide
6
34
S values of less than 5%o could distinguish barren pyritic
altered zones from more prospective base and precious metal-
rich VHMS systems in the Mount Read province.
There are few published accounts of S-isotope compositions
as direct vectors in mineral exploration. However, existing data
for Rosebery (Davidson et al., 2000) and Hellyer (Gemmell
and Large, 1992) suggest broad halos of 6
34
S enrichment in
disseminated pyrite in footwall zones lateral to the main up-
flow zones, which could be used to increase exploration target
sizes and zero-in on Zn-rich VHMS deposits, particularly in
permeable volcaniclastic successions (Large et al., 2001c).
At regional scales, two recent studies of deeply covered
areas have promoted S-isotope compositions of sulfates in
groundwater as potential indicators of buried oxidising Pb-
Zn-Ag sulfide deposits, in the Broken Hill region of New
South Wales (Waring et al., 1998) and Gawler Craton in
South Australia (Kirste et al., 2003). In the latter case, sulfates
from oxidising sulfide deposits with low 5
34
S signatures (-2.5
to +5.6%o) appear to have contributed to anomalous low 5
34
S
values in groundwater sulfates, detectable several hundred
metres downstream from the Menninnie Dam prospect.
Background 6
34
S values of sulfates in ground waters are 16 to
18%o in the Gawler region and -13.5%o in the Broken Hill
region. The concept is probably less applicable to exploration
for sulfide deposits with higher 6
34
S signatures (e.g. Tasmanian
VHMS deposits, 8 to 17%o, Solomon et al., 1988), which
would provide less contrast against background groundwater
compositions. Furthermore there are many, typically difficult
to determine, hydrological and geochemical factors that
complicate interpretations of local groundwater isotopic
anomalies. This new application of S-isotope geochemistry is
one that will probably appeal only to the most persistent of
under-cover mineral explorers.
Carbon isotopes, like sulfur, are used for interpreting
fluid sources and hydrothermal conditions but have not
been widely applied as exploration vectors. Huston's (1999)
review of stable isotopes in VHMS systems found carbonate
6
13
C values in most deposits occupy a narrow range of-5 to
0%o, consistent with seawater dissolved bicarbonate sources.
Low fractionation factors, and the limited occurrences
of carbonates in massive sulfide deposits (Ohmoto and
Goldhaber, 1997) restrict the applications of C isotopes,
except in conjunction with O isotopes. For example, Khin
Zaw and Large (1992) interpreted a coupled positive trend
of 6
13
C and 6
18
O data in Mn-rich carbonates at South
Hercules, Tasmania, as temperature-related, and then, with
additional fluid inclusion temperature data, estimated the
isotopic compositions of the hydrothermal fluid. Although
their paper did not describe spatial zonation of isotopic data,
the genetic discussion speculated that mineralised and altered
facies were zoned according to variations in temperature and
hydrothermal fluid-seawater mixing ratios, controlled by
permeabilities in the volcaniclastic succession. In these types
of deposits associated with lateral carbonate facies, isotopic
data could provide prospect-scale exploration vectors if the
hydrothermal temperature gradients were consistent.
Callaghan's (2001) study of the Henty-Mount Julia gold
deposit, Tasmania, used carbonate 6
13
C and 6
18
O data in
a boomerang shaped trend for intensive modelling of fluid
compositions and genetic concepts. The data, crudely divided
into proximal and distal carbonates, lie on two trends joined
at an abrupt inflection. Both of the fluid mixing or fluid-
limestone interaction models proposed by Callaghan (2001)
to account for the trends offer potential for prospect to district-
scale isotopic vectors, or at least methods of discriminating
hydrothermal and sedimentary carbonates.
Whole-rock O-isotope vectors in VHMS exploration
In most cases, the proximal altered zones of VHMS systems
show significant 6
18
O depletion, partly attributable to high-
fluid temperatures and low-fractionation factors of some
minerals (e.g. chlorite) in seafloor hydrothermal discharge
zones, and partly due to the contrast with 6
18
O enrichment
caused by low-temperature seawater-rock reactions in normal
submarine volcanic successions.
A classic semi-regional study by Cathles (1993) in the
Noranda district, Canada, discovered a low 6
18
O anomaly
(<6%o) in volcanic rocks around the Flavrian felsic pluton. The
pluton is surrounded by a discontinuous halo of high whole-
rock 6
18
O anomalies (>9%o) 10-15 km from the intrusion.
Several narrow finger-like zones of low 6
18
O values extend
radially from the inner
18
O-depleted zone through the high
6
18
O halo, in the directions of most of the known VHMS
deposits in the district (Fig. 8.2). These low 5
18
O zones record
areas of high hydrothermal-fluid flow and concentrated
discharge, which are favourable for mineral deposits. The
concentric zones of
18
O depletion and enrichment around
the pluton closely match the isotopic zonation patterns of
Cathles' (1983) numerical model. He concluded that whole-
rock 8
18
O sampling, at 0.5 km intervals along traverses
adjacent to the margins of plutons, could identify plutons
with sufficient energy to drive long-lived hydrothermal
systems, and favourable settings for detailed massive sulfide
exploration.
Another district-scale study, in the Panorama area of
Western Australia, showed a similar pattern of low whole-rock
6
18
O around the perimeter of a large subvolcanic intrusion
FINDING ORE DEPOSITS IN ALTERED VOLCANIC ROCKS | 2 4 9
FIGURE 8.2 | Map showing spatial
relationships between felsic plutons, whole-rock
6
18
O anomalies, and massive sulfide deposits in
the Noranda area, Abitibi belt, Canada (modified
after Cathles, 1993).
(Brauhart et al., 2000). The granitoid pluton underlies a
1.5 km thick mixed mafic to felsic volcanic succession that
hosts several small polymetallic massive sulfide deposits
and prospects along a single favourable horizon at the
stratigraphic top of the sequence (Fig. 6.6). Narrow radial
zones of low 6
18
O point to most of the known deposits and
prospects (Fig. 8.3). These low 6
18
O zones coincide with
intense feldspar-destructive sericite + quartz and chlorite +
quartz zones. Brauhart et al. (2000) calculated hydrothermal
temperatures from the 6
18
O data. They used fractionation
factors calculated to suit the specific modal mineralogy of
each sample, an initial fluid 6
18
O value of+2%o and assumed
high water-rock ratios. The resulting calculated temperature
distribution closely matched the O-isotopic pattern, the low
6
18
O zones coinciding with temperatures greater than about
300C (Fig. 8.3). This indicates that temperature was the
main control on low whole-rock 8
18
O. It is consistent with
increased temperature with depth in the volcanic succession,
and in the transgressive discharge or feeder zones beneath
the sulfide deposits. The authors concluded that whole-rock
O-isotope mapping could be used as a regional exploration
vector, and pointed to additional favourable targets in the
Panorama district.
Green and Taheri (1992) followed up the Hokuroku
work of Green et al. (1983) with several empirical isotopic
studies of alteration systems in the Mount Read province. The
altered footwall zones beneath the Hellyer deposit exhibit a
subtle whole-rock 6
18
O anomaly with values ranging from
8.3 1.3%o in the central stringer zone, through 9.8 1.7%o
in the enclosing sericitic zone, to background values around
11.30.9%o in adjacent least-altered footwall andesites.
There is also a subtle depletion anomaly of 10.6 1.2%o
in the basalts immediately above the deposit, compared to
background values of 11.82.2%o. However, the 5
18
O-
depletion zone is narrow, reflecting the strong fault or fracture
control on hydrothermal-fluid flow. This limits its utility
in exploration. The Hercules alteration system also shows
a range of whole-rock 6
!8
O values from 6.8%o in footwall
zones to background values of 14.0 to 15.5%o. There are
some unexpectedly high values (around 15%o) in relatively
proximal parts of the footwall and lowvalues (down to 6.8%o)
FIGURE 8.3 | Distribution of whole-rock 6
18
O values and estimated
hydrothermal temperatures in the Panorama district of the Pilbra region, Western
Australia (modified after Brauhart et al., 2000).
2 5 0 | CHAPTER!
in the apparent hanging wall rocks to the east. These may
be partly due to fault displacements that dismembered the
alteration system. They highlight the difficulty of applying
broad-scale geochemical exploration techniques in deformed
and structurally complex terrains.
In contrast to Hellyer and Hercules, the apparently barren
Boco altered zone has whole-rock 6
18
O values that are not
significantly different to background values (9.9 1.0 and
10.5 l.l%o, respectively). This is consistent with Green and
Taheri's (1992) interpretation that the Boco alteration facies
formed in a low-temperature (<200
c
C) seawater hydrothermal
system, incapable of transporting base metals and reducing
seawater sulfate. Alternatively, the 6
18
O values could indicate
a higher temperature, isotopically heavier fluid (>280C,
--5%o), representing either evolved seawater or mixed seawater
and magmatic water. The presence of advanced argillic
assemblages in parts of the Boco system implies highly acidic
fluid conditions, which supports a magmatic fluid input
(Herrmann et al., 2004).
The least-altered volcanic rocks in VHMS-hosting
successions typically have anomalously high background
whole-rock 6
18
O values (>9 or 10%o), which are attributable
to low-temperature diagenetic alteration. The curves in
Figure 4.17 indicate that re-equilibration with quite small
proportions of cold seawater can produce large positive shifts
in volcanic rock 6
18
O values. On the other hand, zones of
low 6
18
O reflect high-temperature hydrothermal alteration
at high water-rock ratios. The empirical data from Noranda
and Panorama show that low 8
18
O zones may be regionally
extensive at depths of greater than 1 km below favourable
horizons, and in narrow finger-like zones that point toward
favourable sites for hydrothermal discharge. They may form
relatively broad halos around massive sulfide deposits.
Despite these promising research results, VHMS
explorers have been less than enthusiastic about O-isotope
vectors and there are few examples of successful application
in Australia. This may be largely attributable to the expense
of isotopic analysis (currently around US$150 per sample)
and the recognition that interpretation of isotopic data is not
straightforward.
A notable exception is the case of the Thalanga West 45
deposit, documented by Miller et al. (2001). These authors took
a similar approach to Brauhart et al. (2000), using estimates
of modal mineralogy to determine tailor-made fractionation
factors for each sample, to calculate isotopic equilibration
temperatures from whole-rock 6
18
O data, at assumed high
water-rock ratios and fluid isotopic composition. They found
that zones of apparent high temperatures (>230C) coincided
with the known Central, East and Orient massive sulfide lenses.
The existence of an additional isotopic-temperature anomaly,
about 1 km west of the known resources, stimulated further
exploration that turned up a favourable REE geochemical
anomaly in the same sector. Subsequent exploratory drilling
discovered a 0.23 Mt polymetallic massive sulfide lens. It
remains sub-economic, but may represent the first successful
VHMS exploration application of O-isotope geochemistry in
Australia.
I REFERENCES
251
Abbott, P. D. B., 1992. Geology of a barite-galena occurrence
exposed in the Anthony Power Development Tunnel,
western Tasmania: Unpublished Honours thesis, University
of Tasmania, 62 p.
Ade-Hall, J. ML, Khan, A., Dagley, P., and Wilson, R. L., 1968.
A detailed opaque petrological and magnetic investigation
of a single Tertiary lava flow from Skye, Scotland - I iron-
titanium oxide petrology: Geophysical Journal of the Royal
Astronomical Society, v. 16, p. 37588.
Alabaster, T, and Pearce, J. A., 1985. The inter-relationship between
magmatic and ore-forming hydrothermal processes in the
Oman Ophiolite: Economic Geology, v. 80, p. 1-16.
Allen, R. L., 1988. False pyroclastic textures in altered silicic lavas,
with implications for volcanic-associated mineralisation:
Economic Geology, v. 83, p. 142446.
Allen, R. L., 1990. Subaqueous welding, or alteration, diagenetic
compaction and tectonic dissolution? [abs.]: IAVCEI
International Volcanological Congress, Mainz, Germany,
September 3-8, 1990, Abstracts volume, p. 2.
Allen, R. L., 1991. Structure, stratigraphy and volcanology of the
Rosebery-Hercules Zn-Pb-Cu-Au massive sulfide district,
Tasmania: Results 1988-1990: Unpublished Report to
Pasminco Exploration and Mining, 3 volumes, 54 p.
Allen, R. L., 1994a. Volcanic facies analysis indicates large pyroclastic
eruptions, sill complexes, synvolcanic grabens and subtle
thrusts in the Cambrian 'Central Volcanic Complex' volcanic
centre, western Tasmania [abs.], in Cooke, D. R., and
Kitto, P. A., eds., Contentious issues in Tasmanian geology,
a symposium: Hobart, Australia, November 3-4, 1994,
Geological Society of Australia Abstracts, v. 39, p. 413.
Allen, R. L., 1994b. Synvolcanic, subseafloor replacement model for
Rosebery and other massive sulfide ores [abs.], in Cooke, D.
R., and Kitto, P. A., eds., Contentious issues in Tasmanian
geology, a symposium: Hobart, Australia, November 34,
1994, Geological Society of Australia Abstracts, v. 39,
p. 107-8.
Allen, R. L., 1997. Rosebery alteration study and regional alteration
studies in the Mount Read Volcanics - the record of diagenetic
alteration in the strongly deformed, felsic volcaniclastic
succession enclosing the Rosebery and Hercules massive
sulfide deposits: AMIRA/ARC Project P439, Centre for Ore
Deposit and Exploration Studies, University of Tasmania,
Unpublished Report 5, October 1997, p. 135-73.
Allen, R. L., and Cas, R. A. E, 1990. The Rosebery controversy:
distinguishing prospective submarine ignimbrite-like units
from true subaerial ignimbrites in the Rosebery-Hercules
Zn-Cu-Pb massive sulfide district, Tasmania [abs.]: 10
th
Australian Geological Convention, Gondwana: terranes and
resources, Hobart, Australia, February 4-9, 1990, Geological
Society of Australia Abstracts, v. 25, p. 31-2.
Allen, R. L., and Large, R., 1996. Rosebery alteration study:
AMIRA/ARC Project P439, Centre for Ore Deposit and
Exploration Studies, University of Tasmania, Unpublished
Report 3, October 1996, p. 143-52.
Alt, J., C, 1995a. Sulfur isotopic profile through the oceanic crust -
sulfur mobility and seawater-crustal sulfur exchange during
hydrothermal alteration: Geology, v. 23, p. 58588.
Alt, J. C, 1995b. Subseafloor processes in mid-ocean ridge
hydrothermal systems: AGU Geophysical Monograph,
v. 91, p. 85-114.
Alt, J. C, 1999. Hydrothermal alteration and mineralisation of
oceanic crust - mineralogy, geochemistry and processes:
Reviews in Economic Geology, v. 8, p. 133-55.
Alt, J. C, and Honnorez, J., 1984. Alteration of the upper oceanic
crust, Deep Sea Drilling Project site 417 mineralogy and
chemistry: Contributions to Mineralogy and Petrology,
v. 87, p. 149-69.
Alt, J. C, Honnorez, J., Laverne, C, and Emmermann, R., 1986.
Hydrothermal alteration of a 1-km section through the
upper oceanic crust, Deep Sea Drilling Project hole 504B
- mineralogy, chemistry, and evolution of seawater-basalt
interactions: Journal of Geophysical Research, v. 91,
p.10309-35.
Alt-Epping, P., and Smith, L., 1997. Geochemical patterns associated
with the onset of convective circulation at a sedimented
ridge crest [abs.]: Geological Society of America Abstracts
with Programs, v. 29, p. 74.
Amstutz, G. C, and Park, W. C, 1967. Stylolites of diagenetic age
and their role in the interpretation of the southern Illinois
fluorspar deposits: Mineralium Deposita, v. 2, p. 4453.
Anderson, J. E. J., 1969. Development of snowflake texture in a
welded tuff, Davis Mountain, Texas: Geological Society of
America Bulletin, v. 80, p. 2075-80.
Anonymous, 1997. Mineral analysis guide, ANALABS, Bentley,
Western Australia, 41 p.
Ashley, P. M., Dudley, R. J., Lesh, R. H., Marr, J. M., and Ryall,
A. W, 1988. The Scuddles Cu-Zn prospect, an Archaean
volcanogenic massive sulfide deposit, Golden Grove District,
Western Australia: Economic Geology, v. 83, p. 91851.
Bailes, A. H., and Galley, A. G., 1999. Evolution of the
Palaeoproterozoic Snow Lake arc assemblage and geodynamic
setting for associated volcanic-hosted massive sulfide deposits,
Flin Flon Belt, Manitoba, Canada: Canadian Journal of
Earth Sciences, v. 36, p. 1789-805.
2 5 2 | REFERENCES
Baillie, P. W., 1989. Stratigraphy, sedimentology and structural
setting of the Cambrian Sticht Range Formation, western
Tasmania: Tasmania Geological Survey Bulletin, v. 65, p. 1
34.
Baker, J. H., 1985. The petrology and geochemistry of 1.81.9 Ga
granitic magmatism and related subseafloor hydrothermal
alteration and ore-forming processes, W. Bergslagen, Sweden:
GUA Papers of Geology, v. 1, 203 p.
Banwell, C. J., Cooper, E. R., Thompson, G. E. K., and McCree,
K. J., 1957. Physics of the New Zealand thermal area:
Department of Scientific and Industrial Research Report,
no. 123, 109 p.
Baragar, W. R. A., Plant, A. G., Pringle, G. J., and Schau, M., 1977.
Petrology and alteration of selected units of Mid-Atlantic
Ridge basalts sampled from sites 332 and 335, Deep Sea
Drilling Project: Canadian Journal of Earth Sciences, v. 14,
p. 837-74.
Baragar, W. R. A., Plant, A. G., Pringle, G. J., and Schau, M., 1979.
Diagenetic and post-diagenetic changes in the composition
of an Archaean pillow: Canadian Journal of Earth Sciences,
v. 16, p. 2102-21.
Barker, A. J., 1990. Introduction to metamorphic textures and
microstructures: Glasgow, Blakie and Sons Ltd, 162 p.
Barnes, H. L., ed., 1979. Geochemistry of hydrothermal ore deposits,
3
rd
edition: New York, USA, John Wiley and Sons, 798 p.
Barrett, T. J., and MacLean, W. H., 1991. Chemical, mass, and
oxygen isotope changes during extreme hydrothermal
alteration of an Archaean rhyolite, Noranda, Quebec:
Economic Geology, v. 86, p. 406-14.
Barrett, T. J., and MacLean, W. H., 1994a. Chemostratigraphy and
hydrothermal alteration in exploration for VHMS deposits
in greenstones and younger volcanic rocks, in Lentz, D. R.,
ed., Alteration and alteration processes associated with ore
forming systems: Geological Society of Canada Short Course
Notes, v. 11, p. 433-67.
Barrett, T. J., and MacLean, W. H., 1994b. Mass changes in
hydrothermal alteration zones associated with VMS deposits
of the Noranda area: Exploration and Mining Geology, v. 3,
p. 131-60.
Barrett, T. J., and MacLean, W. H., 1999. Volcanic sequences,
lithogeochemistry and hydrothermal alteration in some
bimodal volcanic-associated massive sulfide systems: Reviews
in Economic Geology, v. 8, p. 10131.
Barrett, T. J., Jarvis, I., and Jarvis, K. E., 1990. Rare earth element
geochemistry of massive sulfides-sulfates and gossans on the
Southern Explorer Ridge: Geology, v. 18, p. 583-6.
Barrett, T. J., Cattalani, S., and MacLean, W H., 1993. Volcanic
lithogeochemistry and alteration at the Delbridge massive
sulfide deposit, Noranda, Quebec: Journal of Geochemical
Exploration, v. 48, p. 135-73.
Barrett, T. J., MacLean, W. H., andTennant, S. C, 2001. Volcanic
sequence and alteration at Parys Mountain volcanic-
hosted massive sulfide deposit, Wales, United Kingdom
applications of immobile element lithogeochemistry:
Economic Geology, v. 96, p. 1279305.
Barriga, F. J. A. S., and Fyfe, W. S., 1988. Giant pyritic base-metal
deposits - the example of Feitais (Aljustrel, Portugal):
Chemical Geology, v. 69, p. 331-43.
Barriga, E J. A. S., and Kerrich, R., 1984.
18
O-enriched volcanics
and
18
O-evolved marine water, Aljustrel, Iberian pyrite belt
- transition from high to low Raleigh number convective
regimes: Geochimica et Cosmochimica Acta, v. 48, p. 1021
32.
Barriga, F. J. A. S., Carvalho, D., and Oliveira, J. T, 1983.
Carboniferous volcanogenic sulfide mineralisations in
South Portugal, Iberian pyrite belt: Memorias dos Servicos
Geologicos de Portugal, v. 29, p. 99113.
Bateman, A. M., 1951. Economic mineral deposits: New York, USA,
John Wiley and Sons, 916 p.
Beams, S. D., Dronseika, E. V, and Doyle, M. G., 1998. The
exploration history, geology and geochemistry of the
polymetallic Highway-Reward deposit, Mt Windsor
Subprovince [abs.], in Beams, S. D., ed. Economic geology
of northeast Queensland, the 1998 perspective: Townsville,
Australia, July 6-10, 1998, Geological Society of Australia
Abstracts, v. 49 , p. 189-206.
Beane, R. E., 1982. Hydrothermal alteration in silicate rocks -
southwestern North America, in Titley, S. R., ed., Advances
in geology of porphyry copper deposits - southwestern
North America: Tucson, USA, University of Arizona Press,
p. 117-37.
Beane, R. E., and Titley, S. R., 1981. Porphyry copper deposits - Part
II, hydrothermal alteration and mineralisation: Economic
Geology, 75
th
anniversary volume, p. 23569.
Beaty, D. W, and Taylor, H. P., Jr., 1979. Oxygen-isotope
geochemistry of the Abitibi greenstone belt, Ontario; evidence
for seawater-rock interaction and implications regarding the
isotopic composition and evolution of the ocean and oceanic
crust [abs.]: Geological Society of America, Abstracts with
Programs, v. 11, p. 386.
Beaudoin, G., and Therrien, P., 1999. Isotopic fractionation
calculator, Laboratoie de geochimie isotopique, Universite
Laval, Quebec <www.ggl.ulaval.ca/personnel/beaudoin/
labo.html>.
Bell, T. H., and Etheridge, M. A., 1973. Microstructure of mylonites
and their descriptive terminology: Lithos, v. 6, p. 337-48.
Berry, R. E, and Crawford, A. J., 1988. The tectonic significance
of Cambrian allochthonous mafic-ultramafic complexes
in Tasmania: Australian Journal of Earth Sciences, v. 35,
p. 523-33.
Berry, L. G., Mason, B., and Dietrich, R. V, 1983. Mineralogy-
concepts, descriptions, determinations: San Francisco, USA,
WH. Freeman and Co., 561 p.
Berry, R. E, Huston, D. L, Stolz, A. J., Hill, A. P., Beams, S. D.,
Kuronen, U., and Taube, A., 1992. Stratigraphy, structure,
and volcanic-hosted mineralisation of the Mount Windsor
Subprovince, north Queensland, Australia: Economic
Geology, v. 87, p. 739-63.
Bettison-Varga, L., Varga, R. J., and Schiffman, P., 1992. Relation
between ore-forming hydrothermal systems and extensional
deformation in the Solea Graben spreading centre, Troodos
Ophiolite, Cyprus: Geology, v. 20, p. 987-90.
Bevins, R. E., and Robinson, D., 1992. Low-grade metamorphism:
Geology Today, JanuaryFebruary, p. 237-
Bierlein, F. P., Arne, D. C, McKnight, S., Lu, J., Reeves, S., Besanko,
J., Marek, J., and Cooke, D., 2000. Wall-rock petrology and
geochemistry in alteration halos associated with mesothermal
gold mineralisation, central Victoria, Australia: Economic
Geology, v. 95, p. 283-311.
Bierwirth, P., Huston, D., and Blewett, R. S., 2002. Hyperspectral
mapping of mineral assemblages associated with gold
mineralisation in the central Pilbara, Western Australia:
Economic Geology, v. 97, p. 81926.
Bigger, S. E., and Hanson, R. E., 1992. Devitrification textures and
related features in the Carlton Rhyolite in the Blue Creek
Canyon area, Wichita Mountains, southwestern Okalahoma:
Okalahoma Geological Notes, v. 52, p. 12442.
Bischoff J. L., and Dickson, F. W, 1975. Seawater-basalt interaction
at 200
c
C and 500 bars: implications for origin of seafloor
heavy metal deposits and regulation of seawater chemistry:
Earth and Planetary Science Letters, v. 25, p. 38597.
Bischoff, J. L, and Seyfried, W E., Jr., 1978. Hydrothermal
chemistry of seawater from 25C to 350C: American
Journal of Science, v. 278, p. 838-60.
Bish, D. L., and Aronson, J. L., 1993. Paleogeothermal and
paleohydrologic conditions in silicic tuff from Yucca
mountains, Nevada: Clays and Clay Minerals, v. 41, p. 148
61.
Blatt, H., Middleton, G., and Murray, R., 1972. Origin of
sedimentary rocks: New Jersey, USA, Prentice-Hall Inc.,
634 p.
Boda, S. P., 1991. The geology, structural setting and genesis of the
Chester Mine, northwest Tasmania: Unpublished Honours
thesis, Australian National University, 111 p.
Bodon, S. B., and Cooke, D. R., 1998. Geochemical modelling of
low temperature (5 to 150C) seawater-andesite interaction
implications for regional alteration assemblages in VHMS
districts: AMIRA/ARC Project P439, Centre for Ore
Deposit and Exploration Studies, University of Tasmania,
Unpublished Report 3, October 1996, p. 121-33.
Bodon, S. B., and Valenta, R. K., 1995. Primary and tectonic
features of the Currawong Zn-Cu-Pb(-Au) massive sulfide
deposit, Benambra, Victoria implications for ore genesis:
Economic Geology, v. 90, p. 1694-721.
Bohlke, J. K., Honnorez, J., and Honnorez-Guerstein, B. M., 1980.
Alteration of basalts from Site 396B, Deep Sea Drilling Project
petrographic and mineralogical studies: Contributions to
Mineralogy and Petrology, v. 73, p. 341-64.
Boles, J. R., 1974. Structure, stratigraphy and petrology of mainly
Triassic rocks, Hokonui Hills, Southland, New Zealand:
New Zealand Journal of Geology and Geophysics, v. 17,
p.337-74.
Boles, J. R., 1977. Zeolites in low-grade metamorphic rocks, in
Mumpton, F. A., ed., Mineralogy and geology of natural
zeolites: Mineral Society of America Short Course Notes,
v. 4, p. 103-35.
Boles, J. R., 1982. Active albitization of plagioclase, Gulf Coast
Tertiary: American Journal of Science, v. 282, p. 165-80.
Boles, J. R., and Coombs, D. S., 1975. Mineral reactions in zeolitic
Triassic tuff, Hokonui Hills, New Zealand: Geological
Society of America Bulletin, v. 86, p. 163-73.
Boles, J. R., and Coombs, D. S., 1977. Zeolite facies alteration
of sandstone in the Southland Syncline, New Zealand:
American Journal of Science, v. 277, p. 9821012.
Bonatti, E., 1965. Palagonite, hyaloclastites and alteration of
volcanic glass in the ocean: Bulletin Volcanologique, v. 28,
p. 257-70.
Boulter, C. A., 1993. Comparison of RioTinto, Spain, and Guaymas
Basin, Gulf of California - an explanation of a supergiant
massive sulfide deposit in an ancient sill-sediment complex:
Geology, v. 21, p. 801-4.
Braithwaite, R. L., 1974. The geology and origin of the Rosebery ore
deposit, Tasmania: Economic Geology, v. 69, p. 1086-101.
Branney, M. J., and Sparks, R. S. J., 1990. Fiamme formed by
diagenesis and burial compaction in soils and subaqueous
sediments: Journal of the Geological Society of London,
v. 147, p. 919-22.
Branney, M. J., and Suthren, R. J., 1988. High-level peperitic sills in
the English Lake District distinction from block lavas and
implications for Borrowdale Volcanic Group stratigraphy:
Geological Journal, v. 23, p. 171-87.
Brauhart, C. W., Groves, D. I., and Morant, P., 1998. Regional
alteration systems associated with volcanogenic massive
sulfide mineralisation at Panorama, Pilbara, Western
Australia: Economic Geology, v. 93, p. 292-302.
Brauhart, C. W., Huston, D. L., and Andrew, A. S., 2000. Oxygen
isotope mapping in the Panorama VMS district, Pilbara
Craton, Western Australia applications to estimating
temperatures of alteration and to exploration: Mineralium
Deposita, v. 35, p. 727-40.
Brauhart, C. W, Huston, D. L., Groves, D. I., Mikucki, E. J., and
REFERENCES | 2 5 3
Gardoll, S. J., 2001. Geochemical mass-transfer patterns as
indicators of the architecture of a complete volcanic-hosted
massive sulfide hydrothermal alteration system, Panorama
District, Pilbara, western Australia: Economic Geology,
v. 96, p. 1263-78.
Brey, G., and Schmincke, H-U., 1980. Origin and diagenesis of
the Roque Nublo Breccia, Gran Canaria (Canary Islands)
- petrology of Roque Nublo volcanics, II: Bulletin
Volcanologique, v. 43, p. 1533.
Browne, P. R. L., 1978. Hydrothermal alteration in active geothermal
fields: Annual Review of Earth and Planetary Sciences, v. 6,
p. 229-50.
Burnham, C. W, 1979. Magmas and hydrothermal fluids, in Barnes,
H. L., ed., Geochemistry of hydrothermal ore deposits, 3
rd
edition: New York, USA, John Wiley and Sons, p. 71136.
Burnham, C. W, and Ohmoto, H., 1980. Late-stage processes of
felsic magmatism: Mining Geology, v. 8, p. 1-11.
Busby-Spera, C. J., and White, J. D. L., 1987. Variation in peperite
textures associated with differing host-sediment properties:
Bulletin of Volcanology, v. 49, p. 765-75.
Butler, B. S., and Burbank, W S., 1929. The copper deposits of
Michigan: United States Geological Survey Professional
Paper 144, 238 p.
Butterfield, D. A., 2000. Deep ocean hydrothermal vents, in
Sigurdsson, H., ed., Encyclopaedia of volcanoes, San Diego,
California, USA, Academic Press, p. 857-75.
Cagatay, M. N., 1993. Hydrothermal alteration associated with
volcanogenic massive sulfide deposits examples from
Turkey: Economic Geology, v. 88, p. 606-21.
Callaghan, T, 1998. Geology and alteration of the Mt Julia deposit,
Henty gold mine, Tasmania: Unpublished MSc thesis,
University of Tasmania, 78 p.
Callaghan, T, 2001. Geology and host-rock alteration of the Henty
and Mt Julia gold deposits, western Tasmania: Economic
Geology, v. 96, p. 1073-88.
Campana, B., and King, D., 1963. Palaeozoic tectonism,
sedimentation and mineralisation in west Tasmania: Journal
of Geological Society of Australia, v. 10, p. 1-54.
Campbell, I. H., Franklin, J. M., Gorton, M. P., Hart, T. R., and
Scott, S. D., 1981. The role of subvolcanic sills in the
generation of massive sulfide deposits: Economic Geology,
v. 76, p. 2248-53.
Cann, J. R., 1969. Spilites from the Carlsberg Ridge, Indian Ocean:
Journal of Petrology, v. 10, p. 119.
Carmichael, I. S. E., 1979. Glass and glassy rocks, in Yoder, H. S.
J., ed., The evolution of igneous rocks: Princeton, USA,
Princeton University Press, p. 233-44.
Carozzi, A. V, 1960. Microscopic sedimentary petrography: New
York, USA, John Wiley and Sons, 485 p.
Carson, D. J. T, 1973. Petrography, chemistry, age and emplacement
of the plutonic rocks of Vancouver Island: Canada Geological
Survey Paper, v. 7244, 69 p.
Cartwright, I., 1999. Regional oxygen isotope zonation at Broken
Hill, New South Wales, Australia - large-scale fluid flow
and implications for Pb-Zn-Ag mineralisation: Economic
Geology, v. 94, p. 357-73.
Carvalho, C. M. N., and Barriga, E J. A. S., 2000. Preliminary
report on the detailed mineralogical study at the alteration
zones surrounding the Feitais VMS ore body (Aljustrel,
Portugal, IPB) [abs.], in Gemmell, J. B., and Pongratz, J.,
eds., Volcanic environments and massive sulfide deposits,
Hobart, Australia, November 16-19, 2000, CODES Special
Publication, v. 3, p. 19-21.
Carvalho, D., Barriga, E J. A. S., and Munha, J., 1999. The Iberian
pyrite belt of Portugal and Spain examples of bimodal
siliciclastic systems: Reviews in Economic Geology, v. 8,
p. 385-418.
2 5 4 | REFERENCES
Cas, R. A. E, and Wright, J. V., 1987. Volcanic successions-modern
and ancient - a geological approach to processes, products
and successions: Oxford, UK, Unwin Hyman Inc., 528 p.
Casey, W. H., and Bunker, B., 1990. Leaching of mineral and glass
surfaces during dissolution: Reviews in Mineralogy, v. 23,
p. 397-426.
Cathelineau, M., and Nieva, D., 1985. A chlorite solid solution
geothermometer the Los Azufres (Mexico) geothermal
system: Contributions to Mineralogy and Petrology, v. 91,
p. 235-44.
Cathles, L. M., 1977. An analysis of the cooling of intrusives by
ground-water convection, which includes boiling: Economic
Geology, v. 72, p. 804-26.
Cathles, L. M., 1981. Fluid flow and genesis of hydrothermal ore
deposits: Economic Geology, 75
th
anniversary volume,
p. 424-57.
Cathles, L. M., 1983. An analysis of the hydrothermal system
responsible for massive sulfide deposition in the Hokuroku
Basin of Japan: Economic Geology Monographs, v. 5,
p. 439-87.
Cathles, L. M., 1993. Oxygen isotope alteration in the Noranda
mining district, Abitibi greenstone belt, Quebec: Economic
Geology, v. 88, p. 1483-511.
Cerling, T. E., Brown, E H., and Bowman, J. R., 1985- Low-
temperature alteration of volcanic glass hydration, Na, K,
18
O and Ar mobility: Chemical Geology, v. 52, p. 281-93.
Christiansen, R. L., and Lipman, P. W., 1966. Emplacement and
thermal history of a rhyolite lava flow near Fortymile
Canyon, southern Nevada: Geological Society of America
Bulletin, v. 77, p. 671-84.
Coombs, D. S., 1954. The nature and alteration of some Triassic
sediments from Southland, New Zealand: Transactions of
the Royal Society of New Zealand, v. 82, p. 65-109.
Coombs, D. S., 1974. On the mineral facies of spilitic rocks and
their genesis, in Amstutz, G. C, ed., Spilites and spilitic
rocks: New York, USA, Springer-Verlag, p. 373-86.
Coombs, D. S., Ellis, A. J., Fyfe, W. S., and Taylor, A. M., 1959.
The zeolite facies, with comments on the interpretation
of hydrothermal syntheses: Geochimica et Cosmochimica
Acta, v. 17, p. 53-107.
Corbett, K. D., 1979. Stratigraphy, correlation and evolution of the
Mt Read Volcanics in the Queenstown, Jukes-Darwin and
Mt Sedgwick areas: Geological Survey Tasmania Bulletin,
v. 58, 74 p.
Corbett, K. D., 1981. Stratigraphy and mineralisation in the Mt
Read Volcanics, western Tasmania: Economic Geology,
v. 76, p. 209-30.
Corbett, K. D., 1986. The geological setting of mineralisation in the
Mt Read Volcanics, in Large, R. R., ed., The Mount Read
Volcanics and associated ore deposits: Burnie, Australia,
November 20-21, 1986, Geological Society of Australia,
Tasmanian Division, Proceedings, p. 110.
Corbett, K. D., 1989. Stratigraphy, palaeogeography and
geochemistry of the Mt Read Volcanics, in Burnett, C. E,
and Martin, E. L., eds., Geology and mineral resources of
Tasmania: Geological Society of Australia Special Publication,
v. 15, p. 86-106.
Corbett, K. D., 1992. Stratigraphic-volcanic setting of massive
sulfide deposits in the Cambrian Mount Read Volcanics,
Tasmania: Economic Geology, v. 87, p. 56486.
Corbett, K. D., 1994. Stratigraphic mapping, Tyennan connections,
Cambrian orogenies, the Arthur Lineament, and the
tectonic context of the Mount Read Volcanics - a fresh look
at western Tasmania [abs.], in Cooke, D. R., and Kitto, P. A.,
eds., Contentious issues in Tasmanian geology, a symposium:
Hobart, Australia, November 3-4, 1994, Geological Society
of Australia Abstracts, v. 39, p. 35-7.
Corbett, K. D., 2001. New mapping and interpretations of the
Mount Lyell mining district, Tasmania - a large hybrid Cu-
Au system with an exhalative Pb-Zn top: Economic Geology,
v. 96, p. 1089-122.
Corbett, K. D., 2002. Western Tasmanian regional minerals
program, Mount Read Volcanics compilation - updating
the geology of the Mount Read Volcanics belt: Tasmanian
Geological Survey Record, no. 19, 31 p.
Corbett, K. D., and Komyshan, P., 1989. Geology of the Hellyer-Mt
Charter area: Mt Read Volcanics Project Geological Report,
Tasmania Department of Mines, no. 1, 48 p.
Corbett, K. D., and Lees, T. C, 1987. Stratigraphic and structural
relationships and evidence for Cambrian deformation at
the western margin of the Mt Read Volcanics, Tasmania,
Australia: Australian Journal of Earth Sciences, v. 34, p. 45-
67.
Corbett, K. D., and Solomon, M., 1989. Cambrian Mt Read
Volcanics and associated mineral deposits, in Burrett, C.
E, and Martin, E. L., eds., Geology and mineral resources
of Tasmania: Geological Society of Australia Special
Publication, v. 15, p. 84-153.
Corbett, K. D., Reid, K. O., Corbett, E. B., Green, G. R., Wells,
K., and Sheppard, N. W, 1974. The Mount Read Volcanics
and Cambro-Ordovician relationships at Queenstown,
Tasmania: Journal of Geological Society of Australia, v. 21,
p. 173-86.
Costa, U. R., Barnett, R. L., and Kerrich, R., 1983. The Mattagami
Lake mine Archaean Zn-Cu sulfide deposit, Quebec
- hydrothermal co-precipitation of talc and sulfides in a
seafloor brine pool evidence from geochemistry,
ls
O/
i6
O
and mineral chemistry: Economic Geology, v. 78, p. 1144
203.
Costa, I. A., Barriga, E J. A. S., and Fouquet, Y., 1995. Active
hydrothermal sites and ancient massive sulfides - Lucky
Strike versus Iberian pyrite belt [abs.]: Terra Abstracts, v. 7,
p. 212.
Coutts, B. P. C, 1990. The geology, geochemistry and hydrothermal
alteration of the Hollway Andesite, western Tasmania:
Unpublished Honours thesis, University of Tasmania, 76 p.
Cox, S. F., 1981. The stratigraphic and structural setting of the Mt
Lyell volcanic-hosted sulfide deposits: Economic Geology,
v. 76, p. 231-45.
Craig, M. A., 2001. Regolith mapping for geochemical exploration
in the Yilgarn Craton, Western Australia: Geochemistry
- Exploration, Environment, Analysis, v. 1, p. 383-90.
Craig, J. R., and Vaughan, D. J., 1981. Ore microscopy and ore
petrology: New York, USA, John Wiley and Sons, 406 p.
Crawford, A. J., and Berry, R. F., 1992. Tectonic implications of
Late Proterozoic-Early Palaeozoic igneous rock associations
in western Tasmania: Tectonophysics, v. 214, p. 37-56.
Crawford, A. J., Corbett, K. D., and Everard, J. L, 1992.
Geochemistry of the Cambrian volcanic-hosted massive
sulfide-rich Mount Read Volcanics, Tasmania, and some
tectonic implications: Economic Geology, v. 87, p. 597-
619.
Creasey, S. C, 1959. Some phase relations in hydro thermally altered
rocks of porphyry copper deposits: Economic Geology,
v. 54, p. 351-73.
CSIRO, 2002. Automated mineralogical logging of drill core, chips
and powders: CSIRO/AMIRA Project P685, <www.syd.dem.
csiro.au/research/MMTG/P685%20Core%20Logging/
P685.htm>.
Cudahy, T. J., Okada, K., and Brauhart, C, 2000. Targeting VMS-
style Zn mineralisation at Panorama, Australia, using
airborne hyperspectral VNIR-SW1R Hymap data: 14
th
International Conference on Applied Geologic Remote
Sensing: Ann Arbor, Las Vegas, USA, November 6-8, 2000,
Proceedings, v. 14, p. 395-402.
Dana, E. S., 1957. A textbook of mineralogy, with an extended
treatise on crystallography and physical mineralogy, (4
th
edition, revised and enlarged by W. E. Ford): New York,
USA, Wiley, 851 p.
Date, J., Watanabe, Y., Iwaya, S., and Horiuchi, M., 1979. A
consideration on the alteration of dacite below the Fukazawa
ore deposits, Fukazawa Mine, Akita Prefecture: Mining
Geology, v. 29, p. 187-96.
Date, J., Watanabe, Y., and Saeki, Y., 1983. Zonal alteration around
the Fukazawa Kuroko deposits, Akita Prefecture, northern
Japan: Economic Geology Monographs, v. 5, p. 365-86.
Davidson, P., 1998. The Murchison Granite: Unpublished Honours
thesis, University of Tasmania, 128 p.
Davidson, G. J., and Kitto, P., 1997. The sulfur isotope signature
of Cambrian faults in the Mt Read Volcanic Belt and their
implications for sulfur enrichment in submarine volcano-
sedimentary rift basins: AMIRA/ARC Project P291A,
Centre for Ore Deposit Research, University of Tasmania,
Unpublished Final Report, March 1997, p. 121-69.
Davidson, G. J., Garven, G., Kitto, P. A., and Berry, R., 2000.
Geochemically discrete fluid bodies formed by convection
at the heated edge of porous seafloor aquifers [abs.], in
Woodhead, J. D., Hergt, J. M., and Noble, W P., eds.,
Beyond 2000, new frontiers in isotope geoscience: Lome,
Australia, January 30-February 4, 2000, Abstracts and
Proceedings, p. 39-40.
Davies, J. F., Whitehead, R. E., Huang, J., and Nawaratne, S.,
1990. A comparison of progressive hydrothermal carbonate
alteration in Archaean metabasalts and metaperidotites:
Mineralium Deposita, v. 25, p. 65-72.
Davis, L. W, 1990. Silver-lead-zinc-copper mineralisation in the
Captains Flat-Goulburn synclinorial zone and the Hill
End synclinorial zone: Australasian Institute of Mining and
Metallurgy Monograph, v. 14, p. 1375-84.
Davis, B. K., andMcPhie, J., 1996. Spherulites, quench fractures and
relict perlite in a Late Devonian rhyolite dyke, Queensland,
Australia: Journal of Volcanology and Geothermal Research,
v. 71, p. 1-11.
deRonde, C.E.J., 1995. Fluid chemistry and isotopic characteristics of
seafloor hydrothermal systems and associated VMS deposits:
potential for magmatic contributions, in Thompson, J. F.
H., ed., Magmas, fluids and ore deposits: Victoria, Canada,
Mineralogical Association of Canada, v. 23, p. 479-509.
de Ronde, C. E. J., de Wit, M. J., and Spooner, E. T. C, 1994. Early
Archaean (>3.2 Ga) Fe-oxide-rich, hydrothermal discharge
vents in the Barberton greenstone belt, South Africa:
Geological Social of America Bulletin, v. 106, p. 86-104.
de Rosen-Spence, A. F., Provost, G., Dimroth, E., Gochnauer, K.,
and Owen, V, 1980. Archaean subaqueous felsic flows,
Rouyn-Noranda, Quebec, Canada, and their Quaternary
equivalents: Precambrian Research, v. 12, p. 43-77.
Deelman, J. C, 1975. 'Pressure solution' re-examined: Neues
Jahrbuch fuer Geologie und Palaeontologie Abhandlungen,
v. 149, p. 384-97.
Deelman, J. C, 1976. Lithification analysis-experimental
observations: Geologishe Rundschau, v. 65, p. 105578.
Deer, W. A., Howie, R. A., and Zussman, J., 1966. An introduction
to the rock forming minerals: London, UK, Longman
Scientific and Technical, 528 p.
Denniss, A. M., Colman, T. B., Cooper, D. C, Hatton, W. A.,
and Shaw, M. H., 1999. The combined use of PIMA and
Vulcan technology for mineral deposit evaluation at the
Parys Mountain Mine, Anglesea, UK: 13* International
Conference on Applied Geologic Remote Sensing, Vancouver,
Canada, March 1-3, 1999, Proceedings, p. 125-32.
Destrigneville, C, Schott, J., Caristan, Y, and Agrinier, P., 1991.
REFERENCES | 2 5 5
Evidence of an early alteration process driven by magmatic
fluid in Mururoa Volcano: Earth and Planetary Science
Letters, v. 104, p. 119-39.
Dickinson, W. R., 1962. Petrology and diagenesis of Jurassic
andesitic strata in central Oregon: American Journal of
Science, v. 260, p. 481-500.
Dickson, F. W, and Potter, J. M., 1982. Rock-brine chemical
interactions: Electric Power Research Institute Project 653-
2, Final Report, AP-2258, 89 p.
Dill, H. G., Bosse, H. R., Henning, K. H., Fricke, A., and Ahrendt,
H., 1997. Mineralogical and chemical variations in hypogene
and supergene kaolin deposits in a mobile fold belt in the
Central Andes of northwestern Peru: Mineralium Deposita,
v. 32, p. 149-63.
Dimroth, E., and Lichtblau, A. P., 1979. Metamorphic evolution of
Archaean hyaloclastites, Noranda area - Part 1, comparison
of Archaean and Cenozoic seafloor metamorphism: Canadian
Journal of Earth Science, v. 16, p. 131540.
Downs, R. C, 1993. Syndepositional fault controls on the Hellyer
volcanic-hosted massive sulfide deposit: Unpublished MSc
thesis, University of Tasmania, 63 p.
Doyle, M. G., 1990. The geology, mineralisation and alteration
of the Jukes Proprietary prospect, Tasmania: Unpublished
Honours thesis, University of Tasmania, 114 p.
Doyle, M. G., 1997. A Cambro-Ordovician submarine volcanic
succession hosting massive sulfide mineralisation: Mount
Windsor Subprovince, Queensland: Unpublished PhD
thesis, University of Tasmania, 264 p.
Doyle, M. G., 2001. Volcanic influences on hydrothermal and
diagenetic alteration - evidence from Highway-Reward,
Mount Windsor Subprovince, Australia: Economic Geology,
v. 96, p. 1133-48.
Doyle, M. G., and Allen, R. L., 2003. Subseafloor replacement
in volcanic-hosted massive sulfide deposits: Ore Geology
Reviews, v. 23, p. 183-222.
Doyle, M. G., and Huston, D. L., 1999. The subseafloor
replacement origin of the Ordovician Highway-Reward
volcanic-associated massive sulfide deposit, Mount Windsor
Subprovince, Australia: Economic Geology, v. 94, p. 825-
43.
Doyle, M. G., and McPhie, J., 2000. Facies architecture of a
silicic intrusion-dominated volcanic centre at Highway-
Reward, Queensland, Australia: Journal of Volcanology and
Geothermal Research, v. 99, p. 79-96.
Drysdale, D. J., 1991. Perlitic texture and other fracture patterns
produced by hydration of glassy rocks: Department of
Geology, University of Queensland Scientific Paper, v. 12,
no. 3, p. 278-85.
Dudas, F. O., Campbell, I. H., and Gorton, M. P., 1983. Geochemistry
of igneous rocks in the Hokuroku District, northern Japan:
Economic Geology Monographs, v. 5, p. 115-33.
Dugdale, J. S., 1992. Lithostratigraphy of the White Spur area,
western Tasmania: Unpublished Honours thesis, University
of Tasmania, 93 p.
Duhig, N. C, Stolz, J., Davidson, G. J., and Large, R. R., 1992.
Cambrian microbial and silica gel textures in silica-iron
exhalites from the Mount Windsor volcanic belt, Australia
their petrography, chemistry, and origin: Economic Geology,
v. 87, p. 764-84.
Eastoe, C. J., Solomon, M., and Walshe, J. L., 1987. District-
scale alteration associated with massive sulfide deposits in
the Mount Read Volcanics, western Tasmania: Economic
Geology, v. 82, p. 1239-58.
Eggleton, R. A., and Keller, J., 1982. The palagonitization of
limburgite glass - a TEM study: Neues Jahrbuch fur
Mineralogie Monatshefte, p. 321-36.
Eilu, P., Mikucki, E. J., and Groves, D. I., 1997. Wallrock alteration
2 5 6 | REFERENCES
and primary geochemical dispersion in lode-gold exploration:
4
th
biennial meeting of the Society for Geology Applied to
Mineral Deposits, Turku, Finland, August 11-13, 1997,
Short course notes, p. 65.
Einsele, G., 1985. Basaltic sill-sediment complexes in young
spreading centres genesis and significance: Geology, v. 13,
p. 249-52.
Einsele, G., Gieskes, J. M., Curray, J., Moore, D. M., Aguayo, E.,
Aubry, M., Fornari, D., Guerrero, J., Kastner, M., Kelts, K.,
Lyle, M., Matoba, Y., Molina-Cruz, A., Niemitz, J., Rueda,
J., Saunders, A., Schrader, H., Simoneit, B., and Vacquier, V.,
1980. Intrusion of basaltic sill into highly porous sediments
and resulting hydrothermal activity: Nature, v. 283, p. 441
5 .
Eldridge, C. S., Barton, P. B. J., and Ohmoto, H., 1983. Mineral
textures and their bearing on formation of the Kuroko ore
bodies: Economic Geology Monographs, v. 5, p. 241-81.
Elliott-Meadows, S. R., and Appleyard, E. C., 1991. The alteration
geochemistry and petrology of the Lar Cu-Zn deposit, Lynn
Lake area, Manitoba, Canada: Economic Geology, v. 86,
p.486-505.
Ellis, A. J., and Mahon, W. A. J., 1964. Natural hydrothermal systems
and experimental hot-water/rock interactions: Geochimica
et Cosmochimica Acta, v. 28, p. 132357.
Ewart, A., 1979. A review of the mineralogy and chemistry of
Tertiary-Recent dacitic, latitic, rhyolitic, and related salic
volcanic rocks, in Barker, E, ed., Trondhjemites, dacites, and
related rocks: Amsterdam, Netherlands, Elsevier Scientific
Publication, p. 13121.
Faure, G., 1986. Principles of isotope geology: New York, USA,
John Wiley and Sons, 589 p.
Finlow-Bates, T, and Strumpfl, E. E, 1981. The behaviour of so-
called immobile elements in hydrothermally altered rocks
associated with volcanogenic submarine-exhalative ore
deposits: Mineralium Deposita, v. 16, p. 319-28.
Fisher, R. V, I960. Classification of volcanic breccias: Geological
Society of America Bulletin, v. 71, p. 973-82.
Fisher, R. V., 1961. Proposed classification of volcaniclastic sediments
and rocks: Geological Society of America Bulletin, v. 72,
p.1409-14.
Fisher, R. V., and Schmincke, H-U., 1984. Pyroclastic rocks: New
York, USA, Springer-Verlag, 472 p.
Fiske, R. S., 1969. Recognition and significance of pumice in marine
pyroclastic rocks: Geological Society of America Bulletin,
v. 80, p. 1-8.
Fitzgerald, F. G., 1974. The Primrose Pyroclastics in the Hercules-
White Spur area, western Tasmania: Unpublished Honours
thesis, University of Tasmania, 188 p.
Floyd, P. A., and Winchester, J. A., 1978. Identification and
discrimination of altered and metamorphosed volcanic rocks
using immobile elements: Chemical Geology, v. 21, p. 291-
306.
Folk, R. L., 1965. Some aspects of recrystallisation in ancient
limestones: Society of Economic Palaeontologists and
Mineralogists Special Publication, v. 13, p. 1448.
Fournier, R. O., 1985. The behaviour of silica in hydrothermal
solutions: Reviews in Economic Geology, v. 2, p. 45-61.
Fournier, R. O., and Potter, R. W, II, 1982. An equation correlating
the solubility of quartz in water from 20 to 900C at
pressures up to 10,000 bars: Geochimica et Cosmochimica
Acta, v. 46, p. 1969-73.
Fowler, A. D., Jensen, L. S., and Peloquin, S. A., 1987. Varioles in
Archaean basalts products of spherulitic crystallization:
The Canadian Mineralogist, v. 25, p. 275-89.
Franklin, J. M., Kasarda, J., and Poulsen, K. H., 1975. Petrology
and chemistry of the alteration zone of the Mattabi massive
sulfide deposit: Economic Geology, v. 70, p. 6379.
Franklin, J. M., Gibb, W, Poulsen, K. H., and Severin, P., 1977.
Archean metallogeny and stratigraphy of the South Sturgeon
Lake area: Ontario, Canada, Institute Lake Superior Geology,
73 p.
Franklin, J. M., Sangster, D. M., and Lydon, J. W, 1981. Volcanic-
associated massive sulfide deposits: Economic Geology, 75
th
anniversary volume, p. 485627.
Franklin, J. M., Sangster, D. E, Roscoe, S. M., and Loveridge, W
D., 1983. Lead isotope studies in Superior and Southern
provinces: Geological Survey of Canada Bulletin, v. 351,
60 p.
Friedman, I., and Long, W, 1984. Volcanic glasses their origins
and alteration processes: Journal of Non-Crystalline Solids,
v. 67, p. 127-33.
Friedman, I., and Smith, R. L., 1958. The deuterium content of
water in some volcanic glasses: Geochimica et Cosmochimica
Acta, v. 15, p. 218-28.
Friedman, I., Smith, R. L., and Long, W. D., 1966. Hydration of
natural glass and formation of perlite: Geological Society of
America Bulletin, v. 77, p. 323-8.
Fritsch, K. V, and Reiss, W, 1868. Geologische Beschreibung der
Insel Teneriffa: Winterthur, Verlag von Wurster and Co.,
494 p.
Fuller, R. E., 1938. Deuteric alteration controlled by the jointing of
lavas: American Journal of Science, v. 35, p. 16171.
Furbish, W. J., and Schrader, E. L., 1980. Secondary minerals of
deuteric and diagenetic origin filling voids and encrusting
surfaces on basalts from Deep Sea Drilling Project Leg 54:
Proceedings of the Ocean Drilling Program, Initial Reports,
v. 54, p. 807-18.
Furnes, H., 1975. Experimental palagonitization of basaltic glasses
of varied composition: Contributions to Mineralogy and
Petrology, v. 50, p. 105-13.
Fyfe, W. S., Turner, F. J., and Verhoogen, J., 1958. Metamorphic
reactions and metamorphic facies: Geological Society of
America Memoir, v. 73, 259 p.
Galley, A. G., 1993. Characteristics of semi-conformable alteration
zones associated with volcanogenic massive sulfide districts:
Journal of Geochemical Exploration, v. 48, p. 17599.
Galley, A. G., 1995. Target vectoring using lithogeochemistry -
applications to the exploration for volcanic-hosted massive
sulfide deposits: CIM Bulletin, v. 88, p. 15-27.
Galley, A. G., 2003. Composite synvolcanic intrusions associated
with Precambrian VMS-related hydrothermal systems:
Mineralium Deposita, v. 38, p. 443-73.
Galley, A. G., Bailes, A., Hannington, M., Hoik, G., Katsube, J.,
Paquette, F., Paradis, S. J., Santaguida, F., Taylor, B., and
Hillary, B., 2002. Database for CAMIRO Project 94E07
- inter-relationships between subvolcanic intrusions, large-
scale alteration zones and VMS deposits: Geological Survey
of Canada Open File 4431 (CD).
Galloway, J. N., 1979. Alteration of trace metal geochemical cycles
due to the marine discharge of wastewater: Geochimica et
Cosmochimica Acta, v. 38, p. 20718.
Garwin, S., 2002. The geologic setting of intrusion-related
hydrothermal systems near the Batu Hijau porphyry copper-
gold deposit, Sumbawa, Indonesia: SEG Special Publication,
v. 9, p. 333-66.
Gary, M., McAfee, R., Jr., and Wolf, C. L, eds., 1974. Glossary
of geology: Washington, D.C., American Geological
Institution, 805 p.
Geiskes, J. M., Kastner, M., Einsele, G., Kelts, K., and Niemitz, J.,
1982. Hydrothermal activity in the Guaymas Basin, Gulf of
California - a synthesis: Proceedings of the Ocean Drilling
Program, Initial Reports, v. 64, p. 1159-68.
Gemmell, J. B., and Fulton, R., 2001. Geology, genesis, and
exploration implications of the footwall and hanging-
|i
wall alteration associated with the Hellyer volcanic-hosted
massive sulfide deposit, Tasmania, Australia: Economic
Geology, v. 96, p. 1003-35.
Gemmell, J. B., and Herrmann, W, 2001. Preface to a special issue
devoted to alteration associated with volcanic-hosted massive
sulfide deposits, and its exploration significance: Economic
Geology, v. 96, p. 909-12.
Gemmell, J. B., and Large, R. R., 1992. Stringer system and
alteration zones underlying the Hellyer volcanogenic massive
sulfide deposit, Tasmania, Australia: Economic Geology,
v. 87, p. 620-49.
Gemmell, J. B., and Large, R. R., 1993. Evolution of a VHMS
hydrothermal system, Hellyer deposit, Tasmania, Australia
sulphur isotope evidence, in Ishihara, S., Urabe, T, and
Ohmoto, H., eds., Mineral resources symposia: Tokyo,
Japan, August 24, 1992, Society of Resource Geologists of
Japan Special Issue, v. 17, p. 108-19.
Gemmell, J. B., Large, R. R., and Khin Zaw, 1998. Palaeozoic
volcanic-hosted massive sulfide deposits: AGSO Journal of
Australian Geology and Geophysics, v. 17, p. 12937.
Ghiara, M. R., Franco, E., Petti, C, Stanzione, D., and Valentino,
G. M., 1993. Hydrothermal interaction between basaltic
glass, de-ionized water and seawater: Chemical Geology,
v. 104, p. 125-38.
Gibson, H. L., 1989. The mine sequence of the Central Noranda
Volcanic Complex - geology alteration, massive sulfide
deposits and volcanological reconstruction: Unpublished
PhD thesis, Carleton University, 715 p.
Gibson, H. L., and Watkinson, D. H., 1990. Volcanogenic massive
sulfide deposits of the Noranda cauldron and shield volcano,
Quebec: Canadian Institute of Mining and Metallurgy
Special Issue, v. 43, p. 119-32.
Gibson, H. L, Watkinson, D. H., and Comba, C. D. A., 1983.
Silicification hydrothermal alteration in an Archaean
geothermal system within the Amulet Rhyolite Formation,
Noranda, Quebec: Economic Geology, v. 78, p. 954-71.
Gibson, H. L., Morton, R. L., Hudak, G. J., and Harper, G. D., 1999.
Submarine volcanic processes, deposits, and environments
favourable for the location of volcanic-associated massive
sulfide deposits: Reviews in Economic Geology, v. 8, p. 13-
51.
Gibson, H. L., Santaguida, F., Paquette-Mihalasky, F. I., and
Watkinson, D. H., 2000. Evolution of regional semi-
conformable alteration assemblages within an Archaean
subseafloor hydrothermal system, and relationship to VMS
deposits, at Noranda, Quebec, Canada [abs.], in Gemmell,
J. B., and Pongratz, J., eds., Volcanic environments and
massive sulfide deposits: Hobart, Australia, November 16-
19, 2000, CODES Special Publication, v. 3, p. 61-3.
Gifkins, C. C, 2001. Submarine volcanism and alteration in the
Cambrian, northern Central Volcanic Complex, western
Tasmania: Unpublished PhD thesis, University of Tasmania,
239 p.
Gifkins, C. C, and Allen, R. L., 2001. Textural and chemical
characteristics of diagenetic and hydrothermal alteration
in glassy volcanic rocks - examples from the Mount Read
Volcanics, Tasmania: Economic Geology, v. 96, p. 973-
1002.
Gifkins, C. C, and Allen, R. L., 2002. Voluminous, submarine,
intracaldera pumice breccia generated by explosive
eruptions - the Cambrian Mount Black and Kershaw
Pumice Formations, western Tasmania [abs.], in Chapman
Conference Explosive subaqueous volcanism: Dunedin,
New Zealand, January 21-25, 2002, Abstracts Volume,
p. 26.
Gifkins, C. C, and Kimber, B., 2004. Revolutionising our
understanding of an ancient volcanic succession an
REFERENCES | 2 5 7
interactive map of the Mount Read Volcanics, western
Tasmania [abs.], in McPhie, J., and McGoldrick, P. J., eds.,
Dynamic earth past, present and future: Hobart, Australia,
February 8-13, 2004, Geological Society of Australia
Abstracts, v. 73: p. 81.
Gifkins, C. C, McPhie, J., and Allen, R. L, 2002. Pumiceous
rhyolitic peperite in ancient submarine volcanic successions:
Journal of Volcanology and Geothermal Research, v. 114,
p. 181-203.
Gifkins, C. C, Allen, R. L., and McPhie, }., in press. Apparent
welding textures in altered pumice-rich rocks: Journal of
Volcanology and Geothermal Research.
Giggenbach, W. E, 1997. The origin and evolution of fluids in
magmatic hydrothermal systems, in Barnes, H. L., ed.,
Geochemistry of hydrothermal ore deposits: New York,
USA, John Wiley and Sons, p. 737-96.
Gillis, K. M., and Robinson, P. T, 1990. Patterns and processes of
alteration in the lavas and dykes of the Troodos Ophiolite,
Cyprus: Journal of Geophysical Research, v. 95, p. 21523-
48.
Gillis, K. M., Thompson, G., and Kelley, D. S., 1993. A view of
the lower crustal component of hydrothermal systems at the
Mid-Atlantic Ridge: Journal of Geophysical Research, v. 98,
p. 19597-619.
Gilmour, P., 1965. The origin of the massive sulfide mineralisation
in the Noranda district, northwestern Quebec: Proceedings
of the Geological Association of Canada, v. 16, p. 63-81.
Goldfarb, M. S., Converse, D. R., Holland, H. D., and Edmond,
J. M., 1983. The genesis of hot spring deposits on the East
Pacific Rise, 21 N: Economic Geology Monographs, v. 5,
p. 184-97.
Goodfellow, W D., and Franklin, J. M., 1993. Geology, mineralogy,
and chemistry of sediment-hosted clastic massive sulfides in
shallow cores, Middle Valley, northern Juan de Fuca Ridge:
Economic Geology, v. 88, p. 2033-64.
Goodfellow, W. D., and McCutcheon, S. R., 2003. Geologic and
genetic attributes of volcanic sediment-hosted massive
sulfide deposits of the Bathurst mining camp, northern New
Brunswick - a synthesis: Economic Geology Monographs,
v. 11, p. 245-301.
Grant, J. A., 1986. The isocon diagram a simple solution to
Gresens' equation for metasomatic alteration: Economic
Geology, v. 81, p. 1976-82.
Grapes, R. H., 1976. Low-grade alteration of basic rocks in
Hokkaido, Japan: Journal of the Faculty of Science,
Hokkaido University, v. 17, p. 373-85.
Green, G. R., and Iliff, G. D., 1989. Rosebery, in Burrett, C. E,
and Martin, E. L., eds., Geology and mineral resources of
Tasmania: Geological Society of Australia Special Publication,
v. 15, p. 132-7.
Green, G. R., andTaheri, J., 1992. Stable isotopes and geochemistry
as exploration indicators, in Baillie, P., Dix, M., and
Richardson, R., eds., Tasmania an island of potential:
Geological Survey Bulletin, v. 70, p. 84-91.
Green, G. R., Solomon, M., andWalshe, J. L., 1981. The formation
of the volcanic-hosted massive sulfide deposit at Rosebery,
Tasmania: Economic Geology, v. 76, p. 304-38.
Green, G. R., Ohmoto, H., Date, J., and Takahashi, T, 1983.
Whole-rock oxygen isotope distribution in the Fukuzawa-
Kosaka area, Hokuroku District, Japan, and its potential
application to mineral exploration: Economic Geology
Monographs, v. 5, p. 395-411.
Gregory, R. T, and Taylor, H. P., Jr., 1981. An oxygen isotope profile
in a section of Cretaceous oceanic crust, Samail Ophiolite,
Oman - evidence for d
18
O buffering of the oceans by deep
(>5 km) seawater-hydrothermal circulation at mid-ocean
ridges: Journal of Geophysical Research, v. 86, p. 273755.
2 5 8 | REFERENCES
Gregory, P. W., Hartley, J. S., and Wills, K. J. A., 1990. Thalanga
zinc-lead-copper-silver deposit, in Hughes, F. E., ed.,
Geology of the mineral deposits of Australia and Papua New
Guinea: Australasian Institute of Mining and Metallurgy
Monograph, v. 14, p. 1527-37.
Gresens, R. L., 1967. Composition-volume relationships of
metasomatism: Chemical Geology, v. 2, p. 47-65.
Gramme, C. S., Wright, T. L, and Peck, D. L., 1969. Magnetic
properties and oxidation of iron-titanium oxide minerals
in Alae and Makaoputi lava lakes, Hawaii: Journal of
Geophysical Research, v. 74, p. 5277-93.
Guilbert, J. M., and Park, C. F. J., 1986. The geology of ore deposits:
New York, USA, W.H. Freeman and Co., 985 p.
Gulson, B. L., and Porritt, P. M., 1987. Base metal exploration
of the Mount Read Volcanics, western Tasmania Part II,
lead isotope signatures and genetic implications: Economic
Geology, v. 82, p. 291-307.
Gustafson, L. B., and Hunt, J. P., 1975. The porphyry copper
deposit at El Salvador, Chile: Economic Geology, v. 70,
p. 857-912.
Hajash, A., 1975. Hydrothermal processes along mid-ocean ridges -
an experimental investigation: Contributions to Mineralogy
and Petrology, v. 53, p. 20526.
Hajash, A., 1977. Experimental seawater/basalt interactions -
effect of water/rock ratio and temperature gradient [abs.]:
Geological Society of America Abstracts with Programs, v. 9,
p. 1002.
Hajash, A., and Archer, P., 1980. Experimental seawater/basalt
interactions - effects of cooling: Contributions to Mineralogy
and Petrology, v. 75, p. 1-13.
Hajash, A., and Chandler, G. W, 1981. An experimental investigation
of high-temperature interactions between seawater and
rhyolite, andesite, basalt and peridotite: Contributions to
Mineralogy and Petrology, v. 78, p. 240-54.
Halley, S. W, and Roberts, R. H., 1997. Henty - a shallow water
gold-rich volcanogenic massive sulfide deposit in western
Tasmania: Economic Geology, v. 92, p. 43847.
Hannington, M. D., and Herzig, P. M., 2000. Submarine epithermal
deposits and the VMS-epithermal transition a new
exploration target [abs.], in Gemmell, J. B., and Pongratz,
J., eds., Volcanic environments and massive sulfide deposits:
Hobart, Australia, November 16-19, 2000, CODES Special
Publication, v. 3, p. 75-7.
Hannington, M. D., and Scott, S. D., 1989. Gold mineralisation
in volcanogenic massive sulfides implications of data from
active hydrothermal vents on the modern seafloor: Economic
Geology Monographs, v. 6, p. 491-507.
Hannington, M. D., Poulsen, K. H., and Thompson, J. F. H., 1999.
Volcanogenic gold in the massive sulfide environment:
Reviews in Economic Geology, v. 8, p. 325-56.
Hannington, M. D., Santaguida, E, Kjarsgaard, I. M., and Cathles,
L. M., 2003a. Regional-scale hydrothermal alteration in the
Central Blake River Group, western Abitibi subprovince,
Canada - implications for VMS prospectivity: Mineralium
Deposita, v. 38, p. 393-422.
Hannington, M. D., Kjarsgaard, I. M., Galley, A. G., and Taylor,
B. E., 2003b. Mineral-chemical studies of metamorphosed
hydrothermal alteration in the Kristineberg volcanogenic
massive sulfide district, Sweden: Mineralium Deposita,
v. 38, p. 423-43.
Hanor, J. S., 1979. The sedimentary genesis of hydrothermal fluids,
in Barnes, H. L., ed., Geochemistry of hydrothermal ore
deposits, 3
rd
edition: New York, USA, John Wiley and Sons,
p. 137-72.
Hanson, R. E., 1991. Quenching and hydroclastic disruption of
andesitic to rhyolitic intrusions in a submarine island-arc
sequence, northern Sierra Nevada, California: Geological
Society of America Bulletin, v. 103, p. 804-16.
Hart, S. R., 1969. K, Rb, Cs contents and K/Rb, K/Cs ratios of fresh
and altered submarine basalts: Earth and Planetary Science
Letters, v. 6, p. 295-303.
Hashiguchi, H., Yamada, R., and Inoue, T, 1983. Practical
application of low Na
2
O anomalies in footwall acid lava for
delimiting promising areas around the Kosaka and Fukazawa
Kuroko deposits, Akita Prefecture, Japan: Economic Geology
Monographs, v. 5, p. 387-94.
Hawkins, D. B., 1981. Kinetics of glass dissolution and zeolite
formation under hydrothermal conditions: Clays and Clay
Minerals, v. 29, p. 331-40.
Hay, R. L., 1963. Stratigraphy and zeolitic diagenesis of the John Day
Formation of Oregon: University of California Publications
in Geological Sciences, v. 42, p. 199262.
Hay, R. L., 1966. Zeolites and zeolitic reactions in sedimentary
rocks: Geological Society of America Special Paper, v. 85,
130 p.
Hay, R. L., 1977. Geology of zeolites in sedimentary rocks in
Mumpton, F. A., ed., Mineralogy and geology of natural
zeolites: Mineralogical Society of America Short Course
Notes 4, p. 53-63
Hay, R. L., 1978. Geological occurrence of zeolites, in Sand, L. B.,
and Mumpton, F. A., eds., Natural zeolites occurrence,
properties, use: Oxford, UK, Peregamon Press, p. 135-43.
Hay, R. L., and Guldman, S. G., 1987. Diagenetic alteration of silicic
ash in Searles Lake, California: Clays and Clay Minerals,
v. 35, p. 449-57.
Hay, R. L., and Iijima, A., 1968a. Nature and origin of palagonite
tuffs of the Honolulu Group on Oahu, Hawaii: Geological
Society of America Memoir, v. 116, p. 338-76.
Hay, R. L., and Iijima, A., 1968b. Petrology of palagonite tuffs of
KOKO Craters, Oahu, Hawaii: Contributions to Mineralogy
and Petrology, v. 17, p. 141-54.
Heaton, T. H. E., and Sheppard, S. M. F, 1977. Hydrogen and
oxygen isotope evidence for seawater-hydrothermal
alteration and ore deposition, Troodos Complex, Cyprus
[abs.]: London, UK, January, 1976, Institute of Mining and
Metallurgy Monograph - Volcanic processes in ore genesis,
p. 21-22.
Hein, J. R., and Scholl, D. W, 1978. Diagenesis and distribution of
late Cenozoic volcanic sediment in the southern Bering Sea:
Geological Society of America Bulletin, v. 89, p. 197-210.
Hellman, P. L., Smith, R. E., and Henderson, P., 1977. Rare earth
element investigation of the Cliefden outcrop, NSW,
Australia: Contributions to Mineralogy and Petrology, v. 65,
p. 155-64.
Helmold, K. P., and van de Kamp, P C , 1984. Diagenetic mineralogy
and control on albitization and laumontite formation in
Palaeogene arkoses, Santa Ynez Mountains, California, USA:
AAPG Memoir, v. 37, p. 239-76.
Hemley, J. J., and Jones, W. R., 1964. Chemical aspects of
hydrothermal alteration with emphasis on hydrogen
metasomatism: Economic Geology, v. 59, p. 538-67.
Henderson, R. A., 1986. Geology of the Mt Windsor Subprovince
a Lower Palaeozoic volcano-sedimentary terrane in the
northern Tasman Orogenic Zone: Australian Journal of
Earth Sciences, v. 33, p. 343-64.
Henley, R. W, and Brown, K. L., 1985. A practical guide to the
thermodynamics of geothermal fluids and hydrothermal ore
deposits: Reviews in Economic Geology, v. 2, p. 2544.
Henley, H. R., and Ellis, A. J., 1983. Geothermal systems ancient
and modern - a geochemical review: Earth Science Reviews,
v. 19, p. 1-50.
Henley, R. W., and McNabb, A., 1978. Magmatic vapour plumes and
ground-water interaction in porphyry copper emplacement:
Economic Geology, v. 73, p. 1-19.
Henley, R. W., andThornley, P., 1979. Some geothermal aspects of
polymetallic massive sulfide formation: Economic Geology,
v. 74, p. 1600-12.
Henneberger, R. C, and Browne, P. R. L., 1988. Hydrothermal
alteration and evolution of the Ohakuri Hydrothermal
System, Taupo Volcanic Zone, New Zealand: Journal of
Volcanology and Geothermal Research, v. 34, p. 21131.
Herrmann, W., and Berry, R. E, 2002. MINSQ - a least-squares
spreadsheet method for calculating mineral proportions
from whole rock major element analyses: Geochemistry:
Exploration, Environment, Analysis, v. 2, p. 3618.
Herrmann, W., and Hill, A. P., 2001. The origin of chlorite-
tremolite-carbonate rocks associated with the Thalanga
volcanic-hosted massive sulfide deposit, north Queensland,
Australia: Economic Geology, v. 96, p. 114973.
Herrmann, W., and MacDonald, G., 1996. Volcanic facies, alteration
and exploration targets in EL 8/96 South Henty, Tasmania:
Resolute Ltd, Unpublished Report, November 1996, 43 p.
Herrmann, W., Blake, M. D., Doyle, M. G., Huston, D. L., Kamprad,
J., Merry, N., and Pontual, S., 2001. Short wavelength
infrared (SWIR) spectral analysis of hydrothermal alteration
zones associated with base metal sulfide deposits at Rosebery
and Western Tharsis, Tasmania and Highway-Reward,
Queensland: Economic Geology, v. 96, p. 93955.
Herrmann, W., Green, G. R., and Barton, M. D., 2004. Boco
prospect - mineralogic and isotopic implications for VHMS
exploration [abs.], in McPhie, J., and McGoldrick, P. J., eds.,
Dynamic earth - past, present and future: Hobart, Australia,
February 813, 2004, Geological Society of Australia
Abstracts, v. 73, p. 87.
Herzig, P. M., Petersen, S., Hannington, M. D., and Jonasson, I.
R., 2000. Conical Seamount - a submarine analogue of the
Ladolam epithermal gold deposits on Lihir Island, Papua
New Guinea [abs.], in Gemmell, J. B., and Pongratz, J.,
eds., Volcanic environments and massive sulfide deposits:
Hobart, Australia, November 16-19, 2000, CODES Special
Publication, v. 3, p. 83-84.
Hill, A. P., 1996. Structure, volcanic setting, hydrothermal alteration
and genesis of the Thalanga massive sulfide deposit:
Unpublished PhD thesis, University of Tasmania, 404 p.
Hill, A. P., and Orth, K., 1994. Textures and origin of carbonate
associated with some Australian VHMS deposits [abs.]:
AMF Symposium: Australian Research on Ore Genesis,
Adelaide, Australia, 1994.
Hill, A. P., and Orth, K., 1995. Textures and origin of carbonate
associated with the volcanic-hosted massive sulfide deposit at
Rosebery, Tasmania: AMIRA/ARC Project, Centre for Ore
Deposit and Exploration Studies, University of Tasmania,
Unpublished Report 1, November 1995, p. 129-41.
Hodges, D. J., and Manojlovic, P. M., 1993. Application of
lithogeochemistry to exploration for deep VMS deposits
in high-grade metamorphic rocks, Snow Lake, Manitoba:
Journal of Geochemical Exploration, v. 48, p. 201-24.
Hoefs, J., 1973. Stable isotope geochemistry: Berlin, Germany,
Springer-Verlag, 140 p.
Honnorez, J., 1969. La formation actuelle d'un gisement sous-
marin de sulfures fumerolliens a Vulcano (mer tyrrhenienne)
- Partie 1, les Mineraux sulfures des tufs immerges a faible
profondeur, Translated Title: Present-day formation of
a submarine deposit of fumarolic sulfides on Vulcano
(Tyrrhenian sea) - Part 1, sulfide minerals of submerged tuffs
at shallow depths: Mineralium Deposita, v. 4, p. 114-31.
Honnorez, J., 1978. Generation of phillipsites by palagonitization
of basaltic glass in seawater and the origin of K-rich deep-
sea sediments, in Sand, L. B., and Mumpton, E A., eds.,
Natural zeolites: occurrence, properties and use: New York,
USA, Pergamon Press, p. 245-58.
REFERENCES | 2 5 9
Honnorez, J., Bohlke, J. K., and Honnorez-Guerstein, B. M.,
1979. Petrographical and geochemical study of the low-
temperature submarine alteration of basalt from Hole 396B,
Leg 46: Proceedings of the Ocean Drilling Program, Initial
Reports, v. 46, p. 299-318.
Horikoshi, E., 1969. Volcanic activity related to the formation of
the Kuroko-type deposits in the Kosaka district, Japan:
Mineralium Deposita, v. 4, p. 321-45.
Hudak, G. J., Morton, R. L., Peterson, D. M., and Franklin, J. M.,
2000. The Sturgeon Lake caldera complex, northwestern
Ontario - volcanological evolution of an Archaean shallow
water VHMS belt [abs.], in Gemmell, J. B., and Pongratz,
J., eds., Volcanic environments and massive sulfide deposits,
Hobart, Australia, November 16-19, 2000, CODES Special
Publication, v. 3, p. 89-91.
Hughes, C. J., 1973. Spilites, keratophyres, and the igneous
spectrum: Geological Magazine, v. 6, p. 513-27.
Humphris, S. E., and Thompson, G., 1978. Hydrothermal alteration
of oceanic basalts by seawater: Geochimica et Cosmochimica
Acta, v. 42, p. 107-25.
Hunns, S. R., and McPhie, J., 1999. Pumiceous peperite in
a submarine volcanic succession at Mount Chalmers,
Queensland, Australia: Journal of Volcanology and
Geothermal Research, v. 88, p. 239-54.
Huston, D. L., 1993. The effect of alteration and metamorphism on
wall rocks to the Balcooma and Dry River South volcanic-
hosted massive sulfide deposits, Queensland, Australia:
Journal of Geochemical Exploration, v. 48, p. 277307.
Huston, D. L., 1999. Stable isotopes and their significance for
understanding the genesis of volcanic hosted massive sulfide
deposits a review: Reviews in Economic Geology, v. 8,
p. 157-79.
Huston, D. L., and Kamprad, J., 2000. The Western Tharsis deposit,
a high sulfidation Cu-Au deposit in the Mt Lyell field, of
possible Ordovician age: AGSO Research Newsletter, v. 32,
p. 2-6.
Huston, D. L., and Kamprad, J., 2001. Zonation of alteration facies
at Western Tharsis implications for the genesis of Cu-
Au deposits, Mt Lyell field, western Tasmania: Economic
Geology, v. 96, p. 1123-32.
Huston, D. L., and Large, R. R., 1987. Genetic and exploration
significance of the zinc ratio (100 Zn/(Zn + Pb)) in massive
sulfide systems: Economic Geology, v. 82, p. 152139.
Huston, D. L., and Large, R. R., 1989. A chemical model for
the concentration of gold in volcanogenic massive sulfide
deposits: Ore Geology Reviews, v. 4, p. 171200.
Huston, D. L., Taylor, T, Fabray, J., and Patterson, D. J., 1992.
A comparison of the geology and mineralization of the
Balcooma and Dry River South volcanogenic massive sulfide
deposits, northern Queensland: Economic Geology, v. 87,
p. 785-811.
Huston, D. L., Kamprad, J., and Brauhart, C, 1999. Definition of
high-temperature alteration zones with PIMA - an example
from the Panorama VHMS district, central Pilbara Craton:
AGSO Research Newsletter, v. 30, p. 10-12.
Iijima, A., 1974. Clay and zeolitic alteration zones surrounding
Kuroko deposits in the Hokuroku District, Northern Akita,
as submarine hydrothermal-diagenetic alteration products:
Mining Geology Special Issue, v. 6, p. 26789.
Iijima, A., 1978. Geological occurrences of zeolites in marine
environments, in Sand, L. B., and Mumpton, F. A., eds.,
Natural zeolites occurrence, properties, use: Oxford, UK,
Peregamon Press, p. 17598.
Iijima, A., and Hay, R. L., 1968. Analcime composition in tuffs
of the Green River Formation of Wyoming: American
Mineralogist, v. 53, p. 184-200.
Iijima, A., and Utada, M., 1971. Present-day zeolitic diagenesis of
2 6 0 | REFERENCES
the Neogene geosynclinal deposits in the Niigata oil field,
Japan: American Chemical Society Advances in Chemistry
Series, v. 101, p. 342-9.
Iijima, A., and Utada, M., 1972. A critical review on the occurrence
of zeolites in sedimentary rocks in Japan: Japan Journal of
Geology and Geography, v. 42, p. 61-84.
Ineson, P. R., 1989. Introduction to practical ore microscopy: New
York, USA, John Wiley and Sons, 181 p.
Ishikawa, Y., 1983. Volcanic activities and mineralisation of the
Fukazawa-Ezuri Kuroko deposits area: Akita University
Mining College Scientific and Technological Report, no. 4,
p. 23-32.
Ishikawa, Y., Sawaguchi, T., Iwaya, S., and Horiuchi, M., 1976.
Delineation of prospecting targets for Kuroko deposits based
on modes of volcanism of underlying dacite and alteration
halos: Mining Geology, v. 26, p. 105-17.
Ito,T., Matsumoto, R., Kano, K., and Sakuyama, M., 1984. Welding
of acidic tuff and altered rhyolite fragments caused by the
intrusion of magma - examples in the Neogene Shirahama
Group in the southern part of Izu Peninsula: Journal of the
Geological Society of Japan, v. 90, p. 191-205.
Jack, D. J., 1989. Hellyer host rock alteration: Unpublished MSc
thesis, University of Tasmania, 182 p.
Jago, J. B., Reid, K. O., Quilty, P. G., Green, G. R., and Daily, B.,
1972. Fossiliferous Cambrian limestone from within the Mt
Read Volcanics, Mt Lyell mine area, Tasmania: Journal of
Geological Society of Australia, v. 19, p. 37982.
Jakobsson, S. P., 1972. On the consolidation and palagonitization
of the tephra of the Surtsey volcanic island, Iceland: Surtsey
Research Progress Report, v. 6, p. 121-8.
Jakobsson, S. P., 1978. Environmental factors controlling the
palagonitization of the Surtsey tephra, Iceland: Geological
Society of Denmark Bulletin, v. 27, p. 95-105.
Jakobsson, S. P., and Moore, J. G., 1986. Hydrothermal minerals
and alteration rates at Surtsey Volcano, Iceland: Geological
Society of America Bulletin, v. 97, p. 648-59.
Jenner, G. A., 1998. Trace element geochemistry of volcanic rocks
a practical guide to geochemical nomenclature and analytical
geochemistry, Centre for Ore Deposit Research, University
of Tasmania, Master of Economic Geology course work
manual, v. 17, p. 1148.
Jensen, M. L., and Bateman, A. M., 1981. Economic mineral
deposits: New York, USA, John Wiley and Sons, 593 p.
Jezek, P. A., and Noble, D. C, 1978. Natural hydration and ion
exchange of obsidian: American Mineralogist, v. 63, p. 266
73.
Jiang, S., Palmer, M. R., Li, Y, and Xue, C, 1996. Ba-rich micas
from the Yindongzi-Daxigou Pb-Zn-Ag and Fe deposits,
Qinling, northwestern China: Mineralogical Magazine,
v. 60, p. 433-45.
Jolly, W. T., and Smith, R. E., 1972. Degradation and metamorphic
differentiation of the Keweenawan tholeiitic lavas of northern
Michigan, USA: Journal of Petrology, v. 13, p. 273-309.
Jones, A. T, 1993. The geology, geochemistry and structure of the
Mount Darwin-South Darwin Peak area, western Tasmania:
, Unpublished Honours thesis, University of Tasmania,
106 p.
Jones, A. T, 1999. Volcanology and geochemistry of the Cambrian
Mount Read Volcanics in the Basin Lake area, western
Tasmania: Unpublished MSc thesis, University of Tasmania,
190 p.
Jones, S., Herrmann, W, and Gemmell, J. B., in prep. SWIR
spectral characteristics of the HW horizon: implications for
exploration at Myra Falls VHMS camp, Vancouver Island,
British Columbia, Canada: Economic Geology.
Kajiwara, Y, 1973. Significance of cyclic seawater as a possible
determinant of rock alteration facies in the earth's crust:
Geochemical Journal, v. 7, p. 23-36.
Kano, K-L, 1989. Interactions between andesitic magma and
poorly consolidated sediments examples in the Neogene
Shirahama Group, South Izu, Japan: Journal of Volcanology
and Geothermal Research, v. 37, p. 59-75.
Karkhanis, S. N., Bancroft, G. M., Fyfe, W S., and Brown, J. D.,
1980. Leaching behaviour of rhyolite glass: Nature, v. 284,
p. 435.
Kastner, M., 1982. Evidence for two distinct hydrothermal systems
in the Guaymas Basin: Proceedings of the Ocean Drilling
Program, Initial Reports, v. 64, p. 1143-57.
Kastner, M., and Gieskes, J. M., 1976. Interstitial water profiles and
sites of diagenetic reactions, Leg 35, Bellingshausen abyssal
plain: Earth and Planetary Science Letters, v. 33, p. 11-20.
Keating, G. N., and Geissman, J. W, 1998. Cooling history of
shallow-level mafic intrusions and host tuffs, Paiute Ridge,
Nevada field paleomagnetic, and thermal modelling
results [abs.]: Geological Society of America Abstracts with
Programs, v. 30, p. 107.
Keene, J. B., Clague, A., and Nishimori, R. K., 1976. Experimental
hydrothermal alteration of tholeiitic basalt - resultant
mineralogy and textures: Journal of Sedimentary Research,
v. 46, p. 647-53.
Kelley, D. S., and Gillis, K. M., 1993. Fluid evolution in submarine
magma-hydrothermal systems at the Mid-Atlantic Ridge:
Journal of Geophysical Research, v. 98, p. 19579-96.
Kelley, D. S., Robinson, P. T, and Malpas, J. G., 1992. Processes of
brine generation and circulation in the oceanic crust - fluid
inclusion evidence from the Troodos Ophiolite, Cyprus:
Journal of Geophysical Research, v. 97, p. 9307-22.
Kennedy, G. C, 1950. A portion of the system silica-water:
Economic Geology, v. 45, p. 62953.
Kerr, D. J., and Gibson, H. L., 1993. A comparison of the Home
volcanogenic massive sulfide deposit and intracauldron
deposits of the Mine Sequence, Noranda, Quebec: Economic
Geology, v. 88, p. 1419-42.
Khin Zaw, and Large, R. R., 1992. The precious metal-rich, South
Hercules mineralisation, western Tasmania - a possible
subseafloor replacement volcanic-hosted massive sulfide
deposit: Economic Geology, v. 87, p. 93152.
Khin Zaw, Gemmell, J. B., Large, R. R., Mernagh, T. P., and Ryan,
C. G., 1996. Evolution and source of ore fluids in the
stringer system, Hellyer VHMS deposit, Tasmania, Australia
evidence from fluid inclusion microthermometry and
geochemistry: Ore Geology Reviews, v. 10, p. 251-78.
Khin Zaw, Huston, D. L., and Large, R. R., 1999. A chemical model
for the Devonian remobilization process in the Cambrian
volcanic-hosted massive sulfide Rosebery deposit, western
Tasmania: Economic Geology, v. 94, p. 529-46.
Kirste, D., de Caritat, P., and Dann, R., 2003. The application of the
stable isotopes of sulfur and oxygen in groundwater sulfate
to mineral exploration in the Broken Hill region of Australia:
Journal of Geochemical Exploration, v. 78-79, p. 81-4.
Klein, G. D. V, and Lee, Y. I., 1984. A preliminary assessment of
geodynamic controls on depositional systems and sandstone
diagenesis in back-arc basins, western Pacific Ocean:
Tectonophysics, v. 102, p. 11952.
Knuckey, M. J., Comba, C. D. A., and Riverin, G., 1982. Structure,
metal zoning and alteration at the Millenbach deposit,
Noranda, Quebec, in Hutchinson, R. W, Spence, C. D.,
and Franklin, J. M., eds., Precambrian sulfide deposits:
Geological Association of Canada Special Paper, v. 25,
p. 255-317.
Kokelaar, B. P., 1982. Fluidization of wet sediments during the
emplacement and cooling of various igneous bodies: Journal
of the Geological Society of London, v. 139, p. 21-33.
Komyshan, P., 1986. Geology of the Hellyer-Mt Charter area, in
Large, R. R., ed., The Mount Read Volcanics and associated
ore deposits: Burnie, Australia, November 2021, 1986,
Geological Society of Australia, Tasmanian Division,
Proceedings, p. 53-5.
Kristmannsdottir, H., 1976. Types ofclay minerals in hydro thermally
altered basaltic rocks, Reykjanes, Iceland: Jokull, v. 26,
p. 30-9.
Kuniyoshi, S., and Liou, J. G., 1976. Burial metamorphism of the
Karmutsen volcanic rocks, northeastern Vancouver Island,
British Columbia: American Journal of Science, v. 276,
p. 1096-119.
Kutina, J., Park, C. E, Jr., and Smirnov, V. I., 1965. On the
definition of zoning and on the relation between zoning
and paragenesis: Symposium - Problems of post-magmatic
ore deposition with special reference to the geochemistry of
ore veins, Prague, Czech Republic, 1965, Proceedings, v. 2,
p. 589-95.
Lacroix, A., 1936. Le volcan actif de l'ile de la Reunion et ses
produits: Paris, France, Gauthier-Villars, 297 p.
Lagerbald, B., and Gorbatschev, R., 1985. Hydrothermal alteration
as a control of regional geochemistry and ore formation in
the central Baltic Shield: Geologische Rundschau, v. 74,
p. 33-49.
Landis, C. A., and Coombs, D. S., 1967. Metamorphic belts and
orogenesis in southern New Zealand: Tectonophysics, v. 4,
p. 501-18.
Large, R. R., 1977. Chemical evolution and zonation of massive
sulfide deposits in volcanic terrains: Economic Geology,
v. 72, p. 549-72.
Large, R. R., 1992. Australian volcanic-hosted massive sulfide
deposits - features, styles, and genetic models: Economic
Geology, v. 87, p. 471-510.
Large, R. R., 1997. Variability is the key to understanding the
genesis of Australian Palaeozoic volcanic hosted massive
sulfide deposits [abs.], in Barriga, F. J. A. S., ed. SEG Neves
Corvo field conference: Lisbon, Portugal, May 11-14, 1997,
SEG Abstracts and Program, p. 54.
Large, R. R., 2000. The significance of hybrid deposits in the
continuum of VHMS-epithermal and porphyry systems - an
example from the Mt Read Volcanics [abs.], in Gemmell, J.
B., and Pongratz, J., eds., Volcanic environments and massive
sulfide deposits: Hobart, Australia, November 16-19, 2000,
CODES Special Publication, v. 3, p. 113.
Large, R. R., 2004. Development of an alteration vector diagram for
use in the exploration for BHT systems [abs.], in McPhie,
J., and McGoldrick, P. J., eds., Dynamic earth past,
present and future: Hobart, Australia, February 8-13, 2004,
Geological Society of Australia Abstracts, v. 73, p. 91.
Large, R. R., and Allen, R. L., 1997. Preliminary report on the
Rosebery lithochemical halo study: AMIRA/ARC Project
P439, Centre for Ore Deposit and Exploration Studies,
University of Tasmania, Unpublished Report 4, May 1997,
p. 259-330.
Large, R. R., and Both, R. A., 1980. The volcanogenic sulfide ores at
Mount Chalmers, eastern Queensland: Economic Geology,
v. 75, p. 992-1009.
Large, R. R., Huston, D. L., McGoldrick, P. J., Ruxton, P. A., and
McArthur, G., 1989. Gold distribution and genesis in
Australian volcanogenic massive sulfide deposits and their
significance for gold transport models: Economic Geology
Monographs, v. 6, p. 520-36.
Large, R. R., Doyle, M.G., Raymond, O. L, Cooke, D. R., Jones,
A. T, and Heasman, L., 1996. Evaluation of the role of
Cambrian granites in the genesis of world-class VHMS
deposits in Tasmania: Ore Geology Reviews, v. 10, p. 215-
30.
Large, R. R., Bull, S. W., and McGoldrick, P. J., 2000.
REFERENCES | 2 6 1
Lithogeochemical halos and geochemical vectors to
stratiform sediment hosted Zn-Pb-Ag deposits - Part 2,
HYC deposit, McArthur River, Northern Territory: Journal
of Geochemical Exploration, v. 68, p. 105-26.
Large, R. R., Gemmell, J. B., and Paulick, H., 2001a. The
Alteration box plot - a simple approach to understanding
the relationship between alteration mineralogy and
lithogeochemistry associated with volcanic-hosted massive
sulfide deposits: Economic Geology, v. 96, p. 95771.
Large, R. R., Allen, R. L., Blake, M. D., and Herrmann, W., 2001b.
Hydrothermal alteration and volatile element halos for the
Rosebery K lens volcanic-hosted massive sulfide deposit,
western Tasmania: Economic Geology, v. 96, p. 1055-72.
Large, R. R., McPhie, J., Gemmell, J. B., Herrmann, W., and
Davidson, G. J., 2001c. The spectrum of ore deposit
types, volcanic environments, alteration halos, and related
exploration vectors in submarine volcanic successions
some examples from Australia: Economic Geology, v. 96,
p. 913-38.
Larsen, G., and Chilingar, G. V, 1979. Introduction - diagenesis
of sediments and rocks, in Larsen, G., and Chilingar, G. V,
eds., Diagenesis in sediments and sedimentary rocks, New
York, USA, Elsevier Scientific Publishing Co., v. 1, p. 1-29.
Le Maitre, R. W., 1981. GENMIX - a generalised petrological
mixing model program: Computers and Geosciences, v. 7,
p. 229-47.
Leaman, D. E., and Richardson, R. G., 1989. The granites of west
and northwest Tasmania a geophysical interpretation:
Tasmania Department of Mines, Geological Survey Bulletin,
v. 66, 146 p.
Lee, Y. I., and Klein, G. D., 1986. Diagenesis of sandstones in the
back-arc basins of the western Pacific Ocean: Sedimentology,
v. 33, p. 651-75.
Lees, T. C, 1987. Geology and mineralisation of the Rosebery-
Hercules area, Tasmania: Unpublished MSc thesis, University
of Tasmania, 160 p.
Lees, T. C, Khin Zaw, Large, R. R., and Huston, D. L, 1990.
Rosebery and Hercules copper-lead-zinc deposits, in Hughes,
F. E., ed., Geology of the mineral deposits of Australia and
Papua New Guinea: Australasian Institute of Mining and
Metallurgy Monograph, v. 14, p. 1241-47.
Leistel, J. M., Marcoux, E., Thieblemont, D., Quesada, C, Sanchez,
A., Almodovar, G. R., Pascual, E., and Saez, R., 1998. The
volcanic-hosted massive sulfide deposits of the Iberian pyrite
belt: Mineralium Deposita, v. 33, p. 230.
Lentz, D. R., and Goodfellow, W. D., 1996. Intense silicification
of footwall sedimentary rocks in the stockwork alteration
zone beneath the Brunswick No. 12 massive sulfide deposit,
Bathurst, New Brunswick: Canadian Journal of Earth
Sciences, v. 33, p. 284-302.
Lentz, D. R., Hall, D. C, and Hoy, L. D., 1997. Chemostratigraphic,
alteration, and oxygen isotopic trends in a profile through
the stratigraphic sequence hosting the Health Steele B zone
massive sulfide deposit, New Brunswick: The Canadian
Mineralogist, v. 35, p. 841-74.
Lesher, C. M., Goodwin, A. M., Campbell, I. H., and Gorton, M.
P., 1986. Trace-element geochemistry of ore-associated and
barren, felsic metavolcanic rocks in the Superior province,
Canada: Canadian Journal of Earth Sciences, v. 23, p. 222-
37.
Levi, B., 1970. Burial metamorphic episodes in the Andean
geosyncline, Central Chile: Geologishe Rundschau, v. 59, p.
994-1013.
Lindgren, W., 1933. Mineral deposits: New York, USA, McGraw-
Hill, 936 p.
Lipman, P. W., 1965. Chemical comparison of glassy and crystalline
volcanic rocks: United States Geological Survey Bulletin,
2 6 2 | REFERENCES
1201-D, 24 p.
Lipman, P. W., Christiansen, R. L., and van Alstine, R. E., 1969.
Retention of alkalis by calc-alkalic rhyolites during
crystallisation and hydration: American Mineralogist, v. 54,
p. 286-91.
Lofgren, G., 1970. Experimental devitrification rate of rhyolite glass:
Geological Society of America Bulletin, v. 81, p. 553-60.
Lofgren, G., 1971a. Experimentally produced devitrification textures
in natural rhyolite glass: Geological Society of America
Bulletin, v. 82, p. 111-24.
Lofgren, G., 1971b. Spherulitic textures in glassy and crystalline
rocks: Journal of Geophysical Research, v. 76, p. 5635-48.
Lowell, J. D., and Guilbert, J. M., 1970. Lateral and vertical
alteration-mineralisation zoning in porphyry ore deposits:
Economic Geology, v. 65, p. 373408.
Lydon, J. W., 1984. Volcanogenic massive sulfide deposits - Part 1, a
descriptive model: Geoscience Canada, v. 11, p. 195-202.
Lydon, J. W., 1988. Volcanogenic massive sulfide deposits Part 2,
genetic models: in Sheahan, P. A., and Cherry, M. E., eds.,
Ore deposit models: Geoscience Canada Reprint Series, v. 3,
p. 145-155.
Lydon, J. W, 1996. Characteristics of volcanogenic massive
sulfide deposits interpretations in terms of hydrothermal
convection systems and magmatic hydrothermal systems:
Boletin Geologico y Minero, v. 107, p. 215-64.
Lydon, J. W., 1997. Sedex deposits and global evolution [abs.]:
Geological Society of America Abstracts with Programs,
v. 29, p. 17-18.
Lydon, J. W., and Galley, A., 1986. The chemical and mineralogical
zonation of the Mathiati alteration pipe, Cyprus, and its
genetic significance, in Gallagher, M. J., et al., ed., Metallurgy
of basic and ultra-basic rocks: London, UK, Institution of
Mining and Metallurgy, p. 49-68.
Lydon, J. W, and Jamieson, H. E., 1984. The generation of ore-
forming hydrothermal solutions in the Troodos Ophiolite
Complex some hydrodynamic and mineralogical
considerations: Geological Survey of Canada Paper, v. 84-
1A, p. 617-25.
MacGeehan, P. J., 1978. The geochemistry of altered volcanic rocks
at Matagami, Quebec a geothermal model for massive
sulfide genesis: Canadian Journal of Earth Sciences, v. 15,
p. 551-70.
MacLean, W. H., and Barrett, T. J., 1993. Lithogeochemical
techniques using immobile elements: Journal of Geochemical
Exploration, v. 48, p. 109-33.
MacLean, W. H., and Hoy, L. D., 1991. Geochemistry of
hydrothermally altered rocks at the Home Mine, Noranda,
Quebec: Economic Geology, v. 86, p. 50628.
MacLean, W. H., and Kranidiotis, P., 1987. Immobile elements
as monitors of mass transfer in hydrothermal alteration -
Phelps Dodge massive sulfide deposit, Matagami, Quebec:
Economic Geology, v. 82, p. 95162.
Madeisky, H. E., and Stanley, C. R., 1994. Metasomatism in felsic
volcanic hosted massive sulfide footwall alteration zones
- lithogeochemical patterns, fluid-rock reactions, and their
t application to exploration in metamorphosed terranes [abs.]:
Waterloo, Canada, May 1618,1994, Geological Association
of Canada Program with Abstracts, v. 19, p. 70.
Manning, C. E., and Bird, K. K., 1991. Porosity, evolution and fluid
flow in the basalts of the Skaergaard magma-hydrothermal
system, east Greenland: American Journal of Science, v. 291,
p. 201-57.
Manning, C. E., and Bird, D. K., 1995. Porosity, permeability, and
basalt metamorphism: Geological Society of America Special
Paper, v. 296, p. 123-40.
Marsaglia, K. M., and Tazaki, K., 1992. Diagenetic trends in Leg
126 sandstones: Proceedings of the Ocean Drilling Program,
Scientific Results, v. 126, p. 125-38.
Marshall, C. E., 1961. Reactions of feldspars and micas with aqueous
solutions: Economic Geology, v. 56, p. 1330.
Martin, N. K., 2004. Genesis of the Rosebery massive sulfide
deposit, western Tasmanian, Australia: Unpublished PhD
thesis, University of Tasmania, 273 p.
Marumo, K., 1989. Genesis of kaolin minerals and pyrophyllite
in Kuroko deposits of Japan implications for the origins
of the hydrothermal fluids from mineralogical and stable
isotope data: Geochimica et Cosmochimica Acta, v. 53,
p.2915-24.
Mason, B., 1966. Principles of geochemistry: New York, USA, John
Wiley and Sons, 329 p.
McArthur, G. J., 1989. Hellyer, in Burrett, C. E, and Martin, E. L,
eds., Geology and mineral resources of Tasmania: Geological
Society of Australia Special Publication, v. 15, p. 14448.
McArthur, G. J., 1996. Textural evolution of the Hellyer massive
sulfide deposit (3 volumes): Unpublished PhD thesis,
University of Tasmania, 272 p.
McArthur, G. J., and Dronseika, E. V., 1990. Que River and Hellyer
zinc-lead-silver deposits: Australasian Institute of Mining
and Metallurgy, Monograph, v. 14, p. 1229-39.
McArthur, A. N., Cas, R. A. E, and Orton, G. J., 1998. Distribution
and significance of crystalline, perlitic and vesicular
textures in the Ordovician Garth Tuff (Wales): Bulletin of
Volcanology, v. 60, p. 260-85.
McBride, E. E, 1978. Porosity loss in sandstone by ductile grain
deformation during compaction: AAPG (American
Association of Petroleum Geologists) Bulletin, v. 62,
p. 1761.
McConnell, V. S., Shearer, C. K., Eichelberger, J. C, Keskinen, M.
J., Layer, P. W, and Papike, J. J., 1995. Rhyolite intrusions in
the intracaldera Bishop Tuff, Long Valley Caldera, California:
Journal of Volcanology and Geothermal Research, v. 67,
p. 41-60.
McKibben, J. A., 1993. The geology and geochemistry of the North
Pinnacles Ridge, western Tasmania: Unpublished Honours
thesis, University of Tasmania, 120 p.
McLeod, R. L., 1987. Alteration associated with volcanogenic sulfide
ores at Mount Chalmers, Queensland, Australia: Institution
of Mining and Metallurgy, Transactions, Section B: Applied
Earth Science, v. 96, p. 117-27.
McLeod, R. L., and Stanton, R. L., 1984. Phyllosilicates and
associated minerals in some Palaeozoic stratiform sulfide
deposits of southeastern Australia: Economic Geology, v. 79,
p. 1-22.
McMillan, W J., and Panteleyev, A., 1998. Porphyry copper
deposits, in Roberts, R. G., and Sheahan, P. A., eds., Ore
deposit models, Geoscience Canada Reprint Series, v. 3,
p. 45-58.
McNeill, A. W, and Corbett, K. D., 1992. Geology and
mineralisation of the Mt Murchison area (including Map
4): Mt Read Volcanics Project Geological Report, Tasmania
Department of Mines, no. 3, 39 p.
McPhie, J., 1996. Hall Rivulet Canal-Hercules-Mount Read-Read
Hills-Anthony Dam: AMIRA/ARC Project P439, Centre
for Ore Deposit and Exploration Studies, University of
Tasmania, Unpublished Report 2, May 1996, p. 1928.
McPhie, J., and Allen, R. L., 1992. Facies architecture of mineralised
submarine volcanic sequences Cambrian Mount Read
Volcanics, western Tasmania: Economic Geology, v. 87,
p. 587-96.
McPhie, J., and Allen, R. L., 2003. Submarine, silicic, syneruptive
pyroclastic units in the Mount Read Volcanics, western
Tasmania influence of vent setting and proximity on
lithofacies characteristics: AGU Geophysical Monograph,
v. 140, p. 245-58.
McPhie.J., andHunns, S. R., 1995. Secondary welding of submarine,
pumice-lithic breccia at Mount Chalmers, Queensland,
Australia: Bulletin of Volcanology, v. 57, p. 170-8.
McPhie, J., Doyle, M. G., and Allen, R. L., 1993. Volcanic textures
a guide to the interpretation of textures in volcanic rocks:
Hobart, Australia, Centre for Ore Deposit and Exploration
Studies, University of Tasmania, 197 p.
Meinert, L. D., 1983. Variability of skarn deposits guides to
exploration, in Boardman Shelby, J., ed., Revolution in the
earth sciences - advances in the past half-century: Dubuque,
USA, Kendall/Hunt Publishing Co., p. 301-16.
Meinert, L. D., 1993. Skarns and skarn deposits, in Sheahan, P. A.,
and Cherry, M. E., eds., Ore deposit models, 2, Geoscience
Canada Reprint Series, v. 2, p. 11734.
Merino, E., 1975. Diagenesis in Tertiary sandstones from Kettleman
North Dome, California Part I, diagenetic mineralogy:
Journal of Sedimentary Petrology, v. 45, p. 32036.
Merino, E., Ortoleva, P., and Strickholm, P., 1983. Generation
of evenly spaced pressure-solution seams during (late)
diagenesis a kinetic theory: Contributions to Mineralogy
and Petrology, v. 82, p. 360-70.
Merry, N., and Pontual, S., 1999. Rapid alteration mapping using
field portable infrared spectrometers: Bali, Indonesia,
October 10-13, 1999, Proceedings of PACRIM '99 congress
[modified], p. 693-8.
Mevel, C, and Cannat, M., 1991. Lithospheric stretching and
hydrothermal processes in oceanic gabbros from slow-
spreading ridges, in Peters, T. J. (ed.) Ophiolite genesis and
evolution of the oceanic lithosphere, Ministry of Petroleum
and Minerals, Muscat, Sultante of Oman, p. 293-312.
Meyer, C, and Hemley, J. J., 1967. Wall-rock alteration, in Barnes,
H. L., ed., Geochemistry of hydrothermal ore deposits: New
York, USA, Holt, Rinehart and Winston, p. 166-235.
Miller, C. R., 1996. Geological and geochemical aspects of the
Liontown VHMS deposit, northeastern Queensland:
Unpublished MSc thesis, University of Tasmania, 60 p.
Miller, C. R., Halley, S., Green, G. R., and Jones, M. T, 2001.
Discovery of the West 45 volcanic-hosted massive sulfide
deposit using oxygen isotopes and REE geochemistry:
Economic Geology, v. 96, p. 122737.
Miyashiro, A., and Shido, E, 1970. Progressive metamorphism in
zeolite assemblages: Lithos, v. 3, p. 251-60.
Moore, J. G., 1966. Rate of palagonitization of submarine basalt
adjacent to Hawaii: United States Geological Survey
Professional Paper 55OD, D163-71.
Morrow, D. W, and Mcllreath, I. A., 1990. Diagenesis general
introduction, in Mcllreath, I. A., and Morrow, D. W, eds.,
Diagenesis: Newfoundland, Canada, Geological Association
of Canada, Geoscience Canada Reprint Series, v. 4, p. 1-8.
Morton, R. L., and Franklin, J. M., 1987. Two-fold classification
of Archaean volcanic-associated massive sulfide deposits:
Economic Geology, v. 82, p. 105763.
Mottl, M. J., 1983. Metabasalts, axial hot springs, and the structure
of hydrothermal systems at mid-ocean ridges: Geological
Society of America Bulletin, v. 94, p. 161-80.
Mottl, M. J., and Seyfried, W. E., 1977. Experimental basalt-
seawater interaction - rock- vs. seawater-dominated systems
and the origin of submarine hydrothermal deposits [abs.]:
Geological Society of America Abstracts with Programs, v. 9,
p. 1104-5.
Muehlenbachs, K., and Clayton, R. N., 1972. Oxygen isotope
studies of fresh and weathered submarine basalts: Canadian
Journal of Earth Sciences, v. 9, p. 172-84.
Mungall, J. E., and Martin, R. E, 1994. Severe leaching of trachytic
glass without devitrification, Terceira, Azores: Geochimica et
Cosmochimica Acta, v. 58, p. 75-83.
Munha, J., and Kerrich, R., 1980. Seawater basalt interaction
REFERENCES | 2 6 3
in spilites from the Iberian pyrite belt: Contributions to
Mineralogy and Petrology, v. 73, p. 191-200.
Munha, J., Fyfe, W. S., and Kerrich, R., 1980. Adularia the
characteristic mineral of felsic spilites: Contributions to
Mineralogy and Petrology, v. 75, p. 15-19.
Nakajima, T, 1988. Geology of the Hokuroku basin in Kuroko
deposits and geothermal fields in northern Honshu: The
Society of Mining Geologists of Japan, Guidebook 3.
Nehlig, P., 1993. Interactions between magma chambers and
hydrothermal systems - oceanic and ophiolitic constraints:
Journal of Geophysical Research, v. 98, p. 19621-33.
Nesbitt, B. E. 1996. Applications of oxygen and hydrogen isotopes
to exploration for hydrothermal mineralisation: SEG
Newsletter, v. 27, p. 1 and 8-13.
Neuhoff, P. S., Watt, W S., Bird, D. K., and Pedersen, A. K., 1997.
Timing and structural relations of regional zeolite zones in
basalts of the east Greenland continental margin: Geology,
v. 25, p. 803-6.
Nielsen, T E D., Soper, N. J., Brooks, C. K., Faller, A. M., Higgens,
A. C, and Matthews, D. W, 1981. The pre-basaltic
sediments and the lower basalts at Kangerdlugssuaq, east
Greenland their stratigraphy, lithology, palaeomagnetism
and petrology: Meddelelser om Gronland, Geoscience, v. 6,
p. 25.
Nilsson, C. A., 1968. Wall-rock alteration at the Boliden deposit,
Sweden: Economic Geology, v. 63, p. 47294.
Noble, D. C, 1967. Sodium, potassium and ferrous iron contents of
some secondarily hydrated natural silicic glasses: American
Mineralogist, v. 52, p. 280-6.
Noh, J. H., and Boles, J. R., 1989. Diagenetic alteration of perlite
in the Guryongpo area, Republic of Korea: Clays and Clay
Minerals, v. 37, p. 47-58.
Norton, D. L., 1984. Theory of hydrothermal systems: Annual
Review Earth Planetary Science, v. 12, p. 15577.
Norton, D. L., and Taylor, H. P., Jr., 1979. Quantitative simulation
of the hydrothermal systems of crystallizing magmas on
the basis of transport theory and oxygen isotope data - an
analysis of the Skaergaard intrusion: Journal of Petrology,
v. 20, p. 421-86.
Norton, D. L., Taylor, H. P., Jr., and Bird, D. K., 1984. The
geometry and high-temperature brittle deformation of the
Skaergaard intrusion: Journal of Geophysical Research,
v. 89, p. 10178-92.
Offler, R., and Whitford, D. J., 1992. Wall-rock alteration and
metamorphism of a volcanogenic massive sulfide deposit at
Que River, Tasmania petrology and mineralogy: Economic
Geology, v. 87, p. 686-705.
Oftedahl, C, 1958. A theory of exhalative-sedimentary ores:
Geologiska Foereningen i Stockholm Foerhandlingar, v. 80,
p. 1-19.
Ogihara, S., 1996. Diagenetic transformation of clinoptilolite to
analcime in silicic tuffs of Hokkaido, Japan: Mineralium
Deposita, v. 31, p. 548-53.
Ohmoto, H., 1978. Submarine caldera a key for the formation
of massive sulfide deposits: Mining Geology Special Issue,
v. 28, p. 219-31.
Ohmoto, H., 1986. Stable isotope geochemistry of ore deposits:
Reviews in Mineralogy, v. 16, p. 491-559.
Ohmoto, H., 1996. Formation of volcanogenic massive sulfide
deposits the Kuroko perspective: Ore Geology Reviews,
v. 10, p. 135-77.
Ohmoto, H., and Goldhaber, M. J., 1997. Sulfur and carbon isotopes,
in Barnes, H. L., ed., Geochemistry of hydrothermal ore
deposits, 3
rd
edition: New York, USA, John Wiley and Sons,
p. 517-612.
Ohmoto, H., and Rye, R. O., 1974. Hydrogen and oxygen isotopic
compositions of fluid inclusions an introduction and
2 6 4 | REFERENCES
summary of new findings: Economic Geology Monographs,
v. 5, p. 1-8.
Ohmoto, H., and Skinner, B. J., 1983. The Kuroko and related
volcanogenic massive sulfide deposits introduction and
summary of new findings: Economic Geology Monographs,
v. 5, p. 1-8.
Ohmoto, H., and Takahashi, T., 1983. Submarine calderas and
Kuroko genesis - Part III: Economic Geology Monographs,
v. 5, p. 39-54.
Ohmoto, H., Mizukami, M., Drummond, S. E., Eldridge, C. S.,
Pisutha, A. V., and Lenagh, T. C., 1983. Chemical processes
of Kuroko formation: Economic Geology Monographs, v. 5,
p. 570-604.
O'Keefe, J. A., 1984. Natural glass: Journal of Non-Crystalline
Solids, v. 67, p. 1-17.
O'Neil, J. R., and Silberman, M. L., 1974. Stable isotope relations
in epithermal Au-Ag deposits: Economic Geology, v. 69,
p. 902-9.
Palmasson, G., Arnorsson, S., Fridleifsson, I. B., Kristmannsdottir,
H., Saemundson, K., Stefansson, V., Steingrimsson, B., and
Tomasson, J., 1979. The Iceland crust evidence from drill
hole data on structure and processes, in Talwani, M., et al.,
eds., Deep drilling results in the Atlantic Ocean: oceanic
crust, AGU Maurice Ewing Series, v. 2, p. 43-65.
Paradis, S., Taylor, B. E., Watkinson, D. H., and Jonasson, I. R.,
1993. Oxygen isotope zonation and alteration in the northern
Noranda District, Quebec - evidence for hydrothermal fluid
flow: Economic Geology, v. 88, p. 1512-25.
Paradis, S., Bailes, A. H., and Galley, A. G., in press. Regional-
scale synvolcanic alteration zones associated with Snow
Lake volcanogenic massive sulfide deposits, Flin Flon belt,
Manitoba, Canada - Part II, mineralogical and chemical
characteristics: Mineralium Deposita.
Passaglia, E., Artioli, G., Gualtieri, A., and Carnevali, R., 1995.
Diagenetic mordenite from Ponza, Italy: European Journal
of Mineralogy, v. 7, p. 429-38.
Paulick, H., 1999. The Thalanga sequence - facies architecture,
geochemistry, alteration and metamorphism of felsic
volcanics hosting the Thalanga massive sulfide deposit (Early
Ordovician, northern Queensland, Australia): Unpublished
PhD thesis, University of Tasmania, 240 p.
Paulick, H., and McPhie, J., 1999. Facies architecture of the felsic
lava-dominated host sequence to the Thalanga massive
sulfide deposit, Lower Ordovician, northern Queensland:
Australian Journal of Earth Sciences, v. 46, p. 391-405.
Paulick, H., Herrmann, W., and Gemmell, J. B., 2001. Alteration of
felsic volcanics hosting the Thalanga massive sulfide deposit
(northern Queensland, Australia) and geochemical proximity
indicators to ore: Economic Geology, v. 96, p. 1175-200.
Payne, B., 1991. Geophysical interpretation of the Mt Sedgewick-
Red Hills area, western Tasmania: Unpublished Honours
thesis, University of Tasmania, 107 p.
Peacock, M. A., 1926. The petrology of Iceland - Part I, the basic
tuffs: Royal Society of Edinburgh Transactions, v. 55, p. 53
76.
Pearce, J. A., and Cann, J. R., 1973. Tectonic setting of basic volcanic
rocks determined using trace element analyses: Earth and
Planetary Science Letters, v. 19, p. 290-300.
Pemberton, J., and Corbett, K. D., 1992. Stratigraphic-facies
associations and their relationship to mineralisation in
the Mount Read Volcanics: Bulletin Geological Survey
Tasmania, v. 70, p. 16776.
Pemberton, J., Vicary, M. J., and Corbett, K. D., 1991. Geology of
the Cradle Mountain Link Road-Mt Tor area (maps 7 and
8): Mt Read Volcanics Project Geological Report, Tasmania
Department of Mines, v. 4, 74 p.
Penney, A., 1998. Geology and genesis of the massive pyrite and
carbonate, Henty Gold Mine, Tasmania: Unpublished
Honours thesis, University of Tasmania, 111 p.
Peter, J. M., and Goodfellow, W. D., 1996. Mineralogy, bulk and
rare earth element geochemistry of massive sulfide-associated
hydrothermal sediments of the Brunswick horizon, Bathurst
mining camp, New Brunswick: Canadian Journal of Earth
Sciences, v. 33, p. 25283.
Peterson, D. W., 1979. Significance of the flattening of pumice
fragments in ash-flow tuffs, in Chapin, C. E., and Elston,
W. E., eds., Ash-flow tuffs, Geological Society of America
Special Paper, no. 180, p. 195-204.
Pettijohn, F. J., 1957. Sedimentary rocks: New York, USA, Harper
and Row, 718 p.
Phillips, W. J., 1973. Mechanical effects of retrograde boiling and its
probable importance in the formation of some porphyry ore
deposits: Institute Mining Metallurgy Transcripts, Section B,
v. 82, p. B90-8.
Pichler, H., 1965. Acid hyaloclastites: Bulletin of Volcanology, v. 28,
p. 293-310.
Pichler, H., 1981. Italienische Vulkan-Gebiete III: Lipari, Vulcano,
Stromboli, Tyrrhenisches Meer: Translated Title: Italian
volcanic zones III: Lipari, Volcano, Stromboli, Tyrrhenisches
Sea: Sammlung Geologischer Fuhrer, v. 69, 270 p.
Pichler, T, Giggenbach, W. E, Mclnnes, B. I. A., Buhl, D., and
Duck, B., 1999. Fe-sulfide formation due to seawater-gas-
sediment interaction in a shallow-water hydrothermal system
at Lihir Island, Papua New Guinea: Economic Geology,
v. 94, p. 281-8.
Pisutha-Arnond, V., and Ohmoto, H., 1983. Thermal history, and
chemical and isotopic compositions of the ore-forming
fluids responsible for the Kuroko massive sulfide deposits
in the Hokuroku District of Japan: Economic Geology
Monographs, v. 5, p. 52358.
Plimer, I. R., and de Carvalho, D., 1982. The geochemistry of
hydrothermal alteration at the Salgadinho copper deposit,
Portugal: Mineralium Deposita, v. 17, p. 193-211.
Polya, D. A., 1981. The geology of the Murchison Gorge:
Unpublished Honours thesis, University of Tasmania,
170 p.
Polya, D. A., Solomon, M., Eastoe, C. J., and Walshe, J. L, 1986.
The Murchison Gorge, Tasmania a possible cross-section
through a Cambrian massive sulfide system: Economic
Geology, v. 81, p. 1341-55.
Pontual, S., Merry, N., and Gamson, P., 1997. Spectral interpretation
field manual, G-Mex, volume 1: Melbourne, Ausspec
International Pty. Ltd., 169 p.
Potts, P. J., 1987. A handbook of silicate rock analysis: London, UK,
Blackie and Son Ltd., 622 p.
Puffer, J. H., and Benimoff, A. I., 1997. Fractionation, hydrothermal
alteration, and wall-rock contamination of an Early Jurassic
diabase intrusion, Laurel Hill, New Jersey: Journal of
Geology, v. 105, p. 99-110.
Ramdohr, P., 1980. The ore minerals and their intergrowths: Oxford,
UK, Pergamon Press, 1205 p.
Ratterman, N. G., and Surdam, R. C, 1981. Zeolite mineral
reactions in a tuff in the Laney Member of the Green River
Formation, Wyoming: Clays and Clay Minerals, v. 29,
p. 365-77.
Relvas, J. M. R. S., Barriga, E J. A. S., Bernardino, F. B. C. P.,
Oliveira, V. M. S., and Matos, J. X., 1994. Ore zone
hydrothermal alteration in drill hole IGM-LS1, at Lagoa
Salgada, Grandola, Portugal a first report on pyrophyllite
in a central stockwork, Boletin Sociedad Espanola de
Mineralogia, v. 17, p. 157-8.
Relvas, J. M. R. S., Barriga, F. J. A. S., Ferreira, A., Noiva, P. C,
and Carvalho, P., 1997. Footwall alteration and stringer ores,
Corvo ore body, Neves Corvo, Portugal [abs.], in Barriga,
F. J. A. S., ed. SEG Neves Corvo field conference: Lisbon,
Portugal, May 11-14, 1997, SEG Abstracts and Program,
p. 93.
Relvas, J. M. R. S., Barriga, E J. A. S., Ferreira, A., and Noiva, P. C.,
2000. Ore geology, hydrothermal alteration and replacement
ore-forming mechanisms in the Corvo ore body, Neves
Corvo, Portugal [abs.], in Gemmell, J. B., and Pongratz, J.,
eds., Volcanic environments and massive sulfide deposits,
Hobart, Australia, November 16-19, 2000, CODES Special
Publication, v. 3, p. 167-8.
Reyes, A. R., 1990. Petrology of Philippine geothermal systems and
the application of alteration mineralogy to their assessment:
Journal of Volcanology and Geothermal Research, v. 43,
p.279-309.
Richardson, C. J., Cann, J. R., Richards, H. G., and Cowan, J. G.,
1987. Metal-depleted root zones of theTroodos ore-forming
hydrothermal systems, Cyprus: Earth and Planetary Science
Letters, v. 84, p. 243-53.
Riech, V., 1979. Diagenesis of silica, zeolites and phyllosilicates
at sites 397 and 398: Proceedings of the Ocean Drilling
Program, Initial Reports, v. 47, p. 741-60.
Rittmann, A., 1962. Volcanoes and their activity: New York, USA,
John Wiley and Sons, 305 p.
Riverin, G., 1977. Wall-rock alteration at the Millenbach Mine,
Noranda, Quebec: Unpublished PhD thesis, Queen's
University.
Riverin, G., and Hodgson, C. J., 1980. Wall-rock alteration at the
Millenbach Cu-Zn mine, Noranda, Quebec: Economic
Geology, v. 75, p. 424-44.
Roberts, R. G., and Reardon, E. J., 1978. Alteration and ore-forming
processes at Mattagami Lake Mine, Quebec: Canadian
Journal of Earth Sciences, v. 15, p. 1-21.
Robinson, P., 2001. The place of XRF and ICP-MS in the analysis of
silicate rocks, sulfide ores, ironstones and carbonates [abs.]:
2
nd
international workshop, Siberian Geoanalytical Seminar,
Irkutsk, Russia, July 24-26, 2001, Abstracts volume, p. 56.
Rollinson, H. R., 1993. Using geochemical data - evaluation,
presentation, interpretation: Harlow, UK, Longman
Scientific and Technical, 352 p.
Rona, P. A., and Scott, S. D., 1993. A special issue on seafloor
hydrothermal mineralisation - new perspectives: Economic
Geology, v. 88, p. 1933-2245.
Rona, P. A., Hannington, M. D., Raman, C. V, Thompson, G.,
Tivey, M. K., Humphris, S. E., Lalou, C, and Petersen, S.,
1993. Active and relict seafloor hydrothermal mineralisation
at the TAG hydrothermal field, Mid-Atlantic Ridge:
Economic Geology, v. 88, p. 1987-2013.
Rose, A. W, 1970. Zonal relations of wall-rock alteration and
sulfide distribution at porphyry copper deposits: Economic
Geology, v. 65, p. 920-36.
Rose, A. W, and Burt, D. M., 1979. Hydrothermal alteration,
in Barnes, H. L., ed., Geochemistry of hydrothermal ore
deposits, 3
rd
edition: New York, USA, John Wiley and Sons,
p. 173-235.
Ross, C. S., and Smith, R. L., 1955. Water and other volatiles in
volcanic glass: American Mineralogist, v. 40, p. 107189.
Ross, C. S., and Smith, R. L., 1960. Ash-flow tuffs - their origin,
geologic relations and identification: United States Geological
Survey Professional Paper 366, 54 p.
Saccocia, P. J., Ding, K., Berndt, M. E., Seewald, J. S., and Seyfried,
W. E., Jr., 1994. Experimental and theoretical perspectives
on crustal alteration at mid-ocean ridges, in Lentz, D. R.,
ed., Alteration and alteration processes associated with
ore-forming systems, II: Waterloo, Canada, Geological
Association of Canada short course notes, p. 403-31.
Saeki, Y, and Date, J., 1980. Computer application to the alteration
data of the footwall dacite lava at the Ezuri Kuroko deposits,
REFERENCES | 2 6 5
Akita Prefecture: Mining Geology, v. 30, p. 24150.
Saez, R., Pascual, E., Toscano, M., and Almodova, G. R., 1999. The
Iberian type of volcano-sedimentary massive sulfide deposits:
Mineralium Deposita, v. 34, p. 549-70.
Sakai, H., Shinaki, R., Kishima, M., and Tazaki, K., 1978. Oxygen
isotope exchange and chemical reactions in rhyolite-seawater
system at 300C and 1000 bars: EOS-Transactions, American
Geophysical Union, v. 59, p. 1220-1.
Sales, R. H., and Meyer, C, 1948. Wall-rock alteration at Butte,
Montana: American Institute of Mining, Metallurgical and
Petroleum Engineers Technical Publication, v. 12, no. 3,
25 p.
Sanchez-Espana, J., Velasco, E, and Yusta, I., 2000. Hydrothermal
alteration of felsic volcanic rocks associated with massive
sulfide deposition in the northern Iberian pyrite belt (SW
Spain): Applied Geochemistry, v. 15, p. 1265-90.
Sanchez-Espana, J., Carcia de Cortazar, A., Gil, P., and Velasco, E,
2002. The discovery of the Borobia world-class stratiform
magnesite deposit (Soria, Spain) a preliminary report:
Mineralium Deposita, v. 37, p. 240-3.
Sangster, D. E, 1972. Precambrian volcanogenic massive sulfide
deposits in Canada a review: Geological Society of Canada,
v. 72, p. 22-43.
Sato, T, 1973. A chloride complex model for Kuroko mineralisation:
Geochemical Journal, v. 7, p. 245-70.
Sato, T, 1974. Distribution and geologic setting of the Kuroko
deposits: Kozan Chishitsu Special Issue, v. 6, p. 19.
Sawkins, F. J., and Kowalik, J., 1981. The source of ore metals at
Buchans - magmatic versus leaching models, in Swanson,
E. A., Strong, D. E, and Thurlow, J. G., eds., The Buchans
ore bodies fifty years of geology and mining: Geological
Association of Canada Paper 22, p. 255-67.
Schardt, C, Cooke, D. R., Gemmell, J. B., and Large, R. R., 2001.
Geochemical modelling of the zones footwall alteration pipe,
Hellyer volcanic-hosted massive sulfide deposit, western
Tasmania, Australia: Economic Geology, v. 96, p. 1037-54.
Schardt, C, Yang, J., and Large, R., in press. Numerical heat and
fluid-flow modelling of the Panorama volcanic-hosted
massive sulfide district, Western Australia: Economic
Geology.
Schiffman, P., and Smith, B. M., 1988. Petrology and oxygen isotope
geochemistry of a fossil seawater hydrothermal system within
the Solea Graben, northern Troodos Ophiolite, Cyprus:
Journal of Geophysical Research, v. 93, p. 4612-24.
Schiffman, P., Elders, W A., Williams, A. E., McDowell, S. D., and
Bird, D. K., 1984. Active metasomatism in the Cerro Prieto
geothermal system, Baja California, Mexico a telescoped
low-pressure, low-temperature metamorphic fades series:
Geology, v. 12, p. 12-15.
Schiffman, P., Smith, B. M., Varga, R. J., and Moores, E. M., 1987.
Geometry, conditions and timing of off-axis hydrothermal
metamorphism and ore-deposition in the Solea graben
(Cyprus): Nature, v. 325, p. 423-5.
Schmidt, J., M., 1988. Mineral and whole-rock compositions of
seawater-dominated hydrothermal alteration at the Arctic
volcanogenic massive sulfide prospect, Alaska: Economic
Geology, v. 83, p. 822-42.
Schmincke, H-U., and von Rad, U., 1976. Neogene evolution
of Canary Island volcanism inferred from ash layers and
volcaniclastic sandstones of DSDP Site 397 (Leg 47A):
Proceedings of the Ocean Drilling Program, Initial Reports,
v. 47, part 1, p. 703-26.
Schmincke, H-U., 1967. Fused tuff and peperites in south-central
Washington: Geological Society of America Bulletin, v. 78,
p. 319-30.
Schwartz, G. M., 1959. Hydrothermal alteration: Economic
Geology, v. 54, p. 161-83.
2 6 6 | REFERENCES
Schweitzer, J. K., and Hatton, C. J., 1995. Chemical alteration
within the volcanic roof rocks of the Bushveld Complex:
Economic Geology, v. 90, p. 2218-31.
Scott, S. D., 1997. Submarine hydrothermal systems and deposits,
in Barnes, H. L., ed., Geochemistry of hydrothermal ore
deposits, New York, USA, John Wiley and Sons, p. 797-
876.
Scott, R. B., and Hajash, A., Jr., 1976. Initial submarine alteration of
basaltic pillow lavas a microprobe study: American Journal
of Science, v. 276, p. 480-501.
Secor, D. T., Jr., 1965. Role of fluid pressure in jointing: American
Journal of Science, v. 263, p. 633-46.
Sederholm, J. J., 1929. The use of the term 'deuteric': Economic
Geology, v. 24, p. 869-71.
Seki, Y., 1972. Lower grade stability limit of epidote in the light
of natural occurrences: Journal of the Geological Society of
Japan, v. 78, p. 405-513.
Seki, Y, Oki, Y, Matsuda, T., Mikami, K., and Okumura, K.,
1969. Metamorphism in the Tanzawa mountains, central
Japan: Journal of the Japanese Association of Mineralogists,
Petrologists and Economic Geologists, v. 61, p. 1-24.
Selley, D., 1997. Structure and sedimentology of the Dundas Group,
western Tasmania: Unpublished PhD thesis, University of
Tasmania, 188 p.
Seyfried, W. E., Jr., and Bischoff, J. L., 1977. Hydrothermal transport
of heavy metals by seawater: the role of seawater-basalt ratio:
Earth and Planetary Science Letters, v. 34, p. 71-7.
Seyfried, W. E., Jr., and Bischoff, J. L., 1979. Low temperature basalt
alteration by seawater - an experimental study at 70C and
150C: Geochimica et Cosmochimica Acta, v. 43, p. 1937-
47.
Seyfried, W. E., and Mottl, M. J., 1982. Hydrothermal alteration
of basalt by seawater under seawater-dominated conditions:
Geochimica et Cosmochimica Acta, v. 46, p. 985-1002.
Seyfried, W. E., Jr., Mottl, M. J., and Bischoff, J. L, 1978. Seawater-
basalt ratio effects on the chemistry and mineralogy of spilites
from the ocean floor: Nature, v. 275, p. 21113.
Seyfried, W. E., Jr., Berndt, M. E., and Seewald, J. S., 1988.
Hydrothermal alteration processes at mid-ocean ridges
constraints from diabase alteration experiments, hot spring
fluids and composition of the oceanic crust: Canadian
Mineralogist, v. 26, p. 787-804.
Sharpe, R., 1991. Geology and geochemistry of the barite and
siliceous caps, Hellyer Mine, Tasmania: Unpublished
Honours thesis, University of Tasmania, 114 p.
Sharpe, R., and Gemmell, J. B., 2001. Alteration characteristics
of the Archaean Golden Grove Formation at the Gossan
Hill deposit, Western Australia induration as a focusing
mechanism for mineralising hydrothermal fluids: Economic
Geology, v. 96, p. 1239-62.
Sheppard, R. A., and Gude, A. J., 3
rd
, 1968. Distribution and genesis
of authigenic silicate minerals in tuffs of Pleistocene Lake
Tecopa, Inyo County, California, United States Geological
Survey Professional Paper 597, 38 p.
Sheppard, R. A., and Gude, A. J., 3
rd
, 1973. Zeolites and associated
authigenic silicate minerals in tuffaceous rocks of the Big
Sandy Formation, Mohave County, Arizona, United States
Geological Survey Professional Paper P0830, 36 p.
Sheppard, R. A., Gude, A. J., 3
rd
, and Fitzpatrick, J. J., 1988.
Distribution, characterisation, and genesis of mordenite
in Miocene silicic tuffs at Yucca Mountain, Nye County,
Nevada: United States Geological Survey Bulletin, B1777,
p. 1-22.
Sherwood, G. J., 1988. Rock magnetic studies of Miocene volcanics
in eastern Otago and Banks Peninsula, New Zealand
comparison between Curie temperature and low temperature
susceptibility behaviour: New Zealand Journal of Geology
and Geophysics, v. 31, p. 225-35.
Shikazono, N., 1999. Rare earth element geochemistry of Kuroko
ores and hydrothermally altered rocks implication for
evolution of submarine hydrothermal system at back arc
basin: Resource Geology, v. 20, p. 2330.
Shikazono, N., Hoshino, M., Utada, M., Nakata, and Ueda, A.,
1998. Hydrothermal carbonates in altered wall rocks at the
Uwamuki Kuroko deposits, Japan: Mineralium Deposita,
v. 33, p. 346-58.
Shiraki, R., and Iiyama, J. T, 1990. Na-K ion exchange reaction
between rhyolite glass and (Na, K)C1 aqueous solution under
hydrothermal conditions: Geochimica et Cosmochimica
Acta, v. 54, p. 923-31.
Shiraki, R., Sakai, H., Endoh, M., and Kishima, N., 1987.
Experimental studies on rhyolite- and andesite-seawater
interactions at 300C and 1000 bars: Geochemical Journal,
v. 21, p. 139-48.
Shirozu, H., 1974. Clay minerals in altered wall rocks of the Kuroko-
type deposits: Kozan Chishitsu = Mining Geology, no. 6,
p. 303-10.
Sibson, R. H., 1977. Fault rocks and fault mechanisms: Journal
Geological Society, v. 133, p. 191213.
Sillitoe, R. H., Hannington, M. D., and Thompson, J. F. H., 1996.
High-sulfidation deposits in the volcanogenic massive sulfide
environment: Economic Geology, v. 91, p. 204-12.
Simmons, S. E, and Browne, P. R. L., 2000. Hydrothermal minerals
and precious metals in the Broadlands-Ohaaki geothermal
system - implications for understanding low-sulfidation
epithermal environments: Economic Geology, v. 95, p. 971-
99.
Simmons, S. F., and Christenson, B. W, 1994. Origins of calcite in
a boiling geothermal system: American Journal of Science,
v. 294, p. 361-400.
Simpson, K. A., 2001. Volcanic facies architecture of the Seventy
Mile Range Group, Queensland: Unpublished PhD thesis,
University of Tasmania, 217 p.
Simpson, K. A., andMcPhie, J., 1998. Characteristics of subaqueous
basaltic andesite fire fountain deposits - an example from the
Mount Windsor Volcanics, northern Queensland, Australia
[abs.], in Beams, S. D., ed., Economic geology of northeast
Queensland, the 1998 perspective: Townsville, Australia,
July 6-10, 1998, Geological Society of Australia Abstracts,
v. 49, p. 409.
Sippel, R. E, 1968. Sandstone petrology evidence from luminescence
petrography: Journal of Sedimentary Petrology, v. 38,
p. 530-54.
Sise, J. R., and Jack, D. J., 1984. Exploration case history of the
Hellyer Deposit, in Baillie, P. W, and Collins, P. L. E, eds.,
Mineral exploration and tectonic processes in Tasmania:
Burnie, Australia, November 1984, Geological Society
of Australia, Tasmanian Division, Abstract volume and
excursion guide, p. 489.
Sivell, W. J., 1984. Low-grade metamorphism of the Brook Street
Volcanics, D'Urville Island, New Zealand: New Zealand
Journal of Geology and Geophysics, v. 27, p. 16790.
Skirrow, R. G., and Franklin, J. M., 1994. Silicification and metal
leaching in semi-conformable alteration beneath the Chisel
Lake massive sulfide deposit, Snow Lake, Manitoba:
Economic Geology, v. 89, p. 31-50.
Smith, R. L., I960. Ash flows: Geological Society of America
Bulletin, v. 71, p. 795-842.
Smith, R. E., 1968. Redistribution of major elements in the alteration
of some basic lavas during burial metamorphism: Journal of
Petrology, v. 9, p. 191-219.
Smith, R. E., 1969. Zones of progressive regional burial
metamorphism in part of the Tasman Geosyncline, eastern
Australia: Journal of Petrology, v. 10, p. 144-63.
Smith, R. E., 1974. The production of spilitic lithologies by
burial metamorphism of flood basalts from the Canadian
Keweenawan, Lake Superior, in Amstutz, G. C, ed., Spilites
and spilitic rocks: New York, USA, Springer-Verlag, p. 403
16.
Smith, R. E., 1977. Petrography and geochemistry of epidote
alteration patches in gabbro dykes at Matagami, Quebec
- discussion: Canadian Journal of Earth Sciences, v. 14,
p. 505-7.
Smith, R. B., 1991. Diagenesis and cementation of lower Miocene
pyroclastic sequences in the Sulu Sea, Sites 768, 769 and
771: Proceedings of the Ocean Drilling Program, Scientific
Results, v. 124, p. 181-99.
Smith, R. E., and Smith, S. E., 1976. Comments of the use ofTi, Zr,
Y, Sr, K, P and Nb in classification of basaltic magmas: Earth
and Planetary Science Letters, v. 32, p. 11420.
Smith, R. E., Perdrix, J. L., and Parks, T. C, 1982. Burial
metamorphism in the Hamersley Basin, Western Australia:
Journal of Petrology, v. 23, p. 75-102.
Solomon, M., 1964. The spilite-keratophyre association of west
Tasmania and the ore deposits of Mt Lyell, Rosebery and
Hercules: Unpublished PhD thesis, University of Tasmania,
419 p.
Solomon, M., 1976. 'Volcanic' massive sulfide deposits and their
host rocks a review and an explanation, in Wolf, K. A.,
ed., Handbook of stratabound and stratiform ore deposits
- Part II, regional studies and specific deposits: Amsterdam,
Netherlands, Elsevier, 320 p.
Solomon, M., 1981. An introduction to the geology and metallic
ore deposits of Tasmania: Economic Geology, v. 76, p. 194
208.
Solomon, M., 1999. 'Mineralisation' use and mis-use: SEG
Newsletter, v. 37, p. 48.
Solomon, M., and Gaspar, O. C, 2001. Textures of the Hellyer
volcanic-hosted massive sulfide deposit, Tasmania the
ageing of a sulphur sediment on the seafloor: Economic
Geology, v. 96, p. 1513-34.
Solomon, M., and Groves, D. I., 1994. The geology and origin of
Australia's mineral deposits: Oxford, UK, Oxford University
Press, Monographs in Geology and Geophysics, v. 24,
951 p.
Solomon, M., and Khin Zaw, 1997. Formation on the seafloor of
the Hellyer volcanogenic massive sulfide deposit: Economic
Geology, v. 92, p. 686-95.
Solomon, M., and Khin Zaw, 1999. Formation on the seafloor of
the Hellyer volcanogenic massive sulfide deposit: Economic
Geology, v. 92, p. 686-95.
Solomon, M., and Quesada, C, 2003. Zn-Pb-Cu massive sulfide
deposits brine-pool types occur in collisional orogens,
black smoker types occur in backarc and/or arc basins:
Geology, v. 31, p. 1029-32.
Solomon, M., Rafter, T. A., and Jensen, M. L., 1969. Isotope studies
on the Rosebery, Mount Farrell and Mount Lyell ores,
Tasmania: Mineralium Deposita, v. 4, p. 17299.
Solomon, M., Walshe, J. L., and Palomero, G. F., 1980. Formation
of massive sulfide deposits at Rio Tinto, Spain: Institution
of Mining and Metallurgy, Transactions, Section B: Applied
Earth Science, v. 89, p. 16-24.
Solomon, M, Eastoe, C. J., Walshe, J. L, and Green, G. R, 1988.
Mineral deposits and sulfur isotope abundances in the
Mount Read Volcanics between Que River and Mount
Darwin, Tasmania: Economic Geology, v. 83, p. 130728.
Solomon, M., Tornos, E, and Gaspar, O. C, 2002. Explanation for
many of the unusual features of the massive sulfide deposits
of the Iberian pyrite belt: Geology, v. 30, p. 87-90.
Solomon, M., Gemmell, J. B., and Khin Zaw, 2004. Nature and
origin of the fluids responsible for forming the Hellyer Zn-
REFERENCES | 2 6 7
Pb-Cu, volcanic-hosted massive sulfide deposit, Tasmania,
using fluid inclusions, and stable and radiogenic isotopes:
Ore Geology Reviews, v. 25, p. 89-124.
Sparks, R. S. J., and Wright, J. V, 1979. Welded air-fall tuffs: Geological
Society of America Special Paper, no. 180, p. 155-66.
Sparks, R. S. J., Stasiuk, M. V., Gardeweg, M., and Swanson, D.
A., 1993. Welded breccias in andesite lavas: Journal of the
Geological Society of London, v. 150, p. 897-902.
Spooner, E. T. C, 1977. Hydrodynamic model for the origin of the
ophiolitic cupriferous pyrite ore deposits of Cyprus: London,
UK, January 2122, 1976, Volcanic processes in ore genesis,
Proceedings, Institute of Mining and Metallurgy.
Spooner, E. T. C, and Fyfe, W S., 1973. Subseafloor metamorphism,
heat and mass transfer: Contributions to Mineralogy and
Petrology, v. 42, p. 287-304.
Spooner, E. T. C, Beckinsale, R. D., Fyfe, W S., and Smewing, J.
D., 1974. 8O
18
enriched ophiolitic metabasic rocks from
E. Liguria (Italy), Pindos (Greece), and Troodos (Cyprus):
Contributions to Mineralogy and Petrology, v. 47, p. 4162.
Spry, A., 1976. Metamorphic textures: Exeter, UK, Pergamon Press,
350 p.
Stakes, D. S., and Taylor, H. P., Jr., 1992. The northern Samail
Ophiolite - an oxygen isotope, microprobe, and field study:
Journal of Geophysical Research, v. 97, p. 704380.
Stanley, C. R., and Madeisky, H. E., 1996. Lithogeochemical
exploration for metasomatic zones associated -with
hydrothermal mineral deposits using Pearce element ratio
analysis: Vancouver, Canada, Mineral Deposit Research
Unit, University of British Columbia short course notes,
January 1996, 98 p.
Stanton, R. L., 1955. Lower Palaeozoic mineralisation near Bathurst,
New South Wales: Economic Geology, v. 50, p. 681-714.
Stanton, R. L., 1959. Mineralogical features and possible mode
of emplacement of the Brunswick Mining and Smelting
ore bodies, Gloucester County, New Brunswick: Canadian
Mining and Metallurgical Bulletin, v. 570, no. 62, p. 631-42.
Stanton, R. L., 1984. The direct derivation of cordierite from a
clay-chlorite precursor evidence from the Geco Mine,
Manitouwadge, Ontario: Economic Geology, v. 79, p. 1245
64.
Stanton, R. L., 1985. Stratiform ores and geological processes:
Journal and Proceedings of the Royal Society of New South
Wales, v. 118, p. 77-100.
Stanton, R. L., 1990. Magmatic evolution and the ore type-lava type
affiliations of volcanic exhalative ores: Australasian Institute
of Mining and Metallurgy Monograph, v. 15, p. 101-7.
Starkey, R. J., Jr., and Frost, B. R., 1990. Low-grade metamorphism
of the Karmutsen Volcanics, Vancouver Island, British
Columbia: Journal of Petrology, v. 31, p. 167-95.
Steiner, A., 1953. Hydrothermal rock alteration at Wairakei, New
Zealand: Economic Geology, v. 48, p. 113.
Stolz, A. J., 1995. Geochemistry of the Mount Windsor Volcanics
- implications for the tectonic setting of Cambro-
Ordovician volcanic-hosted massive sulfide mineralisation
in northeastern Australia: Economic Geology, v. 90,
p. 1080-97.
Stolz, A. J., Davidson, G. J., and Blake, M., 1996. Barren alteration
systems - example, Gydgie Central, Mt Windsor Volcanic
Belt, North Queensland: AMIRA/ARC Project P439,
Centre for Ore Deposit and Exploration Studies, University
of Tasmania, Unpublished Report 3, October 1996, p. 1
85.
Strauss, G. K., Roger, G., Lecolle, M., and Lopera, E., 1981.
Geochemical and geologic study of the volcano-sedimentary
sulfide ore body of La Zarza, Province of Huelva, Spain:
Economic Geology, v. 76, p. 1975-2000.
Surdam, R. C, 1968. Origin of native copper and hematite in the
2 6 8 | REFERENCES
Karmutsen Group, Vancouver Island, British Columbia:
Economic Geology, v. 63, p. 9616.
Surdam, R. C, 1973. Low-grade metamorphism of tuffaceous
rocks in the Karmutsen Group, Vancouver Island, British
Columbia: Geological Society of America Bulletin, v. 84,
p. 1911-22.
Surdam, R. C, and Boles, J. R., 1979. Diagenesis of volcanic
sandstones: Society of Economic Palaeontologists and
Mineralogists Special Publication, v. 26, p. 127-39.
Surdam, R. C, and Sheppard, R. A., 1978. Zeolites in saline, alkaline-
lake deposits, in Sand, L. B., and Mumpton, E A., eds.,
Natural zeolites occurrence, properties, use: Oxford, UK,
Peregamon Press, p. 14574.
Syddell, M., 2004. Down to earth: Earthmatters (CSIRO Exploration
and Mining), v. 5, p. 8-9.
Tanimura, S., Date, J.,Takahashi,T., and Ohmoto, H., 1983. Geologic
setting of the Kuroko deposits, Japan - Part II, stratigraphy
and structure of the Hokuroku district: Economic Geology
Monographs, v. 5, p. 24-38.
Taranik, J. V., 2001. New applications of remote sensing to mineral
exploration [abs.]: Geological Society of America Abstracts
with Programs, v. 33, p. 347.
Taube, A., 1986. The Mount Morgan gold-copper mine and
environment, Queensland a volcanogenic massive sulfide
deposit associated with penecontemporaneous faulting:
Economic Geology, v. 81, p. 1322-40.
Taylor, J. C. M., 1950. Pore space reductions in sandstones: Bulletin
of American Association Petrology Geology, v. 34, p. 701
6.
Taylor, H. P., Jr., 1974. The application of oxygen and hydrogen
isotope studies to problems of hydrothermal alteration and
ore deposition: Economic Geology, v. 69, p. 843-83
Taylor, H. P., Jr., 1979. Oxygen and hydrogen isotope relationships
in hydrothermal mineral deposits, in Barnes, H. L., ed.,
Geochemistry of hydrothermal ore deposits, 3
rd
edition:
New York, USA, John Wiley and Sons, p. 236-77.
Taylor, R. G., 1992. Ore textures recognition and interpretation
- Volume 1, infill textures: Townsville, Australia, Economic
Geology Research Unit, Geology Department, James Cook
University, 24 p.
Taylor, R. G., 1996. Ore textures recognition and interpretation
- Volume 3, overprinting textures: Townsville, Australia,
Economic Geology Research Unit, Geology Department,
James Cook University, 52 p.
Taylor, H. P., Jr., and Forester, R. W., 1979. An oxygen and hydrogen
isotope study of the Skaergaard intrusion and its country
rocks - a description of a 55-m.y. old fossil hydrothermal
system: Journal of Petrology, v. 20, p. 355-419.
Taylor, M. W., and Surdam, R. C, 1981. Zeolite reactions in the
tuffaceous sediments at Teels Marsh, Nevada: Clays and
Clay Minerals, v. 29, p. 341-52.
Taylor, B., Fujioka, K., Janecek, T. R., Aitchison, J., Cisowski, S.,
Colella, A., Cooper, P. A., Dadey, K. A., Egeberg, P. K., Firth,
J. V, Gill, J. B., Herman, Y., Hiscott, R. N., Isiminger-Kelso,
M., Kaiho, K., Klaus, A., Koyama, M., Lapierre, H., Lovell,
M. A., Marsaglia, K., Nishimura, A., Pezard, P. A., Rodolfo,
K. S., Taylor, R. N., Tazaki, K., and Torssander, P., 1990.
Sites 788/789: Proceedings of the Ocean Drilling Program,
Initial Reports, v. 126, p. 97-124.
Tazaki, K., and Fyfe, W. S., 1992. Diagenetic and hydrothermal
mineral alteration observed in Izu-Bonin deep-sea sediments,
Leg 126: Proceedings of the Ocean Drilling Program,
Scientific Results, v. 126, p. 101-12.
Thompson, A. B., 1971. Analcite-albite equilibria at low
temperatures: American Journal of Science, v. 271, p. 79
92.
Thompson, G., 1973. A geochemical study of the low-temperature
interaction of seawater and oceanic igneous rocks: EOS-
Transactions, American Geophysical Union, v. 54, p. 1015
19.
Thompson, A. J. B., and Thompson, J. F. EL, 1996. Atlas of alteration
- a field and petrographic guide to hydrothermal alteration
minerals: Vancouver, Canada, Geological Association of
Canada, 119 p.
Thompson, A. J. B., Hauff, P. L., and Robitaille, A. J., 1999.
Alteration mapping in exploration - application of short-
wave infrared (SWIR) spectroscopy: SEG Newsletter, v. 39,
p. 1 and 16-27.
Titley, S. R., 1982a. Advances in geology of porphyry copper
deposits, southwestern North America: Tucson, USA,
University of Arizona Press, 560 p.
Titley, S. R., 1982b. The style and progress of mineralisation and
alteration in porphyry copper systems; American southwest,
in Titley, S. R., ed., Advances in geology of porphyry copper
deposits, southwestern North America: Tucson, USA,
University of Arizona Press, p. 93-116.
Titley, S. R., 1994. Evolutionary habits of hydrothermal and
supergene alteration in intrusion-centred ore systems,
southwestern North America, in Lentz, D. R., ed., Alteration
and alteration processes associated with ore-forming systems,
2: Vancouver, Geological Association of Canada short course
notes, May 13-15, 1994, p. 237-260.
Titley, S. R., Fleming, A. W., and Neale, T. I., 1978. Tectonic
evolution of the porphyry copper system at Yandera, Papua
New Guinea: Economic Geology, v. 73, p. 810-28.
Tiwary, A., and Deb, M., 1997. Geochemistry of hydrothermal
alteration at the Deri massive sulfide deposit, Sirohi District,
Rajasthan, NW India: Journal of Geochemical Exploration,
v. 59, p. 99-121.
Tornos, E, Clavijo, E. G., and Spiro, B., 1998. The Filon Norte
ore body (Tharsis, Iberian pyrite belt) a proximal low-
temperature shale-hosted massive sulfide in a thin-skinned
tectonic belt: Mineralium Deposita, v. 33, p. 150-69.
Tornos, E, Casquet, C, and Relvas, J. M. R. S., in press.
Transpressional tectonics, lower crust decoupling and
intrusion of deep mafic sills - a model for the unusual
metallogenesis of SW Iberia: Ore Geology Reviews.
Torres, M. E., Marsaglia, K. M., Martin, J. B., and Murray, R. W.,
1995. Sediment diagenesis in western Pacific basins: AGU
Geophysical Monograph, v. 88, p. 241-58.
Tsolis-Katagas, P., and Katagas, C, 1989. Zeolites in pre-caldera
pyroclastic rocks of the Santorini volcano, Aegean Sea,
Greece: Clays and Clay Minerals, v. 37, p. 497-510.
Tucker, M. E., 1987. Sedimentary petrology - an introduction:
Oxford, UK, Blackwell Scientific Publications, 252 p.
Turner, F. J., 1980. Metamorphic petrology mineralogical, field and
tectonic aspects, 2
nd
edition: USA, Hemisphere Publishing
Corporation, 524 p.
Upton, B. G. J., and Wadsworth, W. J., 1970. Early volcanic
rocks of Reunion and their tectonic significance: Bulletin
Volcanologique, v. 33, p. 1246-68.
Urabe, T, 1987. The effect of pressure on the partitioning ratios of
lead and zinc between vapour and rhyolite melts: Economic
Geology, v. 82, p. 1049-52.
Urabe, T, and Sato, T., 1978. Kuroko deposits of the Kosaka
Mine, northeast Honshu, Japan - products of submarine
hot springs on Miocene seafloor: Economic Geology, v. 73,
p. 161-79.
Urabe, T, Scott, S. D., and Hattori, K., 1983. A comparison of
footwall-rock alteration and geothermal systems beneath
some Japanese and Canadian volcanogenic massive sulfide
deposits: Economic Geology Monographs, v. 5, p. 34564.
Utada, M., 1970. Occurrence and distribution of authigenic zeolites
in the Neogene pyroclastic rocks in Japan: Scientific Papers
REFERENCES I 2 6 9
of the College of General Education, University of Tokyo,
v. 20, p. 191-262.
Utada, M., 1973. The types of alteration in the Neogene sediments
relating to the intrusion of volcano-plutonic complexes in
Japan: Scientific Papers of the College of General Education,
University of Tokyo, v. 23, p. 167216.
Utada, M., 1991. Zeolitization in the Neogene formations of Japan:
Episodes, v. 14, p. 242-5.
Utada, M., Minato, H., Ishikawa, T, and Yoshizaki, Y., 1974. The
alteration zones surrounding Kuroko deposits in Nishi-
Aizu District, Fukushima Prefecture, with emphasis on the
analcime zone as an indicator in exploration for ore deposits:
Mining Geology, v. 6, p. 291-302.
Utada, M., Aoki, M., Inoue, A., and Kusakabe, H., 1988. A
hydrothermal alteration envelope and alteration minerals
in the Shinzan mineralised area, Akita Prefecture, northeast
Japan: Mining Geology Special Issue, v. 12, p. 6777.
Viereck, L. G., Griffin, B. J., Schmincke, H-U., and Pritchard, R.
G., 1982. Volcaniclastic rocks of the Reydarfjordur drill hole,
eastern Iceland - Part 2, alteration: Journal of Geophysical
Research, v. 87, p. 6459-76.
von Damm, K. L., 1995. Controls on the chemistry and temporal
variability of seafloor hydrothermal fluids: AGU Geophysical
Monograph, v. 91, p. 222-47.
Walshe, J. L., 1986. A six-component chlorite solid solution model
and the conditions of chlorite formation in hydrothermal
and geothermal systems: Economic Geology, v. 81, p. 681
703.
Walshe, J. L., and Solomon, M., 1981. An investigation into the
environment of formation of the volcanic-hosted Mt Lyell
copper deposits using geology, mineralogy, stable isotopes
and a six-component chlorite solid solution model: Economic
Geology, v. 76, p. 246-84.
Walton, A. W, 1975. Zeolitic diagenesis in Oligocene volcanic
sediments, Trans-Pecos Texas: Geological Society of America
Bulletin, v. 86, p. 615-24.
Waring, C. L., Andrew, A. S., and Ewers, G. R., 1998. Use of O,
C, and S stable isotope in regional mineral exploration:
AGSO Journal of Australian Geology and Geophysics, v. 17,
p. 301-13.
Waters, J. C, and Wallace, D. B., 1992. Volcanology and
sedimentology of the host succession to the Hellyer and Que
River volcanic-hosted massive sulfide deposits, northwestern
Tasmania: Economic Geology, v. 87, p. 650-66.
Weaver, S. D., Gibson, I. L, Houghton, B. E, and Wilson, C.
J. N., 1990. Mobility of rare earth and other elements
during crystallization of peralkaline silicic lavas: Journal of
Volcanology and Geothermal Research, v. 43, p. 57-70.
Webster, S. S., and Skey, E. H., 1979. Geophysical and geochemical
case history of the Que River deposit, Tasmania, Australia:
Ottawa Geological Survey of Canada Economic Geology
Report, no. 31, p. 697-720.
Weihed, P., Allen, R. L., and Svenson, S. A., 2000. Metallogeny and
tectonic evolution of the c. 1.9 Ga Skellefte marine, volcanic
arc, northern Sweden [abs.], in Gemmell, J. B., and Pongratz,
J., eds., Volcanic environments and massive sulfide deposits,
Hobart, Australia, November 16-19, 2000, CODES Special
Publication, v. 3, p. 225-226.
Whalen, J. B., McNicoll, V. J., Galley, A. G., and Longstaffe, F. J.,
2004. Tectonic and metallogenic importance of an Archaean
composite high- and low-Al tonalite suite, Western Superior
Province, Canada: Precambrian Research, v. 132, p. 275
301.
Whetten, J. T, and Hawkins, J. W, 1970. Diagenetic origin of
greywacke matrix minerals: Sedimentology, v. 15, p. 347-
61.
White, D. E., 1970. Geochemistry applied to the discovery,
evaluation and exploitation of geothermal energy resources:
Geothermatics Special issue 2, v. 1, p. 58-80.
White, M. J., 1996. Stratigraphy, volcanology and sedimentology
of the Cambrian Tyndall Group, Mount Read Volcanics,
western Tasmania: Unpublished PhD thesis, University of
Tasmania, 196 p.
White, J. D. L., 2000. Subaqueous eruption-fed density currents and
their deposits: Precambrian Research, v. 101, p. 87-109.
White, N. C, andHedenquist,J. W, 1990. Epithermalenvironments
and styles of mineralisation variations and their causes,
and guidelines for exploration: Journal of Geochemical
Exploration, v. 36, p. 445-74.
White, M. J., and McPhie, J., 1996. Stratigraphy and
palaeovolcanology of the Cambrian Tyndall Group, Mt
Read Volcanics, western Tasmania: Australian Journal of
Earth Sciences, v. 43, p. 147-59.
White, M. J., and McPhie, J., 1997. A submarine welded ignimbrite-
crystal-rich sandstone facies association in the Cambrian
Tyndall Group, western Tasmania, Australia: Journal of
Volcanology and Geothermal Research, v. 76, p. 277-95.
White, D. E., and Sigvaldason, G. E., 1962. Epidote in hot-spring
systems, and depth of formation of propylitic epidote in
epithermal ore deposits: United States Geological Survey
Professional Paper 450E, p. 80-4.
Whitmore, D. R. E., 1969. Geology of the Coronation copper
deposit: Geological Survey of Canada Paper, v. 68-5, p. 37
53.
Whitten, D. G. A., and Brooks, J. R. V, 1972. The Penguin
dictionary of geology: London, UK, Penguin Books, 493 p.
Williams, N. C, 2000. The Basin Lake high-sulfidation alteration
system, western Tasmania: Unpublished B.Sc. Honours
thesis, University of Tasmania, 102 p.
Williams, N. C, and Davidson, G., 2004. Possible submarine
advanced argillic alteration at the Basin Lake prospect,
western Tasmania, Australia: Economic Geology, v. 99,
p. 987-1002.
Williams, D., Stanton, R. L., and Rambaud, E, 1975. The Planes-
San Antonio pyritic deposit of Rio Tinto, Spain its
nature, environment and genesis: Institution of Mining and
Metallurgy Transactions, v. 84, p. B73-82.
Williams, H., Turner, F. J., and Gilbert, C. M., 1982. Petrography
an introduction to the study of rocks in thin section, 2
nd
edition: San Fransico, USA, WH. Freeman and Co., 626 p.
Williams, E., McClenaghan, M. P., and Collins, P. L. E, 1989.
Mid-Palaeozoic deformation, granitoids and ore deposit, in
Burrett, C. F., and Martin, E. L., eds., Geology and mineral
resources of Tasmania: Geological Society of Australia Special
Publication, v. 15, p. 238-92.
Wilshire, H. G., 1959. Deuteric alteration of volcanic rocks: Journal
and Proceedings, Royal Society of New South Wales, v. 93,
p. 105-20.
Wilshire, H. G., and Hobbs, B. E., 1962. Structure, sedimentary
inclusions, and hydrothermal alteration of a latite intrusion:
Journal of Geology, v. 70, p. 32841.
Wilson, A. J., Cooke, D. R., and Harper, B. L., 2003. The Ridgeway
gold-copper deposit - a high-grade alkalic porphyry deposit
in the Lachlan Fold Belt, New South Wales, Australia:
Economic Geology, v. 98, p. 163766.
Winkler, H. G. E, 1979. Petrogenesis of metamorphic rocks, 5th
ed.: New York, USA, Springer-Verlag, 348 p.
WoldeGabriel, G., Keating, G. N., and Valentine, G. A., 1999.
Effects of shallow basaltic intrusion into pyroclastic deposits,
Grants Ridge, New Mexico, USA: Journal of Volcanology
and Geothermal Research, v. 92, p. 389-411.
Wyman, W. E, 2001. Cambrian granite-related hydrothermal
alteration and Cu-Au mineralisation in the southern Mt
Read Volcanics, western Tasmania, Australia: Unpublished
2 7 0 | REFERENCES
PhD thesis, University of Tasmania, 343 p.
Yamagishi, H., 1987. Studies on the Neogene subaqueous lavas and
hyaloclastites in southwest Hokkaido: Geological Survey of
Hokkaido Report, v. 59, p. 55-117.
Yamagishi, H., and Dimroth, E., 1985. A comparison of Miocene
and Archaean rhyolite hyaloclastites evidence for a hot and
fluid rhyolite lava: Journal of Volcanology and Geothermal
Research, v. 23, p. 337-55.
Yamagishi, H., and Goto, Y, 1992. Cooling joints of subaqueous
rhyolite lavas at Kuroiwa, Yakumo, southern Hokkaido,
Japan: Bulletin of the Volcanological Society of Japan, v. 37,
p. 205-7.
Yang, K., 1998. Compositional variations of white micas - a
literature review: North Ryde, CSIRO, 23 p.
Yang, J., and Large, R. R., 2001. Computational modelling of
hydrothermal ore-forming fluid migration in complex
earth structures, in Xie, H., Wang, Y., and Jiang, Y, eds.,
Proceedings of the International Symposium on Computer
Applications in the Mineral Industries, Beijing, China, April
25-27, 2001, v. 29, p. 115-20.
Yang, K., and Scott, S. D., 1996. Possible contribution of a metal-
rich magmatic fluid to a seafloor hydrothermal system:
Nature, v. 383, p. 420-3.
Yang, K., Huntington, J. E, Boardman, J. W., and Mason, P., 1999.
Mapping hydrothermal alteration in the Comstock mining
district, Nevada, using simulated satellite-borne hyperspectral
data: Australian Journal of Earth Sciences, v. 46, p. 420-3.
Yardley, B. W. D., 1989. An introduction to metamorphic petrology:
Harlow, UK, Longman Scientific and Technical, 248 p.
Ylagan, R. E, Altaner, S. P., and Pozzuoli, A., 1996. Hydrothermal
alteration of a rhyolitic hyaloclastite from Ponza Island,
Italy: Journal of Volcanology and Geothermal Research,
v. 74, p. 215-31.
Yoshida, K., and Utada, M., 1968. A study on alteration of Miocene
green tuff in the Kuroko-type mineralisation area in Odate
basin, Akita Prefecture: Mining Geology, v. 18, p. 333-42.
Yu, Z., Robinson, P., and McGoldrick, P. J., 2001. An evaluation
of methods for the chemical decomposition of geological
materials for trace element determination using ICP-MS:
Geostandards Newsletter, v. 25, no. 2-3, p. 199-217.
Zhang, X-C, Spiro, B., Halls, C, Stanley, C. J., and Yang, K-Y,
2003. Sediment-hosted disseminated gold deposits in
Southwest Guizhou, PRC their geological setting and
origin in relation to mineralogical, fluid inclusion, and
stable-isotope characteristics: International Geology Review,
v. 45, p. 407-70.
Zhou, Z., and Fyfe, W. S., 1989. Palagonitization of basaltic glass from
DSDP Site 335, Leg 37 - textures, chemical composition
and mechanism of formation: American Mineralogist, v. 74,
p. 1045-53.
I 27 1
INDEX
A page number in bold indicates that the reference is to a
figure. A bold t indicates that the reference is to a table.
AI-CCPI Alteration box plot see Alteration box plot
albite alteration 39, 42-44, 61, 65, 69, 133-5, 165-7, 191,
214,216
Alteration box plot 31-4, 36, 169-70, 245
alteration distribution 22, 63-4
alteration facies
describing and defining 1536
diagenetic 2412
distribution 63-4
hydrothermal 242-3
metamorphic 242
variables 22
alteration fluids 93, 170-3
Alteration Index (AI) 30-2, 34, 169, 245
alteration indices 26-30, 34, 73, 169-70, 244^5
alteration intensity
describing 27t
estimation, integrated approach 336
explained 2536
illustrated 28-29, 36
lithogeochemical indications of 243-4
alteration mineral assemblages see mineral assemblages
alteration nomenclature 16, 19-22, 2It, 22t, 23t
alteration pipes 63, 164-74, 176, 182-3
alteration plumes 63, 167, 168, 191-3
alteration processes 46
alteration rates 6
alteration textures
deformation textures 52, 54-6, 55-7
described 37-63, 37t, 38t, 62t
dissolution textures 41, 50- 1, 52
dynamic recrystallisation textures 52
illustrated 39-40, 62, 103-4, 110, 111-13
infill textures 41, 48- 9
overprinting and false/pseudo textures 37, 54-63,
58-61, 62t
recrystallisation textures 52, 53
replacement textures 37-8, 41, 42-7
static recrystallisation textures 52, 53
alteration timing 69-71, 7It, 172
alteration zonation
boundaries 64
contact altered zones 5, 64, 66-7, 67, 139, 149-56,
242
diagenetic 64, 105-8
facies model 3
greenschist facies zones 115, 116, H6t, 131, 142,
144-5, 152
halos 66-9
Hellyer deposit 178, 182-3, 184-93
Henty deposit 178, 212-3, 214-20
Highway Reward deposit 14, 167, 178, 232, 233-40
Hokuroku Basin 119-27
hydrothermal 5-6, 66-8, 164-9, 243
mapping 243
metamorphic 645, 667
Mount Read Volcanics 128-38
patterns 64-9, 98, 165-8
regional deep semi-conformable 66, 142-8
regional metamorphic 115-17,140
Rosebery deposit 178, 195, 196-201
scales described 64t
Thalanga deposit 14, 178, 222, 223-31
veins and fractures 67, 69
Western Tharsis deposit 202-3, 204-11
amphibolite facies 115-17,140
Amulet deposit see Noranda district
analytical techniques
electron microprobe 19,25,88
field observations 18
ICP-AES (inductively coupled plasma atomic emission
spectrometry) 76-7
ICP-MS (inductively coupled plasma mass
spectrometry 767
isotope geochemistry 925
HyMap 246, 247
lithogeochemical sampling 73-87
mineral chemistry analysis 87-91
NAA (neutron activated analysis) 76
petrography 24-5, 33-4
PIMA 25, 33, 245
2 7 2 | INDEX
SWIR spectroscopy 19, 24, 25, 33, 88, 90, 202-3,
243, 245
X-ray diffraction (XRD) 19, 24, 25, 33, 88
X-ray fluorescence spectrometry (XRF) 76-7
anhydrous minerals 97
B
Bathurst mining camp 164-5,170
Boco prospect 248, 250
burial-related alteration 97-138 see also diagenesis and
submarine environments
c
carbonates
diagenetic 105
in exploration 90-1, 243-4
hydrothermal 47, 91, 166, 178, 188, 201, 217,
229-30
cataclastic texture 52
CCPI (chlorite-carbonate-pyrite index) see also Alteration
box plot
explained 31, 34
exploration, uses for 169,245
cementation 97, 102, 105, 108-10, 132
Central Volcanic Complex see Mount Read Volcanics
chlorite 19-22, 89, 138, 156, 165-8, 187, 239
closure
and alteration indices 26
constant sum effect 2434
explained 78-9, 243-5
mass change anomalies 81
compaction 97, 109-10, 132
compositional nomenclature 20
contact alteration 5, 149-62
corrosion vugs 41, 5 01 , 52
crystallisation
primary 4
textures 52
of zeolite assemblages 105, 110, 114, 121
D
Darwin Granite see Mount Read Volcanics
data sheets
contents of 36
Darwin Granite 1 5 7 -6 2
Hellyer deposit 1 84-93
Henty deposit 21 420
Highway-Reward deposit 233-40
Hokuroku Basin 1 22-7
Mount Read Volcanics 1 33-8
Rosebery deposit 1 96 -201
Thalanga deposit 223-31
Western Tharsis deposit 2041 1
deep, semi-conformable altered zones 1426
deformation textures 52, 54-6, 5 5 -7
detection limit explained 75
deuteric alteration 148
devitrification
explained 4
texture 37, 39, 6 2
zones 151
diagenesis
explained 5
Hokuroku Basin 118-27
isotope geochemistry analysis 93-4
and metamorphism 16, 98, 102, 114, 115
Mount Read Volcanics 128-38
in submarine volcanic successions 97, 10214
diagenetic minerals
carbonates 105
genesis of 108-14
layered silicates 102,105
other diagenetic minerals 105
zeolites 105, 110, 118, 120-7
diagenetic zones
Hokuroku Basin 11827
Mount Read Volcanics 128-38
zonation 64-5, 105-8
discharge zone 1412
dissolution 41, 5 0-1 , 52, 97, 102, 108-10, 114, 132
dynamic recrystallisation textures 52
electron microprobe see analytical techniques
element concentrations 32-3
eutaxitic texture 54, 5 7
exploration
Alteration box plot 31-4, 169-70
alteration identification as tool in 24150
Alteration Index (AI) 30-2
isotope geochemistry in 92-5, 246-50
lithogeochemistry in 73-87
mineral chemistry in 87-91
sulfide mapping 243
use of chlorite in 89
use of white mica in 901
vectors and proximity indicators 32-3, 94-5, 243-50
false textures see pseudotextures
fiamme 54, 5 7
fluid-rock interaction 169, 170-2, 1 7 1 , 173
foliation 52, 54, 70, 711
footwall alteration 163-74, 179, 182-9
fused zones 151
geochemistry see isotope geochemistry, lithogeochemistry
geothermal gradient 98
geothermometers 92-3
glass
alteration in submarine volcanic successions 15, 97-8
common alteration minerals 19t
crystallisation 4
diagenesis 10214
disequilibrium assemblages 24
hydration 4-5, 98-102
reactive quality, 6
Green Tuff Belt see Hokuroku district
INDEX | 2 7 3
H
halos 5-6, 38, 41, 66-9, 149-56, 1 5 7 -6 2, 163-74, 178-
81
hanging wall alteration 163, 164, 167-8, 190-3
Hellyer deposit see also Mount Read province
alteration fades and zonation 178, 182-3, 1 8493
Alteration Index (AI) 31, 183
explained 11, 181-93
exploration 245
geological setting 1812
ore genesis 170, 183
white mica 90
Henty deposit see also Mount Read province
alteration facies and zonation 178, 212-13, 21 4-20
explained 12
geological setting 212
hanging wall alteration 167
isotopic data 248
ore genesis 213
Hercules deposit see also Mount Read province
alteration halo 178
explained 12, 128
geological setting 194
Highway-Reward deposit see also Mount Windsor
Subprovince
alteration facies and zonation 14, 167, 178, 232,
233-40
explained 14
geological setting 12-14, 232
hanging wall alteration 167,178
ore genesis 232
submarine facies associations 13
Hokoroku Basin see Hokuroku district
Hokuroku district
alteration 64-5, 67, 118-27, 1 22-7
geological setting 11820
Green Tuff Belt 52, 64, 1 07, 118, 1 5 0, 151
Kuroko deposits see Kuroko deposits
oxygen isotopes 94-5, 246
size of VHMS deposits 164
hydration of volcanic glass 4-5, 98-102
hydrothermal alteration
boundaries between zones 64
chemical reactions 168
and diagenetic alteration 128
discharge zone 141-2,156
discriminating 1619
explained 4
halos 66-7, 164-74
intensity measures 32-3
intrusion-related 140-61 see also intrusions
metamorphic assemblages 174, 175t
plagioclase destruction 31, 167
recharge zone 141
subseafloor systems 1401
syntectonic 6
tectonic deformation 6
VHMS deposits 5-6, 163-240
zones 5-6, 66-8, 164-78, 243
hydrothermal convection 1, 94, 140-1, 1 40
hydrothermal fluid 67, 172-3
HyMap 246, 247
Iberian pyrite belt 90, 91, 142, 164, 165, 166, 1 6 6 , 174
ICP-AES (inductively coupled plasma atomic emission
spectrometry) 767
ICP-MS (inductively coupled plasma mass spectrometry
76-7
indices
alteration 26-30, 34, 73, 244-5
Alteration box plot see Alteration box plot
Alteration Index (AI) 30-2, 34, 169, 245
CCPI (chlorite-carbonate-pyrite index) 31, 34, 169,
245
molar proportion alteration 30
multi-component and normalised 26, 30
simple ratio 26
induration 1 5 0, 151
infill textures 41 , 48-9
intrusions
halos 667
cryptodomes 2, 6, 66, 128, 139, 143, 153, 212, 232
dykes 2, 6, 66, 141, 139, 149, 152, 153
plutons 139, 143, 153
sills 66, 70, 100, 128, 148-9, 152-3, 1 5 0, 152, 174,
182, 194, 202, 212, 221, 232
in submarine volcanic successions 2, 3, 139
synvolcanic 13962
isotope geochemistry
applications 92-5
carbon 76, 77, 248
exploration 94-5, 246-50
hydrogen 76, 77, 93
oxygen 94-5, 246,248-50
stable isotopes 92-5
sulfur 76, 77-8, 92, 141, 180, 246, 248
water-rock ratios 93
K
kaolinite in VHMS altered zones 88, 150, 174-75, 178-80
keratophyre 98
K-lens see Rosebery deposit
Kuroko deposits see also Hokuroku district
alteration 164, 179
Alteration Index (AI) 31
alteration model 166, 1789
least altered see also alteration intensity
explained 26
alteration indices 32
lithification 97, 108-10
lithogeochemistry
analytical methods 73-8
carbonates 76-7
chemostratigraphy 79-81
C-H-N elemental analyser 76
closure 78-9, 81
compatible elements 7980
2 7 4 I INDEX
europium 87
explained 73-8, 74t
and exploration 243-5
hydrous minerals 76-7
ICP-AES (inductively coupled plasma atomic emission
spectrometry) 76-7
ICP-MS (inductively coupled plasma mass
spectrometry 76-7
immobile elements 79-81, 85, 87
inaccuracies in 77
incompatible elements 79
limit of detection 75
LOI (loss on ignition) 76
mass change 73, 81-7, 85-6, 87, 97, 165, 180-1,
245-6
NAA (neutron activated analysis) 76
precision and accuracy required 75
recalculating to volatile free 77-8
REE (rare earth elements) 73, 79, 81, 87
reporting data 77
sampling methods 738
summing elements 77
use of reference materials 75
XRF (X-ray fluorescence spectrometry) 76-7
M
mass change see lithogeochemistry
massive sulfide 1635, 167
Mattabi deposit 91, 180
metamorphism
burial metamorphism and diagenesis 16, 97, 98, 102,
114, 115-17
contact metamorphism 5, 11, 12, 64, 66-7, 149-54,
242
explained 4, 5, 20, 24
regional metamorphism 5, 64-6, 115, 139, 140-8
of VHMS-related altered zones 174-5
metasomatic alteration 4, 5, 144-6
microanalysis 24
microprobe see analytical techniques
mineral assemblages
alteration assemblages 235, 34, 165
burial effects on 97, 1 09
common assemblages 21t, 22t, 1 09, 165
disequilibrium 23-4, 63
equilibrium 23-4
in exploration 243
igneous 16, 19
isotopic studies of 925
nomenclature 19-20
mineral chemistry 87-91, 245-6
minerals defined 87-8
Mount Lyell field see also Mount Read province
deposits 12,202,243
geological setting 202
halos 202
hanging wall alteration 167
HyMap system 246, 247
mineral zonation 243, 247
ore genesis 203
Western Tharsis deposit see Western Tharsis deposit
Mount Read province
alteration 7-12, 163-164
Chester deposit 90
history 9, 11-12
Hellyer deposit see Hellyer deposit
Henty deposit see Henty deposit
Hercules deposit see Hercules deposit
Mount Lyell field see Mount Lyell field
Mount Read Volcanics see Mount Read Volcanics
oxygen isotopic exploration 24950
Que River deposit see Que River deposit
Rosebery deposit see Rosebery deposit
size of VHMS deposits 164
Western Tharsis deposit see Western Tharsis deposit
Mount Read Volcanics see also Mount Read province
AI and CCPI ranges 32,34
alteration 128-32, 1 33-8
Central Volcanic Complex 9-10, 6 9, 128, 1 30-32,
1 5 7-6 2
chemostratigraphic discrimination and correlation 80
compaction effects at 110
Darwin Granite 154-6, 157-62
geology of 7-12, 128, 1 29
Kershaw Pumice Formation 128,1346
Mount Black Formation 128, 1 33, 1 37 -8
pyritic alteration systems 89
metamorphic assemblages 11
Sterling Valley Volcanics 128
Mount Windsor Subprovince
alteration 14, 164, 222, 232
geology 12-14, 221-40
Highway-Reward deposit see Highway-Reward deposit
Thalanga deposit see Thalanga deposit
N
NAA (neutron activated analysis) 76 see also analytical
techniques
naming altered rock see alteration nomenclature
Noranda district 66, 142-7, 167, 179-81
numerical fluid-flow modelling 172
o
overprinting
textures 37, 70-1, 70t, 711
and false/pseudo textures 37, 54-63, 5 8-6 1 , 62t
relationships 69-71, 7It
palagonite 99, 99-100
Panorama district 180, 246, 248-50
paragenetic sequence 69
perlite 37, 40, 54, 100-1, 1 01
PIMA (portable infrared mineral analyser) 25, 33, 245 see
also analytical techniques
plagioclase destruction 31, 167-9
pseudotextures 16, 37, 54-63, 5 8-6 1 , 62t, 63
pyrophyllite in VHMS systems 174, 202, 207-8
Que River deposit 11-12, 70, 90, 167 see also Mount Read
province
INDEX I 2 7 5
R
regional metamorphism see metamorphism
relict textures 14, 16, 19, 24, 25, 37-38, 141
replacement textures 37-38, 41, 42-7
Rio Tinto deposit see Iberian pyrite belt
Rosebery deposit see also Mount Read province
alteration 70, 128, 131t, 168, 173, 178, 195, 196-
201
Chlorite-carbonate-pyrite index (CCPI) 195
geochemical alteration parameters 168
geological setting 194
explained 11-12, 194-201
hydrothermal carbonates 91
hydrothermal fluid flow 174
K-lens 30, 34, 68, 91, 174, 244
lithogeochemical data in exploration 82, 244
sericite 70-1
Scuddles deposit 180
sericite alteration 20-2, 70-1, 165-78, 185-6, 198-9, 206-
11, 218,225, 231, 238 see also white mica
Short-wavelength infrared spectroscopy (SWIR) see
analytical techniques
siliceous alteration 165-9, 182, 189, 199, 219-20
smectites 102, 109, 117
Snow Lake District 142-6
sodium content in volcanic rocks 346, 2434
solution seams 51, 52
spilite 98
stable isotopes see isotope geochemistry
stockwork see stringer zones
stringer zones 163, 165, 172-3, 179
stylolites 51, 52
submarine environments 2, 97138
submarine facies associations
Mount Read Volcanics 10-11, 214-20
Seventy Mile Range Group 12-13
submarine volcanic successions 1-14, 3, 97-138, 241-9
SWIR spectroscopy see analytical techniques
synvolcanic intrusions see intrusions
Thalanga deposit see also Mount Windsor Subprovince
alteration facies and zonation 14, 178, 222, 223-31
explained 14
geological setting 1214, 2212
mass change estimations 856
ore genesis 222
oxygen isotopic exploration 250
thermodynamic alteration model 170-2
vectors
exploration 24350
isotopic 24650
lithogeochemical 243-4
mass change 245
mineral chemistry 79, 89-91, 245-6
sulfide 243-4
vein-halo alteration 38, 41, 43, 69
VHMS deposits
alteration patterns 16478
classification 1636
common features 1634
comparisons 178-81
exploration 241-50
footwall alteration 163-74
halos 5-6, 67-9, 68, 163-8, 175, 178-81
hanging wall alteration 1634, 1678
Hellyer deposit see Hellyer deposit
Henty deposit see Henty deposit
Highway-Reward deposit see Highway-Reward deposit
kaolinite, presence of 174
major VHMS provinces 164
Mount Lyell see Mount Lyell field
pyrite in 243
pyrophyllite, presence of 174
regional alteration zones 5-6, 139-62
Rosebery deposit see Rosebery deposit
Thalanga deposit see Thalanga deposit
Western Tharsis deposit see Western Tharsis deposit
VMS deposits see VHMS deposits
volcanic facies
alteration processes 26
alteration in submarine environments 97138
associations 2
changes in 108
clastic facies 1,19
coherent facies 1, 6, 16, 19
common clay minerals in 102
crystalline facies 6
volcaniclastic facies 1, 6, 12, 1619
volcanic-hosted massive sulphide deposits see VHMS
deposits
vugs 41, 50- 1, 52
W
water-rock ratios 934
Western Tharsis deposit see also Mount Read province see
also Mt Lyell field
alteration facies and zonation 2023, 20411
case study 202-11
geological setting 202
ore genesis 203
white mica 91
white mica 20, 89-91, 165, 246
Woodlawn deposit 178
X
X-ray diffraction 19, 24, 25, 33, 88
X-ray fluorescence spectrometry (XRF) 767
zeolites 105-6, 106t, 110, 114-16, 118, 120-1
zones see alteration zonation
About the authors
Dr Cathryn Gifkins is a Research Fellow at the Centre for Ore Deposit Research at the University
of Tasmania. Cathryn brings to the publication a strong background in mapping, describing and
interpreting altered and deformed volcanic rocks in submarine successions. Her current research
focusses on the textural, mineralogical and compositional effects of alteration in glassy volcanic
rocks, the link between volcanic centres and mineralising hydrothermal systems, and the facies
architecture and stratigraphy of the Mount Read Volcanics. '
Walter Herrmann is a Research Fellow in economic geology at the Centre for Ore Deposit
Research. Wally's background in mineral exploration in Australian volcanic successions,
principally the Mount Read Volcanics and the Mount Windsor Subprovince is a valuable asset
to the book. He has a special interest in understanding hydrothermal alteration as a method for
discriminating and discovering VHMS and porphyry deposits.
Professor Ross Large is Director of the Centre for Ore Deposit Research and has a long and
celebrated academic and exploration career. Ross has a comprehensive knowledge of VHMS
deposits, and has actively promoted and developed the application of geochemical techniques to
mineral exploration. This innovative approach has recently produced the Alteration box plot, an
alternative way to relate alteration intensity, mineralogy and geochemistry.

You might also like