Download as pdf or txt
Download as pdf or txt
You are on page 1of 329

Evolution of hydrothermal

alteration facies at the


Cerro Corona Cu-Au
porphyry deposit,
Northern Peru
by

Jacob Longridge

Submitted in fulfilment of the requirements for


the degree of Doctor of Philosophy

Imperial College London

Department of Earth Science and Engineering

Submitted: November 2016


Declaration of originality

This thesis comprises of work that is my own except where due referencing and
acknowledgement is made.

Copyright Declaration

The copyright of this thesis rests with the author and is made available under a
Creative Commons Attribution Non-Commercial No Derivatives licence. Researchers
are free to copy, distribute or transmit the thesis on the condition that they attribute it,
that they do not use it for commercial purposes and that they do not alter, transform
or build upon it. For any reuse or redistribution, researchers must make clear to others
the licence terms of this work.

ii
Abstract
The Cerro Corona deposit is an active copper gold porphyry deposit, located in the
Hualgayoc District, Northern Peru. Its magmatic evolution is conformable with the current
regional model of crustal thickening, flat slab subduction and uplift. The deposit comprises
of multiple mineralogically and chemically indistinguishable biotite quartz diorite
porphyry intrusions, followed by intrusions of hornblende granodiorite and andesite
dykes. Igneous amphibole chemistry of these later intrusions suggests an increasing input
of deeper, mafic hydrous melts into a shallower magma chamber, similar to the paragenesis
at the nearby Yanacocha district.

Hydrothermal alteration is similar to other porphyry-type deposits, however the argillic


alteration telescopes onto the entire deposit, with pervasive argillic alteration extending to
a depth of 400m from the current surface. This suggests a significant uplift during
hydrothermal alteration.

Error is propagated through the mass transfer calculations, allowing for a more confident
interpretation of mass transfer models. These models show large net transfer of major and
minor components during potassic, propylitic and sericitic alteration, but less drastic
changes during argillic alteration. A 3D net mass transfer models is done for the first time
on a porphyry deposit, which reveals a previously unidentified zone of potassic alteration
in the north.

Infrared spectroscopy revealed the dominant clay mineral is smectite, with mixture of illite-
and/or kaolinite. Kaolinite-dominant argillic alteration appears to be related to sulphide
oxidation near the surface and telescopes down along preferentially permeable zones,
whereas smectite-dominant argillic alteration occurs at near neutral conditions, deeper
and peripheral to zones of sulphide mineralisation.

Intracrystalline oxygen isotope fractionation indicates illite formed at temperatures of


~300°C and kaolinite >150°C. The calculated fluid compositions suggests illite-formation
fluids are composed of >70% magmatic-derived fluids; kaolinite-formational fluids being
an equal mix of meteoric and magmatic fluids and smectite-formational fluid being
predominantly heated meteoric fluid.

iii
Acknowledgements
I would like to thank NERC and Goldfields La Cima for providing the funding that allowed
for this research, under the Collaborative awards in science and engineering studentship
(NERC CASE no.: NE/I018409/1).

Thank you Andy Wurst and Jamie Wilkinson for setting up this project, ensuring that I could
hit the ground running, helping to keep me on track and patiently reviewing my work.

Thank you to the Gold Fields team at Cerro Corona, especially Eddie Garcia, Angel Uzategui,
Edwin Ayala, Luis Armas, Lucia Torres, Roberto Rocha and Regina Baumgartner for your
help-from organising the logistics of my visits to patiently reviewing my reports.

Thank you to the staff at NHM, especially Jens Najorka for improving my understanding of
the intricacies involved in XRD. Thank you to the staff in the Imaging and Analysis Centre
of the NHM, especially Anton Kearsley and John Spratt for your unrivalled assistance in
answering all my questions related to electron microscopy.

At SUERC, I would like to thank Alison McDonald and Terry Donnelly for teaching me the
fundamentals of O and H isotope geochemistry. Thank you to Adrian Boyce for allowing me
the freedom to try new techniques and patiently editing and reviewing my isotope work.

I would like to thank my family for their strong encouragement that helped keep me going
through the tough days. To my friends and colleagues at Imperial College and the NHM -
cheers for keeping my enthusiasm in check-Will, Ed, Thilo, Jorge, Irina, Matt, Rob, Carmen,
Rachael, Adam, Simon, Han, Sarah and Yvette.

Thank you to my family as well as the Sykes family! Your patience, support, humour, roof
over my head and great craft beer helped get me through the darkest days of this write up.

I am indebted to the unwavering love and support from Meredith. I could not have done
this without you, thank you!

iv
Table of Contents
1 Chapter One: Literature review ........................................................................................................... 1

1.1 Porphyry ore-forming processes ............................................................................................... 1

1.1.1 Tectonic setting of porphyry deposits ........................................................................... 1

1.1.2 Metal zoning in porphyry systems................................................................................... 5

1.1.3 Alteration zoning in porphyry systems ......................................................................... 6

1.2 Techniques used in this thesis ................................................................................................. 11

1.2.1 Infrared spectroscopy ........................................................................................................ 11

1.2.2 Stable isotopes ...................................................................................................................... 14

1.3 References ........................................................................................................................................ 15

2 Chapter Two: Introduction.................................................................................................................. 19

2.1 Generic model for porphyry copper formation ................................................................ 19

2.2 The importance of argillic alteration..................................................................................... 21

2.3 Case study of the Cerro Corona deposit ............................................................................... 22

2.4 References ........................................................................................................................................ 26

3 Chapter Three: Regional Geology ..................................................................................................... 29

3.1 Introduction ..................................................................................................................................... 29

3.2 Geological setting of the Cajamarca district ....................................................................... 32

3.2.1 Stratigraphy ........................................................................................................................... 32

3.2.2 Principal structures and tectonic development ...................................................... 35

3.3 Igneous geochemistry of Cajamarca District...................................................................... 42

3.4 Uplift and erosion .......................................................................................................................... 50

3.5 Discussion ......................................................................................................................................... 50

3.6 References ........................................................................................................................................ 51

4 Chapter Four: Geology of Cerro Corona ......................................................................................... 58

4.1 Introduction ..................................................................................................................................... 58

4.1.1 The Hualgayoc mining district ....................................................................................... 58

4.1.2 Local stratigraphy and intrusive ages ......................................................................... 58

4.1.3 Structural geology ............................................................................................................... 65

v
4.1.4 The Cerro Corona intrusive complex ........................................................................... 66

4.1.5 Hydrothermal alteration ................................................................................................... 68

4.2 New insights into the intrusive history at Cerro Corona .............................................. 70

4.2.1 Petrography............................................................................................................................ 72

4.2.2 Mineral chemistry................................................................................................................ 74

4.2.2.1 Feldspar chemistry.............................................................................................................. 74

4.2.3 Geochemistry of igneous rocks ...................................................................................... 81

4.2.4 General alteration facies ................................................................................................... 86

4.2.5 Vein paragenesis ................................................................................................................ 101

4.2.6 Geological modelling of Cerro Corona....................................................................... 105

4.3 Summary ......................................................................................................................................... 110

4.4 References ...................................................................................................................................... 115

5 Chapter Five: Alteration geochemistry ........................................................................................ 119

5.1 Introduction ................................................................................................................................... 119

5.2 Methods ........................................................................................................................................... 119

5.2.1 Geochemistry ....................................................................................................................... 119

5.2.2 Sample selection for geochemical mass transfer analyses ............................... 119

5.2.3 The isocon method ............................................................................................................ 125

5.2.4 Error estimates ................................................................................................................... 128

5.3 Results ............................................................................................................................................. 129

5.3.1 Potassic alteration ............................................................................................................. 130

5.3.1.2 High Field Strength Elements ....................................................................................... 131

5.3.1.3 Metals...................................................................................................................................... 135

5.3.2 Propylitic alteration .......................................................................................................... 137

5.3.2.1 High Field Strength Elements ....................................................................................... 140

5.3.3 Sericitic alteration ............................................................................................................. 145

5.3.3.1 High Field Strength Elements ....................................................................................... 148

5.3.4 Argillic alteration ............................................................................................................... 153

5.3.4.1 High Field Strength Elements ....................................................................................... 155

5.3.5 Model of chemical mass transfer ................................................................................. 159

vi
5.4 References ...................................................................................................................................... 182

6 Chapter Six: Infrared spectroscopy of clay minerals .............................................................. 183

6.1 Introduction ................................................................................................................................... 183

6.2 Methods ........................................................................................................................................... 183

6.2.1 IR spectroscopy .................................................................................................................. 183

6.3 Results .............................................................................................................................................. 196

6.3.1 Major minerals in infrared spectroscopy................................................................. 199

6.3.2 Mineral variance................................................................................................................. 199

6.3.3 Future work.......................................................................................................................... 219

6.4 References ...................................................................................................................................... 219

7 Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration ............. 223

7.1 Introduction ................................................................................................................................... 223

7.2 Methods ........................................................................................................................................... 223

7.2.1 Clay mineral separation and X-ray Diffraction analysis .................................... 223

7.2.2 Sample homogeneity ........................................................................................................ 227

7.2.3 Single mineral dehydroxylation temperatures ..................................................... 228

7.2.4 Oxygen isotope analysis .................................................................................................. 229

7.2.5 Hydrogen isotope analysis ............................................................................................. 232

7.2.6 Local meteoric waters ...................................................................................................... 234

7.3 Results .............................................................................................................................................. 234

7.3.1 Bulk mineral oxygen isotope results ......................................................................... 234

7.3.2 De-hydroxylated oxygen isotope results ................................................................. 235

7.3.3 Hydrogen isotope results ............................................................................................... 240

7.3.4 Local meteoric waters ...................................................................................................... 241

7.4 Discussion ....................................................................................................................................... 242

7.4.1 Estimation of clay formation temperatures ............................................................ 242

7.4.2 Calculation of fluid temperatures and compositions .......................................... 245

7.4.3 Modelling of exchanged meteoric and magmatic waters and mixing curves
250

7.4.4 Results of modelling exchanged fluids ...................................................................... 252

vii
7.5 Summary ......................................................................................................................................... 260

7.6 References ...................................................................................................................................... 261

8 Chapter Eight: Discussion .................................................................................................................. 265

8.1 Geodynamic setting of Cerro Corona and source magmas ......................................... 265

8.2 Mineralogical controls of chemical mass transfer effects ........................................... 267

8.2.1 Potassic alteration ............................................................................................................. 269

8.2.2 Propylitic alteration .......................................................................................................... 272

8.2.3 Sericitic alteration ............................................................................................................. 273

8.2.4 Argillic alteration ............................................................................................................... 276

8.2.5 3D Mass transfer modelling........................................................................................... 279

8.3 The origin of clay alteration at Cerro Corona .................................................................. 283

8.4 Fluid flow model for the genesis of argillic alteration at Cerro Corona ................ 285

8.5 References ...................................................................................................................................... 285

9 Chapter Nine: Conclusions ................................................................................................................ 289

9.1 Development of mineral-bearing arc volcanism in northern Peru......................... 289

9.2 Implications from mass transfer models of hydrothermal alteration in a porphyry


deposit ............................................................................................................................................................ 290

9.3 Inferences from stable isotope and infrared spectroscopy of clay minerals ...... 293

9.4 Future work ................................................................................................................................... 296

9.4.1 Regional and local geology ............................................................................................. 296

9.4.2 Alteration geochemistry ................................................................................................. 297

9.4.3 Infrared spectroscopy ...................................................................................................... 297

9.4.4 Oxygen and hydrogen isotopes .................................................................................... 297

9.4.5 Generic model...................................................................................................................... 298

9.5 References ...................................................................................................................................... 298

viii
List of Figures
Figure 1.1: Global distribution of porphyry deposits (USGS, 2017) and subduction zones. 1
Figure 1.2: Tectono-magmatic model for the emplacement of porphyry deposits (Richards,
2003). Hot zones at the base of the lower crust cause initial homogenisation of magmas (as
described in the text). Within the shallow crust, magma ascent is controlled by local
stresses. The magma rises to a state of neutral buoyancy, where it pools to form a
subvolcanic pluton. Note the change in scale in the upper crust. .................................................... 2
Figure 1.3: Fluid composition versus temperature diagram illustrating that under full
equilibrium CO2 becomes the major control of fluid acidity below ~350°C (Giggenbach
1997)......................................................................................................................................................................... 4
Figure 1.4: Schematic vertical cross section of metal distribution in porphyry deposits
(Halley et al. 2015). ............................................................................................................................................. 5
Figure 1.5: Idealised distribution of alteration assemblages associated with a porphyry
deposit (Seedorff et al. 2003). ........................................................................................................................ 6
Figure 1.6: Phase diagram for the system K2O-Al2O3-SiO2-H2O-KCl-HCl at atmospheric
pressure. The labelled areas broadly correspond to alteration assemblages. Mineral
abbreviations are: Cord-cordierite; And-andalusite; Ksp-potassium feldspar; Bi-biotite;
Chl-chlorite; Pyro-pyrophyllite; Musc-muscovite; Kaol-kaolinite; Ill-illite; Mont-
montmorillonite. Note that chlorite, cordierite and biotite stability fields are based on a
simple solid solution model for their Mg end members (Seedorff et al. 2005). ........................ 7
Figure 1.7: The stability of aluminosilicates is primarily a function of mHCl/mKCl and
temperature. The red arrow shows the path that a vapour phase released from a degassing
magma is likely to take (i.e. low P degassing) (Giggenbach, 1997). ............................................... 8
Figure 1.8: Fluid pathway for magmatic liquids and vapours involved in the formation of
argillic alteration (Hedenquist et al. 1998) projected onto a phase diagram showing the
stability of minerals in the K2O-Na2O-Al2O3-SiO2-HCl system (Sverjensky et al. 1991).
Dashed grey lines are approximate boundaries between clay minerals. Abbreviations for
minerals are: And=andalusite, Pyro(ph)=pyrophyllite, Dick=dickite, Kaol=kaolinite,
Musc=muscovite, Ill=illite, Smec=smectite and Ksp=potassium feldspar................................. 10
Figure 1.9: An example of the “vuggy quartz” from the Cerro Corona deposit. Note the rock
is dominated by dark silica replacement. ............................................................................................... 10
Figure 1.10: Location of absorption features commonly found that are related to clay
minerals. These absorption features are due to the combination of symmetrical stretching
(v1), H-O-H bending (v2) and asymmetrical OH-stretching (v3) of the water molecule
(shown in blue) and the combination of hydroxyl stretching (vOH in yellow) with metal (Al,
Fe or Mg) hydroxyl bending (δXOH in peach). ..................................................................................... 12

ix
Figure 1.11: An example of physical mixtures between dickite and kaolinite in 20 wt.%
increments (this study). The ~1380 nm absorption feature of dickite is more intense than
that of the equivalent ~1395 nm absorption feature of kaolinite. Thus, when using a linear
mixing algorithm for quantification, the dickite proportion is overestimated. ...................... 13
Figure 2.1: Schematic of the typical hydrothermal alteration pattern encountered in
porphyry-type ore deposits. These deposits have large alteration footprints; however, the
distribution of each type may vary significantly in extent and form from deposit to deposit.
From Sillitoe (2010). ....................................................................................................................................... 20
Figure 2.2: Percentage of global copper production in relation to classified resource type
(Gerst 2008). ....................................................................................................................................................... 21
Figure 2.3: Copper recovery from porphyry ores with increasing clay content. Adapted from
Bulatovic et al. (1998)..................................................................................................................................... 22
Figure 2.4: Maps of polygons coloured by: A) clay content; and B) copper recovery.
Reproduced from quality control maps (Uzategui and Ayala 2012)........................................... 24
Figure 3.1 Northern Peruvian margin during the late Cretaceous. The Western Peruvian
Trough is defined by the Paracas High to the West and Maranon Geanticline to the East. The
Normal faults controlled sedimentation within this miogeocline. These faults were
subsequently reactivated during subsequent orogenic events. Adapted from Scherrenberg
et al. 2012. ............................................................................................................................................................ 29
Figure 3.2: Geological terrains of Northern Peru. The Cajamarca region lies west of the
Marañon Fold Thrust Belt (MFTB), within the Western Cordillera. The Cajamarca curvature
is defined by the change of the trend of predominant fold axial planes from northwest to a
west-northwest. The Huancabamba deflection is a large-scale change in fault trend from
NW (in Peru) to NE (in Ecuador and Colombia). The inset box (expanded in Fig. 3.5) shows
the area of study. Adapted from Mégard (1984). ................................................................................ 31
Figure 3.3: Schematic stratigraphic column for the Cajamarca district. Cretaceous
stratigraphy adopted from Wilson (1984). Cenozoic volcanism is incorporated from Chavez
et al. (2012). Decollement horizons are inferred from Eude (2014) .......................................... 34
Figure 3.4: Average convergence rate between the Nazca and South American plate over
time and approximate plate motion, based on plate reconstruction (Pardo-Casas and
Molnar 1987). Note the periods of high convergence coincide with deformation stages. . 36
Figure 3.5: Map of the Cajamarca district. Maps used: Wilson 1984, Noble et al.1990,
Macfarlane et al. 1990, Davies 2002, Longo 2005, Navarro 2012 ................................................ 37
Figure 3.6: Sequence of magmatic events in the Cajamarca district. Adapted from Davies
2002........................................................................................................................................................................ 41

x
Figure 3.7: Cross section through subduction zone through the Northern Peruvian margin.
Note that location of Miocene magmatism is located on the margin of a thickened
continental crust. Adapted from Eude 2014.......................................................................................... 43
Figure 3.8: Major element vs. SiO2 variation diagrams for igneous rocks the Cajamarca
district. Data compiled from Davies 2002; Longo 2005; Chiaradia et al. 2009; Navarro et al.
2010; INGEMMET 2008; and this study. Colours correlate to magmatic pulses in figure 3.6,
whereby brown represents the Tantara (75-55 Ma); maroon represents the Llama (55-42
Ma); orange represents the Pativilca (42-30 Ma); light orange (gabbroic dykes) represent
the Calamarca (30-24 Ma); yellow, tan, pink represent the evolving Calipuy (24-10 Ma); and
green represent the Negritos magmatic pulse (10-3 Ma). ............................................................... 46
Figure 3.9: A) Sr/Y versus Y and B) La/Yb vs. Yb plots of Cenozoic igneous rocks in the
Cajamarca district. Note that the shift to adakite-like signature occurs around the time of
Calipuy Volcanism (i.e. ~24 Ma). Figure 3.8 contains key of symbols and list of references
for data. ................................................................................................................................................................. 47
Figure 3.10: A) Total Alkali versus silica diagram of regional igneous rocks and B) AFM
diagram indicating a linear calc-alkaline trend. Figure 3.8 contains figure key. .................... 49
Figure 4.1: Stratigraphy of the Hualgayoc district. Adapted from Macfarlane and Petersen
(1990) and Castillo (2001). .......................................................................................................................... 60
Figure 4.2: Regional map of NW-SE trending Hualgayoc anticline with Miocene intrusions
hosted within the hanging wall. Map adapted from Wilson 1984, Noble et al. 1990,
Macfarlane and Petersen 1990, Castillo 2001. ..................................................................................... 62
Figure 4.3: Sharp intrusive contacts between the biotite quartz diorite 1 intrusion and the
host limestone (A, B and E) along with interpretations (A’, B’ and E’). Chalcopyrite and
pyrite mineralisation occurs within silty limestone (C and D) with interpretations (C’ and
D’). ........................................................................................................................................................................... 63
Figure 4.4: Schematic model of intrusions at Cerro Corona. Sillitoe 1997 ............................... 66
Figure 4.5: Surficial geology map of the Cerro Corona Diorite intrusion. The mapped
“porphyry diorite” is QD1 intrusion and “Barren Core(s)” are QD2 intrusions, adapted from
Uzategui and Jacay 2010. ............................................................................................................................... 67
Figure 4.6: Representative examples of the intrusive contacts at Cerro Corona. .................. 71
Figure 4.7: Quartz-Alkali-feldspar-Plagioclase classification diagram with modal mineral
proportions. ........................................................................................................................................................ 72
Figure 4.8: Cross-polarised image showing biotite and altered plagioclase phenocrysts in a
fine-grained groundmass of predominantly quartz and feldspar. Scale bar represents 1mm.
.................................................................................................................................................................................. 73
Figure 4.9: An example of symplectite texture in sample C12.JL.48. The dominant
phenocrysts are plagioclase and quartz. The field of view is 2mm.............................................. 73

xi
Figure 4.10: An example of zoning within zircon (blue). Sample C12.JL.185. Green mineral
is apatite. .............................................................................................................................................................. 74
Figure 4.11: Back scattered images of plagioclase phenocrysts showing oscillatory zoning
with brighter zones having a more anorthitic composition. A) Zoned plagioclase phenocryst
with labradorite zone associated with apatite and magnetite. Scale bar is 400μm; B) The
andesine plagioclase phenocryst in in a quartzofeldspathic matrix showing brighter zones
of labradorite. The sample is from a post-mineralisation hornblende porphyry dyke. Scale
bar is 3mm; C) C12.186 Zoned andesine plagioclase phenocryst within andesite dyke.
Brighter zones are more anorthitic. Scale bar is 2mm ...................................................................... 75
Figure 4.12: Electron microprobe data of feldspar compositions in least altered porphyritic
rocks. Phenocrysts are predominantly andesine-labradorite in composition (blue circles),
whereas oligoclase composition feldspars occur within amphibole, and albite-orthoclase
compositions occur within biotite phenocrysts................................................................................... 76
Figure 4.13: Back scattered electron images of analysed feldspars. Red spots indicate points
of microprobe analysis. A) Poikilitic amphibole phenocrysts with inclusions of apatite,
magnetite and feldspar; B) Poikilitic biotite phenocryst with inclusions of apatite and
feldspar; C) Zoned plagioclase phenocryst; D) Poikilitic amphibole with inclusions of
magnetite, apatite and feldspar in a quartzofeldspathic matrix. Scale bar for all images is
500 μm................................................................................................................................................................... 76
Figure 4.14: Biotite phenocrysts from least altered biotite quartz diorites and post ore-
mineralisation andesite dykes illustrating variable mineral inclusions and textures. A)
Biotite phenocryst from post-mineralisation andesite dyke with inclusions of plagioclase,
apatite and magnetite. Scale bar 500 μm (sample C12.JL.186); B) Biotite phenocryst from
post-mineralisation andesite dyke Scale bar 200 μm (sample C12.JL.183); C) Biotite
phenocryst from hornblende quartz diorite porphyry with inclusions of plagioclase and
apatite. Scale bar is 500 μm (sample C12.JL.47), and; D) Biotite phenocryst being altered to
pycnochlorite from hornblende quartz diorite porphyry with inclusions of plagioclase and
apatite. Scale bar is 500 μm (sample C12.JL.48) .................................................................................. 77
Figure 4.15: Sample C11.178. Back scattered electron images of poikilitic magmatic biotite
with inclusions of A) alkali feldspar (K0.65Na0.35AlSi3O8) with two magnetite inclusions. Scale
bar is 300μm; B) Rutile (darker) exsolution from ilmenite(light) within biotite phenocryst.
Scale bar is 100 μm, and; C) apatite, k-spar (K0.84Na0.16AlSi3O8) magnetite and albite
(Na0.96K0.02Ca0.02AlSi3O8). Scale bar is 300μm .......................................................................................... 78
Figure 4.16: Biotite mineral chemistry projected onto phlogopite-annite-siderophyllite-
eastonite quadrilateral (Speer 1984). The circles represent biotite chemistry from the
hornblende-bearing quartz diorite porphyry and diamonds represent biotite chemistry
from post-mineralisation andesite dykes. .............................................................................................. 78

xii
Figure 4.17: Examples of quartz eyes at Cerro Corona. A) Large quartz eye phenocryst in
argillically altered (kaolinite-illite) quartz biotite diorite (GFD 314, 1337m). Scale bar is 1
cm; B) Cold cathodoluminescence electron image showing the oscillatory zoning in quartz
eye. Scale bar is 1mm, and; C) Sub-rounded quartz eye with a strongly developed
embayment in a fine quartzofeldspathic groundmass. C11.55, GFD 175, 32 m. Scale is 1mm.
.................................................................................................................................................................................. 79
Figure 4.18: Back scattered images of amphiboles from post- (A and B) and syn- (C and D)
mineralisation intrusions demonstrating textures and inclusions. Red spots are location of
electron microprobe analyses. A) Pargasite amphibole with anhydrite and magnetite
inclusions (sample C12.JL.183). Scale bar 200μm; B) Pargasite amphibole with anhydrite
inclusion. (sample C12.JL.186). Scale bar 200μm; C) magnesio-hornblende with abundant
anhydrite, magnetite and feldspar inclusions. (sample C12.JL.180). Scale bar is 500 μm, and;
D) Magnesio -hornblende with abundant anhydrite, magnetite and feldspar inclusions.
(sample C12.JL.47). Scale bar is 500 μm. ................................................................................................ 80
Figure 4.19: Classification diagram of calcium amphibole subgroup, where
B(Ca+∑M2+)/∑B≥0.75 and BCa/∑B ≥ B∑M2+/∑B. Amphiboles from syn-mineralisation
hornblende-quartz diorite porphyry(circles)are predominantly magnesio-hornblende
whereas amphibole from post-mineralisation andesite dykes (diamonds) are pargasite
composition......................................................................................................................................................... 81
Figure 4.20: Plan section through the Cerro corona deposit at ~3600 m elevation
illustrating the sample locations for the least altered biotite diorites. Note samples occur
within the “Western Barren Core” (QD 2) as well as the initial biotite quartz diorite (QD 1).
.................................................................................................................................................................................. 82
Figure 4.21. Total Alkali versus Silica classification (Le Bas et al. 1991) and tholeiitic vs.
calc-alkalic trend plots (Kuno 1968). The green symbols are samples from the early quartz-
biotite diorite (QD1); purple symbols represent samples from the “Western barren core”
(QD2) and black rectangles represent samples from the Las Gordas rhyolite. Brown
symbols represent previously analysed QD1 and QD2 samples (James 1998). Error bars
represent the average analytical errors of analysed samples. ....................................................... 83
Figure 4.22. Zr/Ti vs. Nb/Y discrimination diagram (after Winchester and Floyd 1977). The
key used in this figure is the same as Fig. 4.21. Error bars represent the average analytical
errors of analysed samples. Notice that samples straddle multiple fields and are within
error of one another. ....................................................................................................................................... 84
Figure 4.23: Harker diagrams illustrating the composition of the Cerro Corona intrusions
relative to other igneous rocks within the Cajamarca district. Key from Figure 4.22. Error
bars represent the average analytical errors of analysed samples. Red lines distinguish
high- med- and low-K ...................................................................................................................................... 85

xiii
Figure 4.24: Examples of potassically altered samples. Plagioclase is readily altered to
alkali-feldspar, hornblende is altered to shreddy biotite and there is a close association with
magnetite ± chalcopyrite ± bornite mineralisation. ........................................................................... 87
Figure 4.25: Plan section across the 3810 m level illustrating the logged distribution of K-
feldspar. Purple line represents open pit limit. .................................................................................... 88
Figure 4.26: An example of the gradational contact between intense potassic alteration and
intense propylitic alteration seen at depth (sample C12.JL.220; GFD-314, 1221 m). Note
potassic alteration (K-feldspar and biotite) is associated with anhydrite, fluorite and
quartz. Biotite is altered to chlorite and quartz. .................................................................................. 90
Figure 4.27: Plan section across the 3810 m level showing the distribution of logged
chlorite. ................................................................................................................................................................. 91
Figure 4.28: Hand samples of sericitised biotite quartz diorite. A) Example of intense
sericitisation of plagioclase to white sericite and biotite to grey sericite. B) Chalcopyrite
veinlet with k-feldspar halo grading into sericitic alteration of plagioclase. Scale bar for
both images is 1cm. .......................................................................................................................................... 92
Figure 4.29: An example of a planar quartz-pyrite vein associated with sericite alteration
(sample C12.JL.46). A) Sample showing the alteration is confined to the vein selvage and
predominantly alters feldspar. B) Backscattered electron image of the vein and C) the vein
selvage. Note the sericitisation alters both plagioclase and k-feldspar. Mineralogy was
confirmed by EDX analyses. ......................................................................................................................... 93
Figure 4.30: A) Hand sample of showing vein with associated sericitisation of plagioclase
phenocrysts (C12.JL.170; GFD-314; 808 m).B) Back scattered electron image of relic
plagioclase phenocryst and C) higher magnified area showing relic albite being altered to
sericite and calcite. ........................................................................................................................................... 94
Figure 4.31: Distribution of muscovite across the 3810 m level at Cerro Corona. ................ 95
Figure 4.32: Smectite-dominant argillic alteration overprinting potassic altered biotite
quartz diorite (sample C12.JL.141, GFD-314, 610 m). This smectite is associated with
zeolite-calcite vein which crosscuts quartz-magnetite-anhydrite ± chalcopyrite vein. ...... 96
Figure 4.33: Plan section across the 3810 m level at Cerro Corona showing the distribution
of smectite, as determined by IR spectroscopy. ................................................................................... 97
Figure 4.34: Back scattered image of illitisation of plagioclase phenocrysts and groundmass
sample (C12.JL.210 (GFD 314, 1142 m). ................................................................................................. 98
Figure 4.35: Plan section across the 3810 m level showing illite distribution, as determined
by IR spectroscopy ........................................................................................................................................... 99
Figure 4.36: Kaolinite distribution across the 3810 m level as determined by IR
spectroscopy. .................................................................................................................................................... 100

xiv
Figure 4.37 Representative examples of crosscutting relationships at Cerro Corona and
their association to alteration.................................................................................................................... 101
Figure 4.38: Sample C12.JL. 161. False colour montage of quartz-magnetite-chalcopyrite
vein associated with potassic alteration. Black is predominantly quartz. Minerals are
confirmed by EDX analyses. Bright red is magnetite, darker red is chalcopyrite, and green
is albite plagioclase and blue is alkali-feldspar. Note that the matrix is almost completely
altered to k-feldspar with the larger phenocrysts being relatively unaltered. ..................... 102
Figure 4.39: K-feldspar open space vein with minor infill of chalcopyrite and molybdenite.
Sample C12.JL.168. GFD 314, 780m ........................................................................................................ 102
Figure 4.40: Summary of paragenesis of sample C12.JL.46. A) Hand sample shows
apparently simple quartz-pyrite vein with an associated sericite alteration selvage. Red box
illustrates location of inset E). B) Back scattered electron image of vein illustrating a
sequence of quartz - calcite ± chalcopyrite ± pyrite ± fluorite; C) Montage of SEM-CL image
of vein illustrating multiple generations of quartz growth and re-absorption; D)
interpretation of images B and C. E) Example of SEM-CL image of quartz illustrating
multiple stages of crustiform quartz growth. ..................................................................................... 104
Figure 4.41: Calcium content of rocks at Cerro Corona, based on assay data. Note the drastic
change of slope from ~4 wt.%. Values >4 wt.% CaO are interpreted to be within Cretaceous
limestone. ........................................................................................................................................................... 105
Figure 4.42: Plan showing molybdenum grade shells across the Corona Diorite at 3810 m
a.s.l......................................................................................................................................................................... 107
Figure 4.43: Plan showing copper grade shells across the Corona Diorite at 3810 m a.s.l.
The areas of high grade are concentrically arranged around the “Barren Cores”: BCE -
Barren Core East; BCW - Barren Core West......................................................................................... 108
Figure 4.44: Map illustrating distribution of gold grades across the Cerro Corona deposit at
3810 m a.s.l. ....................................................................................................................................................... 109
Figure 4.45: log fO2 versus temperature diagram with calculated amphibole temperatures
and oxygen fugacities (Ridolfi et al. 2010) from hornblende-quartz diorite porphyries
(circles) and post-mineralisation andesite (diamonds). MH = magnetite-hematite buffer,
FMQ = fayalite-magnetite-quartz buffer (Ming Chou 1978). Magnesio-hornblendes formed
at lower temperature and relatively higher oxygen fugacity than pargasitic amphibole from
late-stage andesitic dykes. .......................................................................................................................... 111
Figure 4.46: Al/(Ca+Na+K) vs. anorthite percentage for feldspar crystals in biotite-quartz
diorite porphyry. Note that phenocrysts appear to be more anorthitic than feldspar
inclusions in amphibole and biotite. Line joins ideal albite and anorthite end-members.
................................................................................................................................................................................ 112

xv
Figure 4.47: Pressure versus temperature of syn-mineralisation hornblende quartz diorite
(circles) and post-mineralisation andesite dykes (diamonds) as calculated from amphibole
thermobarometry (Ridolfi et al. 2010). ................................................................................................. 113
Figure 4.48: H2O of melt versus temperature as determined from amphibole hygrometry
and thermometry of amphiboles (Ridolfi et al. 2010) from syn-mineralisation hornblende-
quartz diorites (circles) and post-mineralisation andesite dykes (diamonds). ................... 114
Figure 5.1: Hand samples of least-altered biotite-quartz diorite used for mass transfer
calculations. ....................................................................................................................................................... 120
Figure 5.2: Map of Cerro Corona showing the location of the least-altered samples (black
dots). Orange lines are the inferred location of major faults which continue through host
limestone. Western and eastern meshes delineate the post-mineralisation, biotite-quartz
diorite porphyries........................................................................................................................................... 121
Figure 5.3: Potassically-altered samples used for mass transfer analysis with abundant
quartz-magnetite “stringer” veins and stockwork............................................................................ 121
Figure 5.4: Propylitically-altered samples used for mass transfer analysis........................... 122
Figure 5.5: Hand samples and corresponding infrared spectra showing sericitic alteration.
The deep, symmetrical absorption features at 1400, 2200 nm along with small absorption
features at ~2350 and 2450 nm are diagnostic of muscovite. A) unaltered biotite-quartz
diorite protolith; B) potassically-altered biotite-quartz diorite with moderate intensity of
sericite alteration; and C) potassically-altered biotite-quartz diorite with very strong
sericitic alteration overprint. ..................................................................................................................... 123
Figure 5.6: Argillically-altered samples with corresponding infrared spectra. .................... 124
Figure 5.7: Plots of TiO2 and Zr versus Al2O3 illustrating conservative element behaviour
and mass addition results in overall dilution of these elements................................................. 126
Figure 5.8: Example of the effect of the arbitrary scaling method on definition of the
ISOCON. Black line represents the unscaled ISOCON. The red and orange lines are the
resultant ISOCON from arbitrary scaling of conservative components. Notice the drastically
different slope that results from the arbitrary scaling. ................................................................... 127
Figure 5.9: Example of an ISOCON for a potassically-altered biotite-quartz diorite sample
(C11.JL.178) showing net gains (green arrows) and losses (red arrow) of major
components using the least squares approach (Baumgartner and Olsen 1995).
Conservative components used to define the ISOCON (red line) are TiO2, Al2O3 and Zr. The
error bars indicate the analytical uncertainty for this sample and the standard deviation of
the average of the least-altered biotite-quartz diorite samples. ................................................. 127
Figure 5.10: Schematic diagram of overprinting alteration relationships used in the mass
transfer analysis reflecting the spectrum of overprinting alteration. Note that each
alteration type has variable intensities. A) Potassic alteration of least-altered biotite-quartz

xvi
diorite; B) propylitic alteration of previously potassically-altered biotite-quartz diorite; C)
propylitic alteration of a least-altered biotite-quartz diorite; D) sericitic alteration of least-
altered biotite-quartz diorite; E) sericitic alteration of potassically-altered biotite-quartz
diorite, with variable sericitic alteration intensity; and F) argillic alteration of an inferred
least- altered biotite-quartz diorite......................................................................................................... 130
Figure 5.11: Bar chart of net changes of major components during potassic alteration.
Uncertainties have been calculated through propagation of errors (see Section 5.2.4). .. 131
Figure 5.12: Bar chart of net changes in HFSE based on comparison of potassically- altered
biotite-quartz diorite with least-altered biotite-quartz diorite................................................... 132
Figure 5.13: Net changes of LILE are based on comparison of potassically-altered biotite-
quartz diorite with least-altered biotite-quartz diorite. ................................................................. 133
Figure 5.14: Net changes of LREE are based on comparison of potassically-altered biotite-
quartz diorite with least-altered biotite-quartz diorite. ................................................................. 134
Figure 5.15: Net changes of HREE are based on comparison of potassically-altered biotite-
quartz diorite with least-altered biotite-quartz diorite. ................................................................. 135
Figure 5.16: Net change of REE-affinity elements Sc and Y based on comparison of
potassically-altered biotite-quartz diorite with least-altered biotite-quartz diorite. ........ 135
Figure 5.17: Net change of copper and gold based on comparison of potassically-altered
biotite-quartz diorite with least-altered biotite-quartz diorite................................................... 136
Figure 5.18: Net change of metals based on comparison of potassically-altered biotite-
quartz diorite with least-altered biotite-quartz diorite. ................................................................. 137
Figure 5.19: Example of an ISOCON for propylitically-altered sample (C11.JL.35) compared
to (A) the average least-altered biotite-quartz diorite, and (B) to the average potassically-
altered biotite-quartz diorite showing net gains and losses of major components using the
least squares approach (Baumgartner and Olsen 1995). Error bars are the relative standard
deviation of the analyses for the propylitically-altered sample and the standard deviation
of the average least-altered biotite-quartz diorite and average potassically-altered biotite-
quartz diorite. ................................................................................................................................................... 138
Figure 5.20: Bar chart of net changes of major components during propylitic alteration
compared with an average potassically-altered biotite-quartz diorite (C11.JL.35, C11.JL.36
and C11.JL.128) or the average least-altered biotite-quartz diorite (C11.JL.92). “VS Prop” is
the average of very strongly propylitically-altered samples compared with the average
potassically-altered biotite-quartz diorite. Error bars were calculated through propagation
of error (see section 4.2.4) .......................................................................................................................... 139
Figure 5.21 Bar chart of net changes of HFSE based on comparison of propylitically-altered
biotite-quartz diorite to potassically-altered protolith (samples C11.JL.35, C11.JL.36 and
C11.JL.128) or least-altered biotite-quartz diorite (sample C11.JL.92). “VS Prop” is the

xvii
average of very strongly propylitically-altered samples compared with the average
potassically-altered biotite-quartz diorite. .......................................................................................... 140
Figure 5.22: Bar chart of net changes of LILE based on comparison of propylitically-altered
biotite-quartz diorite with potassically-altered protolith (samples C11.JL.35, C11.JL.36 and
C11.JL.128) or least-altered biotite-quartz diorite (sample C11.JL.92). “VS Prop” is the
average of very strongly propylitically- altered samples compared with the average
potassically-altered biotite-quartz diorite. .......................................................................................... 141
Figure 5.23: Bar chart of net changes of LREE based on comparison of propylitically-altered
biotite-quartz diorite with potassically-altered protolith (samples C11.JL.35, C11.JL.36 and
C11.JL.128) or least-altered biotite-quartz diorite (sample C11.JL.92). “VS Prop” is the
average of very strongly propylitically-altered samples compared with the average
potassically-altered biotite-quartz diorite. .......................................................................................... 142
Figure 5.24: Bar chart of net changes of LREE based on comparison of propylitically altered
biotite-quartz diorite to potassic altered protolith (samples C11.JL.35, C11.JL.36 and
C11.JL.128) or least-altered biotite-quartz diorite (sample C11.JL.92). “VS Prop” is the
average very strong propylitically altered samples compared with the average potassically-
altered biotite-quartz diorite. .................................................................................................................... 142
Figure 5.25: Bar chart of net change of REE-affinity elements Sc and Y based on comparison
of propylitically-altered biotite-quartz diorite with potassically- altered protolith (samples
C11.JL.35, C11.JL.36 and C11.JL.128) or least-altered biotite-quartz diorite (sample
C11.JL.92). “VS Prop” is the average of very strongly propylitically-altered samples
compared with the average potassically-altered biotite-quartz diorite. ................................. 143
Figure 5.26: Bar chart of largest net changes of metals based on comparison of
propylitically-altered biotite-quartz diorite with potassically-altered protolith (samples
C11.JL.35, C11.JL.36 and C11.JL.128) or least-altered biotite-quartz diorite (sample
C11.JL.92). “VS Prop” is the average of very strongly propylitically-altered samples
compared with the average potassically-altered biotite-quartz diorite. All changes are in
ppm except for Au, which is in ppb. ........................................................................................................ 144
Figure 5.27: Bar chart of net changes of metals based on comparison of propylitically-
altered biotite-quartz diorite with potassically-altered protolith (samples C11.JL.35,
C11.JL.36 and C11.JL.128) or least-altered biotite-quartz diorite (sample C11.JL.92). “VS
Prop” is the average of very strongly propylitically-altered samples compared with the
average potassically-altered biotite-quartz diorite. ......................................................................... 144
Figure 5.28: An ISOCON of the average sericitised biotite-quartz diorite (samples C11.JL.24,
C11.JL.25, C11.JL.138 and C11.JL.190) compared with the average least-altered biotite-
quartz diorite using the least squares approach (Baumgartner and Olsen 1995). Error bars

xviii
are the standard deviation of average sericitised biotite-quartz diorite and the average
least-altered biotite-quartz diorite.......................................................................................................... 145
Figure 5.29: An ISOCON of the average moderately-sericitised biotite-quartz diorite
(samples C11.JL.193, C11.JL.199 and C11.JL.200) compared with the average potassically-
altered biotite-quartz diorite using the least squares approach (Baumgartner and Olsen
1995). Error bars are the standard deviation of average very strongly-sericitised biotite-
quartz diorite and the average potassically-altered biotite-quartz diorite. ........................... 146
Figure 5.30: An ISOCON of the average very strongly-sericitised biotite-quartz diorite
(samples C11.JL.201 and C11.JL.205) compared with the average potassically-altered
biotite-quartz diorite using the least squares approach (Baumgartner and Olsen 1995).
Error bars are the standard deviation of average moderately-sericitised biotite-quartz
diorite and the average potassically-altered biotite-quartz diorite. ......................................... 147
Figure 5.31: Bar chart of net changes of major components during sericitic alteration
compared with the average least-altered biotite-quartz diorite (“Sericite_U”), moderate
sericitisation of an average potassically-altered biotite-quartz diorite (“Sericite_M”) or very
strong sericitisation of an average potassically-altered biotite-quartz diorite
(“Sericite_VS”). Error bars were calculated through propagation of error (section 4.2.4).
................................................................................................................................................................................ 148
Figure 5.32: Bar chart of net changes of HFSE during sericitic alteration compared with the
average least-altered biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an
average potassically-altered biotite-quartz diorite (“Sericite_M”) or very strong
sericitisation of an average potassically-altered biotite-quartz diorite (“Sericite_VS”). Error
bars were calculated through propagation of error (section 4.2.4). ......................................... 149
Figure 5.33: Bar chart of net changes of LILE during sericitic alteration compared with the
average least-altered biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an
average potassically-altered biotite-quartz diorite (“Sericite_M”) or very strong
sericitisation of an average potassically-altered biotite-quartz diorite (“Sericite_VS”). Error
bars were calculated through propagation of error (section 4.2.4). ......................................... 150
Figure 5.34: Bar charts of net changes of REE during sericitic alteration compared with the
average least-altered biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an
average potassically-altered biotite-quartz diorite (“Sericite_M”) or very strong
sericitisation of an average potassically-altered biotite-quartz diorite (“Sericite_VS”). Error
bars were calculated through propagation of error (section 4.2.4). ......................................... 151
Figure 5.35: Bar chart of net change of REE-affinity elements Sc and Y during sericitic
alteration compared with the average least-altered biotite-quartz diorite (“Sericite_U”),
moderate sericitisation of an average potassically-altered biotite-quartz diorite
(“Sericite_M”) or very strong sericitisation of an average potassically-altered biotite-quartz

xix
diorite (“Sericite_VS”). Error bars were calculated through propagation of error (section
4.2.4). ................................................................................................................................................................... 151
Figure 5.36: Bar charts of net changes of metals during sericitic alteration compared with
the average least-altered biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an
average potassically-altered biotite-quartz diorite (“Sericite_M”) or very strong
sericitisation of an average potassically-altered biotite-quartz diorite (“Sericite_VS”). Error
bars were calculated through propagation of error (section 4.2.4). All changes are in ppm
except for Au, where net changes are in ppb. ..................................................................................... 153
Figure 5.37: An ISOCON of the average dickite-dominant argillically-altered biotite-quartz
diorite (samples C11.JL.95, C11.JL.108 and C11.JL.110) compared with the average least-
altered biotite-quartz diorite using the least squares approach (Baumgartner and Olsen
1995). Error bars are the standard deviation from these averages. ......................................... 154
Figure 5.38: Bar chart of net changes of major components during argillic alteration
compared with an average least-altered biotite-quartz diorite. “Dickite ave” is the average
of dickite-dominant argillically-altered samples, “K-S ave” is the average of kaolinite-
smectite-dominant argillically-altered samples and “I-S ave” is the average of illite-
smectite-dominant argillically-altered samples compared with the average least-altered
biotite-quartz diorite. Error bars were calculated through propagation of error (section
4.2.4). ................................................................................................................................................................... 155
Figure 5.39: Bar chart of net changes of HFSE during argillic alteration compared with the
average least-altered biotite-quartz diorite compared with an average least-altered biotite-
quartz diorite. “Dickite ave” is the average of dickite-dominant argillically altered samples,
“K-S ave” is the average of kaolinite-smectite dominant argillically altered samples and “I-
S ave” is the average of illite-smectite dominant argillically altered samples compared with
the average least-altered biotite-quartz diorite. Error bars are calculated through
propagation of error (section 4.2.4) ....................................................................................................... 155
Figure 5.40: Bar chart of net changes of LILE in argillically altered samples compared with
an average least-altered biotite-quartz diorite. “Dickite ave” is the average of dickite-
dominant argillically altered samples, “K-S ave” is the average of kaolinite-smectite
dominant argillically altered samples and “I-S ave” is the average of illite-smectite
dominant argillically altered samples compared with the average least-altered biotite-
quartz diorite. Error bars were calculated through propagation of error (section 4.2.4).
................................................................................................................................................................................ 156
Figure 5.41: Bar charts of net changes of REE during argillic alteration compared with an
average least-altered biotite-quartz diorite. “Dickite ave” is the average of dickite-dominant
argillically altered samples, “K-S ave” is the average of kaolinite-smectite dominant
argillically altered samples and “I-S ave” is the average of illite-smectite dominant

xx
argillically altered samples compared with the average least-altered biotite-quartz diorite.
Error bars were calculated through propagation of error (section 4.2.4.) ............................. 157
Figure 5.42: Bar chart of net changes of elements in argillic alteration compared with the
average least-altered biotite-quartz diorite. A.) Major contributions to metals are seen in
predominantly dickite-dominant alteration assemblages, with less significant gains in
kaolinite-illite and smectite-illite assemblages. B.) Notice the significant gains in metals in
dickite-dominant argillic alteration. ....................................................................................................... 159
Figure 5.43: TiO2 versus Al2O3 for assay data (n = 11755) showing the least-altered average
biotite-quartz diorite (grey) and potassically-altered biotite-quartz diorite (pink) with all
assay data (A) and with assay data for samples attributed just to a potassically-altered
biotite-quartz diorite protolith (B).......................................................................................................... 160
Figure 5.44: Histogram of bulk mass change of altered samples (n=11240). ....................... 161
Figure 5.45: Plan view at 3810 m elevation showing contours of bulk mass change. Black
represents bulk mass change of +10%, orange represents bulk mass change of +60% and
red represents a bulk mass change of +100%. ................................................................................... 162
Figure 5.46: 763300E cross section across the Cerro Corona deposit contoured by bulk
mass change. The black, near horizontal line represents ground surface, and thick dark blue
line is proposed level of the final open pit............................................................................................ 163
Figure 5.47: Histogram of net change in K2O relative to precursor potassic alteration
(n=8025)............................................................................................................................................................. 164
Figure 5.48: Plan view at 3810 m elevation with contours of net change in K2O. ............... 165
Figure 5.49: 763300E cross-section across the Cerro Corona deposit contoured for net
change in K2O. The black, near horizontal line represents topography, and thick dark blue
line is proposed level of the final open pit............................................................................................ 166
Figure 5.50: Histogram of net change in Na2O of drill core samples relative to precursor
potassic alteration (n=9867). .................................................................................................................... 167
Figure 5.51: Plan view at 3810 m elevation with contours of net change in Na2O.............. 168
Figure 5.52: 763300E cross-section across the Cerro Corona deposit contoured by Na2O net
change. The black, near horizontal line represents topography, and thick dark blue line is
proposed level of the final open pit. ........................................................................................................ 169
Figure 5.53: Histogram of net change of CaO in drill core samples relative to precursor
potassic alteration (n=9491). .................................................................................................................... 170
Figure 5.54: Plan view at 3810 m elevation with contours of CaO net change. Note the areas
of CaO net loss are concentric around the post-mineral intrusions. ......................................... 171
Figure 5.55: 763300E cross section across the Cerro Corona deposit contoured by CaO net
change. The black, near horizontal line represents topography, and thick dark blue line is
proposed level of the final open pit. ........................................................................................................ 172

xxi
Figure 5.56: Histogram of net changes in Fe2O3 in drill core samples relative to precursor
potassic alteration (n=9319). .................................................................................................................... 173
Figure 5.57: Plan view at 3810 m elevation with contours of net change in Fe2O3. Note all
the Fe addition is concentrated in the western half of the intrusive complex. ..................... 174
Figure 5.58: 763300E cross-section across the Cerro Corona deposit contoured by Fe2O3
net change. The black, near horizontal line represents topography, and thick dark blue line
is proposed level of the final open pit. ................................................................................................... 175
Figure 5.59: Histogram of net change of P2O5 in drill core samples relative to precursor
potassic alteration. The darker fill with red outline represents the reduced dataset, after
removal of results in which the calculated net change is larger than the error (n=940). The
bars with a light blue fill represent the entire dataset for comparison (n=11755). ........... 176
Figure 5.60: Plan view at 3810 m elevation with contours of net change of P2O5. .............. 177
Figure 5.61: 763300E cross section across the Cerro Corona deposit contoured by P2O5 net
change. The black, near horizontal line represents topography, and thick dark blue line is
proposed level of the final open pit. ........................................................................................................ 178
Figure 5.62: Histogram of calculated net change in S relative to an average potassic altered
protolith (n=3651). ........................................................................................................................................ 179
Figure 5.63: Plan view at 3810 m elevation with contours of net change in S. Note similar
pattern to net Fe2O3 addition. .................................................................................................................... 180
Figure 5.64: 763300E cross section across the Cerro Corona deposit contoured by net
change in S. The black, near horizontal line represents topography, and thick dark blue line
is proposed level of the final open pit. ................................................................................................... 181
Figure 6.1: Example of the full infrared spectrum (red line) and absorption features
extracted from the hull-removed spectrum (dashed black line). Note that the relative
depths of the absorption features have changed in the hull-removed data. .......................... 185
Figure 6.2: Profile view (100) of kaolinite with single Al-octahedral sheet bonded to a Si-
tetrahedral sheet, thus a 1:1 phyllosilicate. Two cavities are occupied by Al ions. Two
hydroxyl groups are seen: 1) the outer hydroxyl groups or “inner surface” hydroxyls; and
2) the inner hydroxyl groups (termed by Frost and Johannson 1998). Si=khaki, Al=blue,
O=red and H=white. Adapted from Geysermans and Noguera (2009). ................................... 186
Figure 6.3 Examples of infrared spectra of kaolin group minerals, arbitrarily offset for
clarity. The dickite spectrum is displayed in red, kaolinite in orange and halloysite in green.
Note the doublets are the distinguishing features of this group of minerals, and are most
intense for dickite. The 1900 nm feature increases in depth from dickite->kaolinite-
>halloysite.......................................................................................................................................................... 187
Figure 6.4: (A) Location of the first overtone stretch (2vOH) bands related to inner surface
and inner hydroxyl vibrations in the kaolin group minerals. Note the position of the inner

xxii
hydroxyl absorption band is similar in all kaolin groups, whereas the position of absorption
bands related to inner surface hydroxyl is slightly shorter for dickite (red), than for
kaolinite (blue). (B)Absorption bands related to the combination mode of OH- stretching
(vOH) of inner surface and inner hydroxyl bands with bending of Al2OH groups................. 187
Figure 6.5: Profile view (100) of illite with single Al-octahedral sheet sandwiched between
two Si-tetrahedral sheets, thus a 2:1 phyllosilicate. The charge imbalance across the layer
is partially balanced by interlayer K+. Modified from Grim (1962). .......................................... 189
Figure 6.6: Example of an illite reflectance spectrum with distinguishing feature ranges
highlighted in red. The dashed red lines are the exact position and depth of these features
in the continuum removed spectrum. The absorption features located at 2340 and 2440 nm
are used to distinguish from smectite, which do not have these features. ............................. 191
Figure 6.7: Profile view (100) of smectite with single Al-octahedral sheet sandwiched
between two Si-tetrahedral sheets, thus a 2:1 phyllosilicate. The charge imbalance across
the layer is partially balanced by interlayer Ca2+, Na+ and H2O.Modified from Grim (1962).
................................................................................................................................................................................ 192
Figure 6.8: Combination bands of stretching (vOH) and bending (δ XAlOH) modes associated
with the smectite group minerals. The position of absorption bands is indicated by dashed
lines and labelled with the band modes. The band assignments are adopted from Sposito et
al. (1983), Post and Noble (1993), Cariati et al. (1983), Bishop et al. (2002) and Madejova
and Komadel (2001)...................................................................................................................................... 193
Figure 6.9: Spectral assignment of overtone and combination bands encountered in
smectite. Sample spectra are show for Ca-montmorillonite (blue), Na-montmorillonite
(pink), nontronite (light blue) and hectorite (green) (Clark et al. 1990). Dashed lines
indicate stretching (v) and bending (δ) combination band positions and the subscript
indicates the bonds of hydroxyl (OH) or water (w). For water molecule absorptions, v’
indicates the stretching of water bonded to the inner layer and v denotes the stretching of
adsorbed water. Assignment of positions of combination bands are from Cariati et al.
(1983), Bishop et al. (2002). The spotted grey and blue lines are unassigned features within
the infrared spectrum. .................................................................................................................................. 194
Figure 6.10: Spectrum of an illite-smectite mixture in which smectite is the dominant clay.
Dashed red lines indicate the depth and position of the absorption features. Refer to text
for further explanation of the characterisation of illite-smectite mixtures............................ 196
Figure 6.11: Cross section across 3810m level of the Cerro Corona intrusion with the
contours of the percentage of total clay content. Dots represent sample locations. .......... 197
Figure 6.12: Cross section along 76300 across the deposit contoured for total clay mineral
content based on SpecCam spectral logging. ....................................................................................... 198

xxiii
Figure 6.13: Pie chart showing the percentages of dominant mineral groups identified in
samples from Cerro Corona using infrared spectroscopy. ............................................................ 199
Figure 6.14: Pie charts of the abundance of the kaolinite group minerals (A) and their
mixtures (B). ..................................................................................................................................................... 200
Figure 6.15: Map of distribution of kaolinite, based on frequency of kaolinite identification
................................................................................................................................................................................ 201
Figure 6.16:The kaolinite content, based on semi-quantitative methods by SpecCam. .... 202
Figure 6.17: Histogram of the crystallinity index for dickite (a) and for kaolin group
minerals (b). There is a slight asymmetry toward a higher ratio in dickite-dominant spectra
and a more normal distribution for the rest of the kaolin group. “All” means all samples
with kaolinite and dickite. ........................................................................................................................... 203
Figure 6.18:Plan section across 3810 contoured with the kaolinite crystallinity. .............. 204
Figure 6.19: Cross section across 763300 with contours of kaolinite crystallinity index.
................................................................................................................................................................................ 205
Figure 6.20: Histogram showing the 2180/2210 absorption band ratio. A) The kaolin group
minerals have a bimodal distribution (grey), which can be separated into dickite (orange)
and kaolinite (maroon). B) The subgroup of the kaolin into kaolinite (maroon) and
halloysite (purple) appears artificial and does not distinguish two mineralogical groups.
................................................................................................................................................................................ 206
Figure 6.21: The location of the 1900 nm absorption feature. There are possibly two
overlapping peaks in this region, the major adsorption feature at ~1912 nm and a
secondary feature at ~1920 nm. .............................................................................................................. 206
Figure 6.22: Plot of the absorption depth ratios between dickite doublets. The dickite
crystallinity index is the ratio of the depth of 1380 nm feature versus the depth of the 1415
nm feature, whereas D2180/D2210 is the ratio of the depth of the 2180 nm feature versus
the depth of the 2210 nm feature. The most crystalline dickite samples have the highest
values in both doublets. ............................................................................................................................... 207
Figure 6.23: Histogram of the distribution of dickite crystallinity index (~1390/1415
absorption depths) and ~2170/2210 nm absorption depths. .................................................... 207
Figure 6.24: Relative proportions of illite-bearing mineral assemblages. The dominant
mixture observed is smectite-illite. ......................................................................................................... 208
Figure 6.25: Plan section across 3810 contoured for the illite content. .................................. 209
Figure 6.26: Cross section across 763300 with contours of calculated illite content. ....... 210
Figure 6.27: Histogram of illite end-members, based on the 2200 nm absorption feature
position. Illite with absorption feature <2204 nm is plotted on the secondary axis for clarity.
................................................................................................................................................................................ 211

xxiv
Figure 6.28: Histogram of the ratio of the depths of the 2200 and 2340 nm features as a
proxy for crystallinity. The smaller the value, the higher the “crystallinity”. ........................ 212
Figure 6.29: Plot of illite “crystallinity” index versus 2200 nm wavelength position for
kaolinite-illite-smectite, illite-smectite and illite clays. A) The red arrow indicates
increasing “crystallinity” with decreasing absorption position within the illite mixtures. B)
The subcategories of pure illite show that illite with < 2204 nm wavelength have a
consistently higher crystallinity (i.e. lower “crystallinity” index) than those with absorption
feature >2204 nm, which have a variable “crystallinity”. .............................................................. 213
Figure 6.30: Plan section of illite crystallinity index. ....................................................................... 214
Figure 6.31: Cross section of illite crystallinity index show that more crystalline illite is
centred in the north and central parts of the host intrusion ........................................................ 215
Figure 6.32: (A) Identified smectite mineral proportions and (B) the most commonly
identified pure smectite. .............................................................................................................................. 216
Figure 6.33:Plan section of overall smectite content across Cerro Corona. ........................... 217
Figure 6.34:Cross section across 763300 with contours of overall smectite content. ...... 218
Figure 7.1: Schematic diagram illustrating how clay minerals were separated from whole
rock samples. .................................................................................................................................................... 224
Figure 7.2: Characteristic X-ray diffraction pattern of dominant clay minerals, showing
location of diagnostic peak positions. .................................................................................................... 227
Figure 7.3: Back scattered electron images of clay mineral separates taken under the FEI
Quanta 650 scanning electron microscope. These images show sample homogeneity in: A)
Kaolinite-dominant sample (88 % kaolinite) JL2.216A. Field of view: 35 μm; B) Kaolinite-
dominant sample (97 % kaolinite) JL2.230F. Field of view: 35 μm, and; C) Illite-dominant
sample (81 % illite) JL2.98F. Field of view: 60 μm ........................................................................... 227
Figure. 7.4: Thermal analysis of predominant clay mineral separates of kaolinite (blue),
illite (orange) and smectite (green) with the specific temperatures at which dehydration
(lower temperatures) and dehydroxylation indicated. .................................................................. 228
Figure 7.5: Clay mineral proportions of samples selected for OH-removed oxygen isotope
analyses............................................................................................................................................................... 229
Figure 7.6: Measured versus true δ18O of all analysed standards (n=115) throughout
isotope study. The coefficient of determination is based on all standards............................. 232
Figure 7.7: Histogram of bulk mineral δ¹⁸O values (i.e. structural and hydroxyl oxygen).
................................................................................................................................................................................ 234
Figure 7.8 Depth versus bulk oxygen isotope values for A) kaolinite; B) illite, and; C)
smectite minerals. Note these are samples where the minerals are >50 wt. %. .................. 235
Figure 7.9: Histogram of the δ18O values of dehydroxylated clay mineral separates. ....... 236

xxv
Figure 7.10: Calculated formational temperatures of kaolinite using non-hydroxyl group -
hydroxyl fractionation method (1=Méheut et al. 2007) and hydroxyl-bulk oxygen isotope
method (2=Zheng. 1993) versus depth. ................................................................................................. 239
Figure 7.11 Histogram of hydrogen isotope results of clay minerals from Cerro Corona.
................................................................................................................................................................................ 241
Figure 7.12 Phase diagram for the system K2O-Al2O3-SiO2-H2O-KCl-HCl as a function of
temperature and K+/H+ activity ratio at a fluid pressure of 1 kbar (Seedorf et al. 2005). This
shows the stability ranges of hydrothermal minerals that typically form in the porphyry
environment from the alteration of aluminosilicate rock. ............................................................ 243
Figure 7.13: Fluid inclusion homogenisation temperatures and salinities for typical
alteration assemblages from 13 porphyry copper deposits (Bodnar et al. 2014). Propylitic
and argillic alteration are grouped together as their composition and homogenisation
temperatures are similar. ............................................................................................................................ 244
Figure 7.14:Vapour-dominant fluid inclusions associated with argillic alteration from Cerro
Corona. The large vapour phase makes identification of homogenisation temperatures
challenging......................................................................................................................................................... 245
Figure 7.15: Fractionation between muscovite and water for hydrogen isotopes, solid green
line, (Suzuoki and Epstein 1976) extended to lower temperatures (dashed green) and
oxygen isotope fractionation between illite and water (solid golden line, Sheppard and Gilg
1996) extended to higher temperatures (dashed golden line). The calculated formation
temperatures are derived from internal oxygen isotope values and are used to calculate the
isotopic differences between the mineral phases and the fluids from which they
precipitated. ...................................................................................................................................................... 247
Figure 7.16: Calculated formational fluids for kaolinite, illite and smectite at specific
temperature for samples in which these clay minerals comprise of at least 50 wt %. The
analytical error is displayed in the top left. ......................................................................................... 249
Figure 7.17: Evolving compositions of fluid from local meteoric, hydrous dacite and a
hydrous dacite that has degassed by ~40 % (low Al-amphibole) sources by interaction with
host rock at specified temperatures and decreasing water: rock ratios. See explanation in
text. ....................................................................................................................................................................... 254
Figure 7.18: Evolving compositions of fluid from local meteoric, hydrous dacite and a
hydrous dacite that has degassed by ~40 % (low Al-amphibole) source by interaction with
host rock at specified temperatures and decreasing water: rock ratios. See explanation in
text. ....................................................................................................................................................................... 257
Figure 7.19: Model of magma mixing between felsic magmatic and meteoric water. Fluid in
equilibrium with smectite is dominated by >70 wt% meteoric waters, whereas kaolinite is

xxvi
dominated by 40-65 wt. % and illite is predominated by magmatic waters. Analytical error
is shown in top left corner. ......................................................................................................................... 259
Figure 8.1: Molar K/Al versus (Na+K)/Al plot showing samples used in mass transfer
analysis. Black squares are ideal mineral compositions: “K-spar” is endmember orthoclase,
“Plag” is end-member albite, “K-mica” is muscovite mica. “Ave K-spar” is the average
analysed composition of K-spar that has replaced [albite – or plagioclase]. “BD ave” is the
bulk rock composition of the least-altered biotite-quartz diorite. For explanation of
alteration types, see Chapter 5. ................................................................................................................. 268
Figure 8.2: Breakdown of amphibole phenocryst. Numbered red spots are locations of
electron microprobe analyses. .................................................................................................................. 270
Figure 8.3: Chondrite-normalized (Sun and McDonough 1989) REE diagram for the average
least-altered biotite-quartz diorite porphyry and the modelled dilution of this REE pattern
by bulk mass gain of 20% increments compared with potassically-altered biotite-quartz
diorite samples. ............................................................................................................................................... 271
Figure 8.4: REE patterns of sericitised biotite-quartz diorites normalised to volatile-free C1
chondrite (Sun and McDonough, 1989). “Sericite_U” is the average of sericitised, previously
unaltered biotite-quartz diorite (“BD least alt”), “Sericite_M” and “Sericite_VS” are the
averages of moderately and very strongly sericitised, previously potassically-altered
biotite-quartz diorites (“Potassic average”) respectively. Modelled bulk mass gains (grey
lines) of 20% and 40% of the biotite-quartz diorite and 60%, 150% and 275% are modelled
for bulk mass gain of the average potassically-altered biotite-quartz diorites. ................... 275
Figure 8.5: Electron microprobe analyses of pyrite from Cerro Corona (sample C11.JL.122)
confirming: (A) cobaltiferous pyrite containing up to 0.57±0.06 wt.%; (B) pyrite with Zn up
to 0.26±0.04 wt.% and sphalerite infill along fractures; (C) arsenian chalcopyrite infilling
fractures with As up to 0.31±0.02 wt.%. Red spots are location of microprobe analyses.
................................................................................................................................................................................ 276
Figure 8.6: REE chondrite-normalised (Sun and McDonough 1989) patterns of average
argillically-altered biotite-quartz diorite compared with least-altered, potassically-altered
and propylitically-altered biotite-quartz diorite (“BD least alt”, Potassic average” and “VS
Prop”, respectively). Grey lines indicate 30% bulk mass gain of least-altered biotite diorite
and 20% and 40% bulk mass gain of potassically-altered biotite-quartz diorite. .............. 278
Figure 8.7: Cross section across 763300 showing inferred intrusion causing drastic bulk
gains in northern area of the deposit and their coincidence with logged pyrite, magnetite
and chalcopyrite, and an inverted cup-shaped Cu grade shell. ................................................... 280
Figure 8.8: N-S cross section showing net loss of Na2O and logged areas where chlorite
>10%. ................................................................................................................................................................... 282

xxvii
Figure 9.1: Calculated net change during potassic alteration in other deposits. Bulk mass
change and SiO2 are plotted on the left axis, as their changes are larger than other
components read off the right hand axis. Abbreviations are in the order of deposit,
alteration and lithology where: BdlA., Arg. = Bajo de la Alumbrera porphyry Cu-Au,
Argentina (Ulrich and Heinrich 2002); B.H., Ind. = Batu Hijau porphyry Cu-Au, Indonesia
(Idrus et al. 2009); Chu, Chi. = Chuquicamata porphyry Cu-Au, Chile (Arnott 2003); Du., T.,
= Duobuza porphyry Cu-Au, Tibet (Li et al. 2011); M-A, Ir.= Maher-Abad porphyry Cu-Au,
Iran (Siahcheshm et al. 2014); Pot. Alt= Potassic alteration; Early P3= Early P3 porphyry;
Late P3=Late P3 porphyry; NW Por=NW porphyry; E Por=East porphyry; Unk.=lithology
uncertain; Late Grd.= Late Granodiorite; QM= Quartz monzonite; And. Clast= Andesitic
volcaniclastics; I.T.= Intermediate tonalite; EQD= Equigranular quartz diorite; Y.T=Young
tonalite. ............................................................................................................................................................... 292
Figure 9.2: Cross section with clay mineral distribution pattern with interpretations of their
distribution based on temperature and acidity. ................................................................................ 295

List of Tables
Table 3.1: Compilation of the nomenclature and ages of major deformation events in the
Cenozoic of Northern Peru. * nomenclature adopted in this thesis............................................. 39
Table 4.1: Radiometric ages of intrusions in the Hualgayoc district. Cerro is abbreviated to
“Cᵒ”. ......................................................................................................................................................................... 64
Table 4.2: Alteration assemblages at Cerro Corona (James 1998)............................................... 69
Table 4.3: Parameters for grade shell construction. ........................................................................ 106
Table 7.1: Summary of treatments and their effects on the basal 001 reflection of clay
minerals. *peak intensities change (001 increase, higher-order peaks 002, 003, 004
decrease) **better comparison of 001 reflections of Kaolinite ~14 Å, Chlorite ~14 Å..... 226
Table 7.2 Standards used for oxygen isotope analyses. See text for descriptions of
standards............................................................................................................................................................ 231
Table 7.3: Standards used for hydrogen isotope analyses. ........................................................... 233
Table 7.4: Calculation of oxygen isotope composition within kaolinite structure and
calculated temperature (1=Méheut et al. 2007 method; 2=Zheng 1993 method). ............... 238
Table 7.5: Calculation of oxygen isotope composition within illite structure and calculated
temperature using logarithmic isotope fractionation method (1=Zheng 1993) and linear
isotope fractionation (2= Bechtel and Hoernes 1990). ................................................................... 240
Table 7.6: The calculated isotope composition of fluids in equilibrium with kaolinite at
temperatures inferred from kaolinite stability field (Seedorf et al. 2005) ............................. 246

xxviii
Chapter One: Literature review

Table 7.7: Calculated water: rock exchange ratios and temperatures required to evolve
meteoric water to compositions comparable to those of kaolinite-forming fluids. ............ 252
Table 7.8: Calculated water: rock exchange ratios and temperatures required to evolve
meteoric water to compositions comparable to those of illite-forming fluids...................... 255
Table 7.9: Calculated water: rock exchange ratios and temperatures required to evolve an
exsolved dacitic fluid to compositions comparable to those of illite-forming fluids .......... 256

xxix
Chapter One: Literature review

1 Chapter One: Literature review

The geological processes involved in forming an economic porphyry deposit are not purely
tectonic, magmatic or hydrothermal, but rather a combination of these processes that
operate at various scales.

This thesis focuses on the final stages of porphyry ore formation, where hydrothermal
processes take centre stage. This chapter aims to review the primary controls of porphyry
formation and places hydrothermal alteration in the context of the wider porphyry deposit
model.

1.1 Porphyry ore-forming processes

1.1.1 Tectonic setting of porphyry deposits


On a global scale, porphyry ore systems are located at convergent margins where
subduction of an oceanic plate is occurring (such as the circum-Pacific), or has known to
have occurred (such as Oyu Tolgoi, Mongolia; Perrelló et al., 2001) (Fig. 1.1).

Figure 1.1: Global distribution of porphyry deposits (USGS, 2017) and subduction zones.

Best (2003) provides a clear understanding of subduction-related magmas, collectively


known as arc magmas. Island arcs are commonly comprised of high K to alkalic suite
magmas whereas continental arcs tend to be more calc-alkaline in composition. Calc-
alkaline rocks are found in mature arcs, with thicker crust which reflect the relative timing
of crystallisation of Fe-Ti oxides, plagioclase and ferromagnesian silicates within a suite of
rocks (Peacock 1931). The term calc-alkaline has evolved over time to provide a

1
Chapter One: Literature review

petrotectonic indicator of a sample; however, the term sensu stricto refers to the variation
of total Fe versus FeO/MgO in evolving subalkaline rock suites and not a single sample
(Arculus 2003).

Within subduction zones, melting of metasomatized mantle generates a basaltic melt which
buoyantly rises to the base of the lower crust where it mixes with sub-crustal melts,
assimilates crustal rocks and is stored and therefore becomes homogeneous (this is known
by the acronym “MASH” zone; Hildreth and Moorbath, 1988;). These homogenised melts
are envisaged to be emplaced into the lower crust via a series of sills whereby silicic,
hydrous magmas are proposed to form by the progressive injection of basaltic melts that
assimilate crustal rocks and supply volatiles (Annen et al. 2006).

Shear zones, and brittle faults at shallower depth, create areas of low pressure in the
overlying lithosphere, which allow these hydrous, silicic magmas to rise as dykes to upper
crustal levels (Fig. 1.2). Local extensional stress regimes aid in the ascent of these dykes to
shallow levels, where they group to form shallow batholiths (Richards, 2003). The pre-
existing crustal stress and strain patterns may be locally important in the final stages of
magma emplacement (Richards, 2003).

Figure 1.2: Tectono-magmatic model for the emplacement of porphyry deposits (Richards, 2003). Hot zones at
the base of the lower crust cause initial homogenisation of magmas (as described in the text). Within the shallow
crust, magma ascent is controlled by local stresses. The magma rises to a state of neutral buoyancy, where it
pools to form a subvolcanic pluton. Note the change in scale in the upper crust.

2
Chapter One: Literature review

During decompression of the magmatic body in the shallow crust, volatiles within it reach
saturation (“first boiling”). This is a critical stage in porphyry ore development as volatiles,
being less dense, are thought to migrate to the upper parts of magma bodies and collect
beneath a crystallising carapace (Cloos 2001).

The increase in volume that occurs during first boiling causes a pressure increase in the
carapace. This increase in pressure causes fracturing in the crystallised magma and host
rock, allowing for rapid escape of fluid into a developing fracture network, known as a
stockwork.

As the typically low to moderate salinity fluid rises out of the magma, a high temperature
hypersaline liquid (brine) may condense out. The remaining fluid may continue its ascent
and can cool within a single-phase stability field (Heinrich 2005).

Crystallisation of the magma can also serve to increase the volatile content of the melt, as
minerals do not generally contain water. This process is known as “second boiling”. The
fluids exsolving from magmas during these latter stages are the agents of metal transport
and deposition and host rock alteration in porphyry ore systems.

The emplacement of a porphyry stock into the upper crust is known to drive hydrothermal
fluid convection which leads to metasomatism of the host rock (i.e. alteration; Giggenbach
1997). The chemistry and temperature of the fluids will dictate the extent and type of
alteration. Alteration associated with porphyry intrusions has been documented and
incorporated into models since Lowell and Guilbert (1970). This alteration is used in
exploration as a vectoring tool for locating porphyry deposits. However, there is still some
controversy in the timing and the source of some of these metasomatic fluids (especially in
phyllic alteration).

The chemistry of the fluid affects the efficiency of stripping out of the key ore-forming
components - chalcophile elements and sulphur - from the magma. As magmatic fluids
evolve into hydrothermal fluids there are four controls on their chemical behaviour, these
are acidity, redox potential, pressure and temperature (Giggenbach 1997).

CO2, HCl and S-derived species (H2SO3, H2SO4, HSO4- and H2S) are the most important
sources of acidity in hydrothermal fluids (Giggenbach 1997). S species, however, are weak
acids and, during porphyry ore formation, most magmatic S is removed from fluids by the
precipitation of sulphides or sulphates. In porphyry systems at temperatures above
~350°C, HCl is the dominant acid, whereas at temperatures below this, CO2 is the major
acidifier (Fig. 1.4).

3
Chapter One: Literature review

100 10
CO2 (initial)

80 Full equilibrium cooling 8


XCl (mol %)

CCO2 (wt %)
60 6
CO2 (aq)

40 4

20 2

0 0
800 600 400 200
Temperature (°C)

Figure 1.3: Fluid composition versus temperature diagram illustrating that under full equilibrium CO2 becomes
the major control of fluid acidity below ~350°C (Giggenbach 1997).

Many porphyry deposits have close spatial and temporal associations with epithermal
deposits. The fluids which form the epithermal deposits are thought to contain a mixture
of magmatic fluids and meteoric waters. The magmatic fluids are either separated into low
salinity vapour and dense hypersaline liquid at depth or during ascent, or may remain in
the single-phase state under certain P-T conditions. In the case of an ascending, low density,
vapour-like fluid, contraction into a liquid-like state can occur (Heinrich et al. 2004).

At the higher temperature end of hydrothermal activity in porphyry systems, dissociation


of HCl leads to the alteration of feldspars to sheet silicates (Giggenbach 1997) as described
by the equation below:

3𝐾𝐴𝑙𝑆𝑖 𝑂 + 2𝐻𝐶𝑙 → 𝐾𝐴𝑙 𝑆𝑖 𝑂 (𝑂𝐻) + 2𝐾𝐶𝑙 + 6𝑆𝑖𝑂

3 𝐴𝑙𝑘𝑎𝑙𝑖 𝑓𝑒𝑙𝑑𝑠𝑝𝑎𝑟𝑠 + 2ℎ𝑦𝑑𝑟𝑜𝑐ℎ𝑙𝑜𝑟𝑖𝑐 𝑎𝑐𝑖𝑑 → 𝑖𝑙𝑙𝑖𝑡𝑒 + 2 𝑝𝑜𝑡𝑎𝑠𝑠𝑖𝑢𝑚 𝑐ℎ𝑙𝑜𝑟𝑖𝑑𝑒 + 6 𝑞𝑢𝑎𝑟𝑡𝑧

Redox potential is controlled by temperature, with cooling fluids becoming more oxidised
(Giggenbach 1987). At magmatic temperatures, mineral buffers (such as FMQ) control the
oxidation state; however, this buffer decreases in importance with decreasing temperature.
At lower temperatures, the redox potential may be governed by redox reactions involving
H2S and SO2 (termed the “gas buffer”).

4
Chapter One: Literature review

1.1.2 Metal zoning in porphyry systems


The metal budget of a porphyry deposit is intrinsically linked to igneous processes
(Wilkinson 2013), which are beyond the scope of this thesis. The process involved in
transporting and precipitating metals into ore are principally controlled by fluid chemistry
and fluid: rock interactions. This is demonstrated by the concentric zoning surrounding
porphyry deposits, which viewed in vertical cross section may show gradational zonation
(Halley et al. 2015, Fig. 1.4). This zonation of metals can be used as a vector toward the
central ore zone and will be further investigated in this thesis.

Figure 1.4: Schematic vertical cross section of metal distribution in porphyry deposits (Halley et al. 2015).

5
Chapter One: Literature review

1.1.3 Alteration zoning in porphyry systems


The importance of distinguishing between mineral assemblages and mineral associations
is commonly neglected. A mineral association is a group of minerals which occur together
and may have been formed during a variety of overprinting events (Seedorff et al. 2005). A
mineral assemblage, however, is a group of minerals that formed contemporaneously and
are stable together. Thus, a mineral assemblage provides a convenient indicator of the local
thermodynamic equilibrium in that system, whereas a mineral association may reflect a
paragenetic overprint of two or more mineral assemblages.

Alteration mineral assemblages are a useful guide in mineral exploration as they provide
an indication of the key components in the hydrothermal environment: temperature;
pressure; chemical composition of host rocks; chemical composition of external fluids, and
structural controls of fluid flow. Genetically, the deepest alteration and most proximal to
an intrusion is sodic-calcic alteration (Fig. 1.5). This grades to potassic alteration which
typically fringes the porphyry intrusion and forms a dome and is often coincident with
hypogene ore mineralisation. Potassic alteration typically grades into sericitic alteration
which may be structurally controlled. Argillic and propylitic alteration assemblages occur
more distal to the porphyry intrusion and mineralisation (Fig. 1.5). The final distribution
of the alteration assemblages is typically more complicated by changes in hydrology
resulting from the emplacement of multiple intrusions, regional uplift and erosion and the
development of structures.

Figure 1.5: Idealised distribution of alteration assemblages associated with a porphyry deposit (Seedorff et al.
2003).

The definition of mineral assemblages - potassic, sericitic (or phyllic), propylitic, argillic
and advanced argillic - are loosely utilised to encompass the minerals that reflect the

6
Chapter One: Literature review

thermodynamic equilibrium reached in the local environment. The controls on the


development of these alteration assemblages can be considered more quantitatively using
phase diagrams (Fig. 1.6). Magmatic fluids have a high KCl/HCl ratio which causes K-
metasomatism of the rock (Fig. 1.6). This K-metasomatism can manifest itself in various
ways, depending on the host rock mineralogy. In mafic host rocks it is characterised by the
replacement of amphiboles and pyroxenes by biotite. In more felsic host rocks, plagioclase
phenocrysts are altered to orthoclase and secondary, shreddy, biotite may form around,
and partly replace, primary igneous biotite.

700

600
KSP 600 °C Bi
Cor 0
And Ksp -1
500 AND
Temperature (°C)

ADVANCED
400
ARGILLIC

Log (aMg+2/a2H+)
PYRO 300°C 6
Bi
300 Chl 5
Pyro Musc Ksp
4
ILL
200 KAOL
150°C Bi 8
100 Chl
ILL-MONT Ksp 9
Kaol Ill
0 2 4 6
0 2 Log (aK+/aH+)
1 3 4 5 6
Log (mKCl+K+/mHCl+H+)

Figure 1.6: Phase diagram for the system K2O-Al2O3-SiO2-H2O-KCl-HCl at atmospheric pressure. The labelled
areas broadly correspond to alteration assemblages. Mineral abbreviations are: Cord-cordierite; And-andalusite;
Ksp-potassium feldspar; Bi-biotite; Chl-chlorite; Pyro-pyrophyllite; Musc-muscovite; Kaol-kaolinite; Ill-illite;
Mont-montmorillonite. Note that chlorite, cordierite and biotite stability fields are based on a simple solid
solution model for their Mg end members (Seedorff et al. 2005).

As the magmatic fluids cool there is a change in the alteration assemblage from
predominantly K-silicates (i.e. K-feldspar and biotite) to hydrated K-silicates (e.g.
muscovite). This transition from potassic to phyllic (or “sericitic”) alteration reflects the
reducing ratio of KCl/HCl in the fluid (Figs. 1.6 and 1.7). This low-pressure degassing is
thought to link to roots of epithermal deposits, such as Lepanto (Hedenquist et al. 1998).

7
Chapter One: Literature review

-1
Volcanic waters Volcanic vapours
ANDALUSITE

Crater lakes
LHK=log(mHCl/mKCl)

-2
Direct
KAOLINITE
degassing
from a
K-MICA
“boiling”
-3
magma

K-FELDSPAR

-4
100 200 300 400 500 600
Temperature (°C)

Figure 1.7: The stability of aluminosilicates is primarily a function of mHCl/mKCl and temperature. The red arrow
shows the path that a vapour phase released from a degassing magma is likely to take (i.e. low P degassing)
(Giggenbach, 1997).

The phyllic alteration assemblage is characterised by fine-grained muscovitic mica (termed


“sericite”), quartz and pyrite. It is important to note that in this thesis “sericite” is used in
its colloquial sense as a general term referring to a fine-grained white mica, typically with
a muscovitic composition and “sericitic” is used to define an alteration assemblage
containing a fine-grained muscovite mica.

The sericitic alteration assemblage is identified by most silicate minerals (except quartz)
being replaced by muscovite and pyrite. K-feldspar is often completely sericitized, and
pyrite occurs both as veins and disseminated throughout the rock (averaging about 5-10
wt. %; Lowell and Guilbert 1970). Quartz associated with sericite and pyrite is commonly
present as sheeted veins or silicified zones (Fig. 1.5). This alteration assemblage can form
under a range of temperatures, depending on the fluid composition (Fig. 1.6). This has
resulted in much controversy as to the ultimate origin of this alteration assemblage, as it
appears that it may form in equilibrium with magmatic and/or meteoric waters.

Propylitic alteration is characterised by the formation of chlorite and epidote with variable
amounts of amphibole (actinolite), smectite and carbonate. Overall, this alteration zone is
weak. There is generally an addition of iron and magnesium, but the majority of the
metasomatism is correlated with calcium addition (Burnham 1962). Although weak,
propylitic alteration is usually the most extensive alteration facies and can extend several

8
Chapter One: Literature review

kilometres from a deposit (Fig. 1.5). This extensive alteration has been subdivided into
several sub-zones, based on the relative abundance of the major phases (Cooke et al. 2014).

Argillic alteration is a common alteration product in porphyry deposits. Spatially, this


alteration assemblage typically occurs above the deposit (Fig. 1.5), although it may be
found lateral to, or telescoped onto it. This assemblage is not always preserved as the clay
minerals that characterize it are easily eroded during uplift and exhumation. When
preserved, the variation within the assemblage provides useful indications of the
temperature and acidity of the fluids from which the minerals formed (Fig. 1.6). This
alteration assemblage is the primary focus of this thesis, as it is telescoped onto the Cerro
Corona deposit and directly impacts the economics of the deposit.

The mineralogy that defines the argillic alteration assemblage is varied. Broadly speaking,
it commonly includes an excess of quartz (which is textually vuggy and dark grey; Fig. 1.9),
illite, smectite and kaolinite group minerals with lesser amounts of alunite, pyrophyllite
and diaspore. This broad group is typically subdivided into mineral assemblages that
reflect temperature and KCl/HCl (Fig. 1.6). At the highest temperature and lowest KCl/HCl
is the “advanced argillic” assemblage, defined by vuggy silica, pyrophyllite, andalusite,
kaolinite, and dickite. This advanced argillic alteration is thought to be a result of the
cooling and condensation of magmatic vapour into groundwaters. These vapours are SO2-
and HCl-rich, and can form highly acidic (pH ~0.5-1.5) solutions (Hedenquist et al. 1994)
when absorbed into water as the acid generating species hydrolyse. In extreme cases, all
silicate minerals (except quartz) are stripped from the rock resulting in a “vuggy”
appearance (Fig. 1.9). to form what has been referred to as a lithocap.

9
Chapter One: Literature review

Figure 1.8: Fluid pathway for magmatic liquids and vapours involved in the formation of argillic alteration
(Hedenquist et al. 1998) projected onto a phase diagram showing the stability of minerals in the K2O-Na2O-Al2O3-
SiO2-HCl system (Sverjensky et al. 1991). Dashed grey lines are approximate boundaries between clay minerals.
Abbreviations for minerals are: And=andalusite, Pyro(ph)=pyrophyllite, Dick=dickite, Kaol=kaolinite,
Musc=muscovite, Ill=illite, Smec=smectite and Ksp=potassium feldspar.

Vuggy texture

Figure 1.9: An example of the “vuggy quartz” from the Cerro Corona deposit. Note the rock is dominated by dark
silica replacement.

10
Chapter One: Literature review

Argillic alteration was initially thought of as a more distal equivalent to the phyllic
alteration zone (Lowell and Guilbert 1970). However, more recent studies have shown that
the exact timing of this alteration is not very well understood. In theory, there is no reason
why an argillic lithocap cannot occur synchronously with deeper potassic alteration
(Hedenquist et al. 1998) if the magmatic fluids are undergoing phase separation and
releasing low density, acidic gases.

Subsequent magmatic fluids which flow through this lithocap may cause mineralisation,
usually forming a high sulfidation epithermal deposit (Sillitoe and Hedenquist, 2003).

1.2 Techniques used in this thesis


Mapping out the distribution of various clay minerals may aid in vectoring toward higher
temperature assemblages, which may be proximal to mineralisation. This critical aspect is
challenging due to the complexities in identification of the clay minerals, largely owing to
their small grain size (usually <2 μm). Their small size inhibits classical methods of mineral
identification and thus relies on more sophisticated tools for identification.

1.2.1 Infrared spectroscopy


Clay mineral identification utilizing chemical composition can be problematic due to the
high-water content in their mineral structure, which volatilises at relatively low
temperatures, leading to inaccurate mineral chemistries. The most common method for
clay mineral identification is X-ray diffraction, whereby the crystallography of the clay is
determined from diffraction angles between aligned atomic sheets (Moore and Reynolds
1989). An alternative approach of mineral identification exploits the vibrational modes
between molecular units in a mineral. These are routinely investigated using the infrared
part of the electromagnetic spectrum.

Reflectance spectroscopy is a well-established method in identification of minerals,


particularly in the remote sensing community (Elachi and Van Zyl 2006, van der Meer et al.
2012). The infrared (IR) part of the spectrum falls between that of X-rays (at higher
frequencies and thus lower wavelength) and microwaves and radio waves (at lower
frequencies and higher wavelength). The IR is subdivided into visible-, near-, mid- and far-
infrared, with the subdivisions depending on the community of users. Herein, the
nomenclature that is adopted is that of the remote sensing community whereby the
spectrum between 350 and 1000 nm is referred to as visible and near-infrared (VNIR) and
between 1000-2500 nm is referred to as short-wave infrared (SWIR). The SWIR region is

11
Chapter One: Literature review

the main region of importance in the identification of clay minerals, owing to the strong
absorption features that these minerals exhibit in this region (Fig. 1.10).

When a bond is excited, it vibrates in a restricted number of motions (or “modes”).


Overtones of this vibration occur at a fundamental frequency corresponding to the
constituent atoms, their spatial geometry and the strength of the bond (Hunt 1977). The
most relevant to this thesis are the overtones and combination of the vibrational modes of
water and hydroxyl groups as these are utilised to differentiate between phyllosilicates.
These overtones and combination bands manifest themselves in the resulting infrared
reflectance spectrum as absorption features, of varying depth and at specific wavelengths.

Absorption features from overtones and combination bands of the water molecule are
caused by symmetrical stretching (v1), H-O-H bending (v2) and asymmetrical OH- stretching
(v3). The presence of liquid water is determined by the presence of absorption features at
approximately 1400 and 1900 nm (Hunt 1977). The location of the absorption feature
associated with the fundamental stretching mode of the OH- group depends primarily on
the atom to which the OH- is combined and the location of the bond within the crystal
lattice, but is generally at ~2750 nm – outside the range of most SWIR spectrometers (Hunt
1977; Petit et al. 1999; Balan et al. 2010). However, the first OH- overtone, located at ~1400
nm and the combination of OH- stretch and a metal-OH- bend, located between 2200-2300
nm, are present in the SWIR range of most spectrometers (Fig. 1.10). In this thesis clay
minerals are identified using a handheld infrared spectrometer.

Figure 1.10: Location of absorption features commonly found that are related to clay minerals. These absorption
features are due to the combination of symmetrical stretching (v1), H-O-H bending (v2) and asymmetrical OH-
stretching (v3) of the water molecule (shown in blue) and the combination of hydroxyl stretching (vOH in yellow)
with metal (Al, Fe or Mg) hydroxyl bending (δXOH in peach).

12
Chapter One: Literature review

The effect of grain size directly impacts the reflectivity of the material; smaller grain sizes
have a higher reflectivity than larger grains. This is due to the smaller grain sizes having a
higher surface-to-volume ratio than larger grains, and thus a higher proportion of
reflectivity per volume. This hinders the use of reflectance spectra for quantification of
mineral modes, unless all materials are of a uniform grain size (Clark and Roush 1984).

There are two important factors to consider in the quantification of clay minerals utilising
infrared spectroscopy: the quantification from spectrum deconvolution and the
quantification from frequency of mineral identification (i.e. mineral mapping).
Deconvolution of mixtures from a spectrum is a post-analytical process, whereby a
spectrum is unmixed into separate known “active” phases of determinable proportions.
The Spectral Analyst® (TSA) software utilises linear mixing between pre-determined
(“library”) end-members to determine the mineral abundances (Berman et al. 1999). This
deconvolution method yields errors of up to 10% under ideal conditions. This error may
be significantly higher when spectra are acquired on dark (i.e. low reflection) material.

The deconvolution method also fails to address the fact that the spectrum is a
representation of the bonds present, and, as such, some bonds show greater intensities of
absorption than others (i.e. that the bonds are not necessarily linearly related). The
relationships between bond strengths, absorption intensities and mineral abundances
have yet to be empirically determined. For example, a series of physical mixtures of pure
dickite and pure kaolinite demonstrate dickite bond absorbance is greater than that of
kaolinite (Fig. 1.11).

Figure 1.11: An example of physical mixtures between dickite and kaolinite in 20 wt.% increments (this study).
The ~1380 nm absorption feature of dickite is more intense than that of the equivalent ~1395 nm absorption
feature of kaolinite. Thus, when using a linear mixing algorithm for quantification, the dickite proportion is
overestimated.

13
Chapter One: Literature review

Quantification by deconvolution is particularly imprecise when a sample is composed of a


range of crystal sizes, as is the case for most geological specimens. Smaller grain sizes lead
to more prominent absorptions so that fine-grained minerals are overestimated (Clark and
King 1987). In this regard, the use of “quantification by deconvolution” should be treated
as a qualitative estimate of the “active” mineral abundance and not a true quantitative
estimate of the proportions of minerals present.

To provide a more robust quantification of the abundance of an IR-active mineral (such as


clay), a second method may be adopted. Instead of analysing a representative sample and
deconvolving its spectrum, the frequency of infrared analyses per interval is increased,
such as by using a core scanner. This increase in analyses may produce a more realistic
quantitative result than single spectrum deconvolution. The best method for quantification
utilising infrared spectroscopy would be a combination of these methods: use of a core
scanner to generate a spectrum for each pixel scanned and a deconvolution of this spectrum
to identify the major phases present.

1.2.2 Stable isotopes


To assess the contribution of magmatic and meteoric components involved in the
formation of argillic alteration, oxygen and hydrogen stable isotope analysis of the minerals
can be carried out (e.g. Taylor 1997). The isotope fractionation between these minerals and
fluid is a function of the:

a) isotope composition of the fluid in equilibrium with the clay at the time of its
formation, and;
b) temperature of the environment in which the clay mineral formed.

Thus, the stable isotope composition provides a proxy for formation conditions, assuming
equilibrium with the fluid and that no post-formation fractionation has occurred.

Classical oxygen isotope analyses of kaolinite, illite and smectite often focus on the bulk
clay mineral (Savin and Epstein 1970). Early isotope fractionation experiments suggested
internal oxygen fractionation was negligible (O’Neil and Kharaka 1976), but subsequent
experimentation into oxygen isotopes of hydroxyl-bearing minerals suggested there may
be an exchange amongst various oxygen sites within these minerals, i.e. structural O in the
tetrahedral site and OH- in the interlayer site (Hamza and Epstein 1980).

Within clay mineral structures, there are two oxygen sites and a single hydrogen site. Any
adsorbed water can be viewed as a contaminant and accounts for a large degree of
variability as it is not structurally bound. In this thesis “hydroxyl” refers to the structural

14
Chapter One: Literature review

OH- and accounts for all hydrogen and the hydroxyl component of the oxygen. The
“structural” oxygen refers to the oxygen that is bonded with Si and Al (i.e. that is not
hydroxyl or adsorbed; see Chapter 6). The “bulk” isotopic composition refers to the sum of
the hydroxyl and structural isotopic compositions.

In this thesis, stable isotopes are used to evaluate the proportions of magmatic and
meteoric waters involved in the formation of argillic alteration.

1.3 References
Annen, C., Blundy, J. D., Sparks, R. S. J. (2006) The Genesis of Intermediate and Silicic
Magmas in Deep Crustal Hot Zones. Journal of Petrology, 47(3) 505–539.

Arculus, R. J. (2003). Use and abuse of the terms calcalkaline and calcalkalic. Journal of
Petrology, 44(5), 929-935.

Balan, E., Delattre, S., Guillaumet, M. and Salje, E. K. H. (2010) Low-temperature infrared
spectroscopic study of OH-stretching modes in kaolinite and dickite. American
Mineralogist. 95 (8-9), 1257-1266.

Berman, M., Bischof, L. and Huntington, J. (1999) Algorithms and software for the
automated identification of minerals using field spectra or hyperspectral
imagery. Thirteenth International Conference on Applied Geologic Remote Sensing. 222-232.

Best, M. G. (2003) Differentiation of magmas. In: Best, M. G. (ed.). Igneous and Metamorphic
Petrology.

Burnham, C. W. (1962) Facies and types of hydrothermal alteration. Economic Geology. 57


(5), 768.

Clark, R. N. and King, T. V. V. (1987) Causes of spurious features in spectral reflectance data.
JPL Proceedings of the 3rd Airborne Imaging Spectrometer Data Analysis Workshop. United
States.

Clark, R. N. and Roush, T. L. (1984) Reflectance spectroscopy: Quantitative analysis


techniques for remote sensing applications. Journal of Geophysical Research: Solid Earth. 89
(B7), 6329-6340.

Cooke, D.R., Hollings, P., Wilkinson, J.J., and Tosdal R.M. (2014) 13.14 - Geochemistry of
Porphyry Deposits, In: Holland, H. D. and Turekian, K.K. (ed.). Treatise on Geochemistry
(Second Edition). Oxford, Elsevier. 357-381.

15
Chapter One: Literature review

Cloos, M. (2001) Bubbling Magma Chambers, Cupolas, and Porphyry Copper Deposits.
International Geology Review. 43 (4), 285-311.

Elachi, C., and Van Zyl, J. J. (2006). Introduction to the physics and techniques of remote
sensing (28). John Wiley and Sons.

Giggenbach, W. F. (1987). Redox processes governing the chemistry of fumarolic gas


discharges from White Island, New Zealand. Applied Geochemistry, 2(2), 143-161.

Giggenbach, W. F. (1997). Relative importance of thermodynamic and kinetic processes in


governing the chemical and isotopic composition of carbon gases in high-heatflow
sedimentary basins. Geochimica et Cosmochimica Acta, 61(17), 3763-3785.

Halley, S., Dilles, J. H., Tosdal, R. M. (2015) Footprints: Hydrothermal Alteration and
Geochemical Dispersion Around Porphyry Copper Deposits. SEG Newsletter (100), 1-17.

Hamza, M. S. and Epstein, S. (1980) Oxygen isotopic fractionation between oxygen of


different sites in hydroxyl-bearing silicate minerals. Geochimica Et Cosmochimica Acta. 44
(2), 173-182.

Hedenquist, J. W., and Lowenstern, J. B. (1994). The role of magmas in the formation of
hydrothermal ore deposits. Nature, 370(6490), 519-527.

Hedenquist, J. W., Arribas, A. and Reynolds, T. J. (1998) Evolution of an intrusion-centered


hydrothermal system; Far Southeast-Lepanto porphyry and epithermal Cu-Au deposits,
Philippines. Economic Geology. 93 (4), 373-404.

Heinrich, C. A. (2005). The physical and chemical evolution of low-salinity magmatic fluids
at the porphyry to epithermal transition: a thermodynamic study. Mineralium Deposita,
39(8), 864-889.

Heinrich, C. A., Driesner, T., Stefánsson, A., and Seward, T. M. (2004). Magmatic vapor
contraction and the transport of gold from the porphyry environment to epithermal ore
deposits. Geology, 32(9), 761-764.

Hildreth, W., and Moorbath, S. (1988). Crustal contributions to arc magmatism in the Andes
of central Chile. Contributions to mineralogy and petrology, 98(4), 455-489.

Hunt, G. (1977) Spectral signatures of particulate minerals in the visible and near
infrared. Geophysics. 42 (3), 501-513.

16
Chapter One: Literature review

Lowell, J. D. and Guilbert, J. M. (1970) Lateral and vertical alteration-mineralization


zoning in porphyry ore deposits. Economic Geology. 65 (4), 373-408.

Moore, D. M. and Reynolds, R. C. (1997) Quantitative analysis. In: Moore, D. M. and


Reynolds, R. C. (eds.). X-ray Diffraction and the Identification and Analysis of Clay
Minerals. 2nd edition. Oxford, New York, Oxford university press Oxford. 298-329.

O’Neil J. R. and Kharaka T. K. (1976) Hydrogen and oxygen isotope exchange reactions
between clay minerals and water. Geochemica et Cosmochimica Acta. (40) 241-246.

Peacock, M. A. (1931) Classification of igneous rock series. The Journal of Geology. 54-67.

Perelló, J., Cox, D., Garamjav, D., Sanjdorj, S., Diakov, S., Schissel, D., Munkhbat, T. and Oyun,
G. (2001) Oyu Tolgoi, Mongolia: Siluro-Devonian Porphyry Cu-Au-(Mo) and High-
Sulfidation Cu Mineralization with a Cretaceous Chalcocite Blanket. Economic Geology. 96
(6), 1407-1428.

Petit, S., Madejová, J., Decarreau, A. and Martin, F. (1999) Characterization of octahedral
substitutions in kaolinites using near infrared spectroscopy. Clays and Clay Minerals. 47
(1), 103-108.

Richards, J. P. (2003) Tectono-Magmatic Precursors for Porphyry Cu-(Mo-Au) Deposit


Formation. Economic Geology. 98 (8), 1515-1533.

Savin, S. M. and Epstein, S. (1970) The oxygen and hydrogen isotope geochemistry of clay
minerals. Geochimica Et Cosmochimica Acta. 34 (1), 25-42.

Seedorf, E., Dilles, J. H., Proffett, J. M., Einaudi, M. T., Zurcher, L., Stavast, W. J. A., Johnson, D.
A. and Barton, M. D. (2005) Porphyry deposits: characteristics and origin of hypogene
features. In: Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J. and Richards, J. P.
(eds.). Economic Geology 100th anniversary volume. 251-298.

Sillitoe, R. H., and Hedenquist, J. W. (2003). Linkages between volcanotectonic settings, ore-
fluid compositions, and epithermal precious metal deposits. Special Publication-Society of
Economic Geologists, 10, 315-343.

Sverjensky, D. A., Hemley, J. and d'Angelo, W. (1991) Thermodynamic assessment of


hydrothermal alkali feldspar-mica-aluminosilicate equilibria. Geochimica Et Cosmochimica
Acta. 55 (4), 989-1004.

17
Chapter One: Literature review

Taylor, H. P. J. (1997) Oxygen and hydrogen isotope relationships in hydrothermal mineral


deposits. In: Barnes, H. L. (ed.). Geochemistry of hydrothermal ore deposits. 3rd edition. New
York, NY, John Wiley and Sons. 229-302.

USGS (2017) Porphyry copper deposits: interactive map. Retrieved


from https://mrdata.usgs.gov/mineral-resources/porcu.html

Van der Meer, F. D., Van der Werff, H. M., van Ruitenbeek, F. J., Hecker, C. A., Bakker, W. H.,
Noomen, M. F., van der Meijde, M., Carranza, J. M., Boudewijn de Smeth, J. and Woldai, T.
(2012). Multi-and hyperspectral geologic remote sensing: A review. International Journal
of Applied Earth Observation and Geoinformation, 14(1), 112-128.

Wilkinson, J. J. (2013) Triggers for the formation of porphyry ore deposits in magmatic
arcs. Nature Geoscience. 6 (11), 917-925.

18
Chapter Two: Introduction

2 Chapter Two: Introduction

2.1 Generic model for porphyry copper formation


The formation of economic porphyry Cu deposits in the Earth’s crust involves a series of
processes that occur at a variety of depths (Wilkinson 2013). At the largest scale, and most
distal to final mineralisation, is the evolution of mantle-derived, basaltic magmas at, or
near, the base of the lithosphere into the intermediate to felsic magmas that typify arcs.
Subsequently, these oxidized magmas are thought to migrate into volumetrically smaller
magma chambers at shallower levels in the crust. This process is inferred to occur by
sequential dyking events and the eventual development of a variably homogeneous
batholith, such as that at Yerington (Dilles 1987). Cooling and crystallisation of these
batholiths result in an increasing water content (from ~3 wt. % to ~6 wt. %), although
significant contributions of water and chlorine from the upper crust are possible
(Chambefort et al. 2013). Cupola failure within the batholith leads to the emplacement of
fluid-saturated intrusions that can focus fluid flow from the deeper batholith into a small
volume of rock, with magma volumes of 36-100 km3 thought to be necessary to source the
water and metals required for an economic porphyry deposit (Shinohara and Hedenquist
1997). The partitioning of metals and sulphur into an exsolving hydrothermal fluid
ultimately leads to precipitation of metal sulphides in a variety of generally inverted cup-
shaped ore shells (Weis et al. 2012). The loci of porphyry emplacement are inferred to
coincide with pre-existing crustal weaknesses (Tosdal and Richards 2001).

The emplacement of porphyry intrusions within the shallow crust drives large-scale
hydrothermal circulation, producing large volumes of hydrothermally-altered rocks. These
alteration patterns provide a useful vectoring tool for exploration in porphyry systems (Fig.
2.1, Sillitoe 2010). The metal-rich fluids may migrate kilometres from the source intrusion,
and can form associated epithermal systems at a significant distance, such as the Lepanto
high-sulphidation orebody which is located up to ~4 km from the Far Southeast porphyry
Cu and Au deposit (Hedenquist et al. 1998). A final important process is that of the uplift
and erosion of the host rock sequence. This process is vital for the exhumation of deposits
to depths at which the metal can be extracted economically. During uplift and erosion,
enrichment by supergene processes can upgrade previously sub-ore grade material into
oxide ore, and elevate well-mineralised zones to exceptional grades (>1.5 wt. % Cu).

19
Chapter Two: Introduction

Figure 2.1: Schematic of the typical hydrothermal alteration pattern encountered in porphyry-type ore deposits.
These deposits have large alteration footprints; however, the distribution of each type may vary significantly in
extent and form from deposit to deposit. From Sillitoe (2010).

Exsolved hydrothermal fluids form zoned alteration assemblages (Lowell and Guilbert
1970). The highest temperature and most proximal to the deeper magma chamber is the
sodic-calcic alteration. This alteration forms at depths between 3.5 to >6 km and is
generally devoid of sulphides (Dilles and Einaudi 1992). Potassic alteration is characterised
by K-silicates (K-feldspar and/or biotite) typically with accessory magnetite and anhydrite
(Lowell and Guilbert 1970, Sillitoe 2010). This alteration assemblage occurs within the core
of the system dominated by magmatic fluids, typically brines (Ulrich et al. 2002). As these
magmatic fluids move away from central up flow zones, the fluids cool and acidify (i.e.
activity of H+ increases). This acidification is reflected in the more distal alteration which is
dominated by feldspar-destructive sericitic alteration (Giggenbach 1984) which can
involve mixing with meteoric fluids (Reynolds and Beane 1985) and/or just magmatic
fluids (Harris and Golding 2002). As the name suggests, this alteration is dominated by
sericite (finely crystalline muscovite mica), quartz and pyrite. This alteration can be
structurally controlled at shallower levels and dominates the roots of high-sulphidation
epithermal deposits (Sillitoe 1999). The argillic and advanced argillic alteration zones,
formed from highly acidic fluids, typically sit above the porphyry centre (Fig. 2.1) or, in
telescoped deposits where exhumation during hydrothermal activity has superimposed
later fluid flow events onto the core of the system, it may overprint earlier, proximal

20
Chapter Two: Introduction

alteration types. Propylitic alteration is peripheral to the higher temperature potassic and
sericitic alteration assemblages and is characterised by chlorite, epidote and carbonate.

The development of clays in porphyry systems, particularly in telescoped deposits where


clays and ore zones are intimately associated, is important in mines for primarily two
reasons: clays affect rock strength so have significant geotechnical impact on slope stability
(Wyllie and Mah 2004) and can significantly influence recovery of copper during mineral
processing (Bulatovic et al. 1998).

2.2 The importance of argillic alteration


Porphyry deposits contribute the largest share of copper (Cu) production in the world
(Gerst 2008, Fig. 2.2). These ores are typically large-tonnage and low-grade and their
increasing importance as copper resources is largely attributed to improvements in
extraction methods, particularly froth-flotation from in the late 19th century (Bunyak
2000). The development of solvent extraction and electrowinning in the 1970’s made
recovery from low grade oxide ores viable (Kordosky 2002). Gold (Au) and/or
molybdenum (Mo) are commonly produced as a by-product of copper concentration by this
method.

Figure 2.2: Percentage of global copper production in relation to classified resource type (Gerst 2008).

A major problem that impacts the economics of copper processing is the contamination of
ores by clay minerals. Clays have an adverse effect on chalcopyrite recovery (Uribe et al.
2016), by reducing grindability and flotation efficiency (Bulatovic 2007). Clay minerals,
which are sub 2 micrometres in size, are easily entrained during floatation and are known

21
Chapter Two: Introduction

to be a significant problem in the processing of porphyry-type mineralization (Jorjani et al.


2011).

Currently, the only method of treatment of high-clay ores is through blending (Peng and
Zhao 2011), as the precise mechanism by which clay minerals affect froth stability is still
not well understood (Farrokhpay and Bradshaw 2012). Even small amounts of clay can
significantly reduce recovery (Fig. 2.3); thus, understanding their distribution within an ore
deposit has a direct impact on the profitability of prospective and current mining
operations.

Figure 2.3: Copper recovery from porphyry ores with increasing clay content. Adapted from Bulatovic et al.
(1998).

2.3 Case study of the Cerro Corona deposit


Cerro Corona, a porphyry Cu-Au system located in the Hualgayoc district of northern Peru,
provides an opportunity to investigate several aspects of hydrothermal alteration,
particularly argillic alteration. Although argillic alteration typically forms relatively late in
the hydrothermal evolution, its origin is controversial and potentially multi-stage. Argillic
alteration is well developed at Cerro Corona but has several possible origins – related to
the primary magmatic-hydrothermal system driven by porphyry emplacement, linked to a
later epithermal overprint because such activity is evident from vein systems in the district

22
Chapter Two: Introduction

(Yanacocha, Tantahuatay, Sipan and La Zanja; Gustafson et al. 2004), or from a late
weathering due to uplift of the Andes.

Preliminary studies produced grade control maps in which polygonal areas were
characterised by grade and refractory mineral content. These maps indicate that areas of
moderate to high clay content (>5%) occur in the east, southeast and within the western
“barren core” of the deposit (denoted by the warm colours in Fig. 2.4-A). These are zones
in which copper recovery is negatively affected with <75% Cu recovery (Fig. 2.4-B).

The use of locally-available hydrothermal clay minerals as a material for use in tailings dam
construction at Cerro Corona is a novel idea that has the potential to reduce the impact of
open pit mining on the environment by reducing the overall carbon footprint. For this
purpose, specific properties are required, specific to clay mineral mixtures. Thus,
understanding the distribution of clay types in the system is important for evaluating the
potential of this approach.

Cerro Corona also presents a unique opportunity to study the bulk mass changes and
element transfers that accompany different stages of porphyry alteration. This is because
fluid flow was highly focused within an intrusive complex emplaced into relatively
impermeable limestones and where there was little variation in the mineralogy or
geochemistry of the igneous host rocks involved. This means that chemical mass transfer
calculations can be done both in detail on selected hand specimens and on a deposit scale
in 3D because of the consistent composition protolith. Furthermore, the isotopic
composition of magmatic waters and likely meteoric compositions are sufficiently distinct
so that the relative roles of these fluids in the alteration system can be assessed effectively.

23
Chapter Two: Introduction

Figure 2.4: Maps of polygons coloured by: A) clay content; and B) copper recovery. Reproduced from quality
control maps (Uzategui and Ayala 2012).

24
Chapter Two: Introduction

The primary questions this thesis aims to resolve are:

1. What are the overall bulk mass changes and net gains and losses of major, minor
and trace components that occur during different stages of hydrothermal alteration
and how do these patterns inform our understanding of the evolution of porphyry
hydrothermal systems?
2. What are the mineral assemblages in the argillic alteration zones and can different
alteration facies be defined that reflect different processes or different stages of an
evolving process?
3. Was the argillic alteration formed from cooling magmatic fluids or heated meteoric
fluids?

The thesis comprises of nine chapters, including this introductory chapter. Chapter one
encompasses the current understanding of hydrothermal alteration and mass balance in
porphyry systems. Chapter two (this chapter) outlines the objectives of this thesis and
places it in the context of current research. Chapter three covers the regional geology and
places the Hualgayoc district in the context of the development of the northern Peruvian
margin. Chapter four focuses on the Cerro Corona deposit, describing the temporal
evolution of the deposit from both magmatic and hydrothermal perspectives, and produces
an updated model for the evolution of the deposit. Chapter five evaluates the mass transfers
that occurred during hydrothermal alteration and models these major component changes
in three dimensions. This chapter improves upon the mass transfer calculations by
propagating error through these calculations, allowing for more confidence in the
interpretation of mass transfer results. Chapter six evaluates the distribution of clay
minerals by utilisation of two methods of infrared spectroscopy, a handheld IR
spectrometer and a core scanner. These results are used to model clay mineral distribution
patterns and propose a fluid evolution model. Chapter seven applies theoretical single-
mineral oxygen isotope geothermometry in kaolinite and illite to calculate the formation
temperatures. These temperatures are utilised to estimate fluid compositions for argillic
alteration. Chapter eight combines the preceding chapters into a discussion of the controls
of the hydrothermal alteration at Cerro Corona and presents a model for the fluid processes
involved in the argillic alteration. Chapter nine condenses the observations and
interpretations from the preceding chapters into a genetic geological model of Cerro
Corona and attempts to answer the questions posed above.

25
Chapter Two: Introduction

2.4 References
Bulatovic, S. M. (2007) Flotation of Copper Sulfide Ores. In: Bulatovic, S. M. (ed.). Handbook
of flotation reagents: chemistry, theory and practice: volume 1: Flotation of sulfide
ores. Netherlands, Elsevier. 235-291.

Bulatovic, S. M., Wyslouzil, D. M. and Kant, C. (1998) Operating practices in the beneficiation
of major porphyry copper/molybdenum plants from Chile: Innovated technology and
opportunities, a review. Minerals Engineering. 11 (4), 313-331.

Bunyak, D. (2000). To Float or Sink: A Brief History of Flotation Milling. The Mining History
Journal. 7 35-44.

Chambefort, I., Dilles, J. H. and Longo, A. A. (2013) Amphibole geochemistry of the


Yanacocha Volcanics, Peru: Evidence for diverse sources of magmatic volatiles related to
gold ores. Journal of Petrology. 54 (5), 1017-1046.

Dilles, J. H. (1987) Petrology of the Yerington Batholith, Nevada; evidence for evolution of
porphyry copper ore fluids. Economic Geology. 82 (7), 1750-1789.

Dilles, J. H., and Einaudi, M. T. (1992). Wall-rock alteration and hydrothermal flow paths
about the Ann-Mason porphyry copper deposit, Nevada; a 6-km vertical reconstruction.
Economic Geology, 87(8), 1963-2001.

Farrokhpay, S. and Bradshaw, D. J. (2012) Effect of clay minerals on froth stability in


mineral flotation: a review. 26th International Mineral Processing Congress: Innovative
Processing for Sustainable Growth-Conference Proceedings. 24 - 28 September, New Dehli,
India. 4601-4611.

Gerst, M. D. (2008) Revisiting the cumulative grade-tonnage relationship for major copper
ore types. Economic Geology. 103 (3), 615-628.

Giggenbach, W. F. (1984) Mass-Transfer in Hydrothermal Alteration Systems - a


Conceptual-Approach. Geochimica Et Cosmochimica Acta. 48 (12), 2693-2711.

Gustafson, L. B., Vidal, C. E., Pinto, R., and Noble, D. C. (2004). Porphyry-epithermal
transition, Cajamarca region, northern Peru. Society of Economic Geologists Special
Publication (11), 279-299.

Harris, A. C. and Golding, S. D. (2002) New evidence of magmatic-fluid–related phyllic


alteration: Implications for the genesis of porphyry Cu deposits. Economic Geology, 30 (4)
335-338

26
Chapter Two: Introduction

Hedenquist, J. W., Arribas, A. and Reynolds, T. J. (1998) Evolution of an intrusion-centered


hydrothermal system; Far Southeast-Lepanto porphyry and epithermal Cu-Au deposits,
Philippines. Economic Geology. 93 (4), 373-404.

Jorjani, E., Barkhordari, H. R., Tayebi Khorami, M. and Fazeli, A. (2011) Effects of
aluminosilicate minerals on copper–molybdenum flotation from Sarcheshmeh porphyry
ores. Minerals Engineering 24, 754-759.

Kordosky, G. A. (2002). Copper recovery using leach/solvent extraction/electrowinning


technology: Forty years of innovation, 2.2 million tonnes of copper annually. Journal of the
South African Institute of Mining and Metallurgy, 102(8), 445-450.

Lowell, J.D. and Guilbert, J.M. (1970). Lateral and vertical alteration-mineralization zoning
in porphyry ore deposits. Economic Geology, 65(4), 373-408.

Peng, Y. and Zhao, S. (2011) The effect of surface oxidation of copper sulfide minerals on
clay slime coating in flotation. Minerals Engineering. 24 (15), 1687-1693.

Reynolds, T.J. and Beane R. E. (1985). Evolution of hydrothermal fluid characteristics at the
Santa Rita, New Mexico, porphyry copper deposit. Economic Geology, 80(5), 1328-1347.

Shinohara, H., and Hedenquist, J. W. (1997). Constraints on magma degassing beneath the
Far Southeast porphyry Cu–Au deposit, Philippines. Journal of Petrology, 38(12), 1741-
1752.

Sillitoe, R. H. (1999) Styles of high-sulphidation gold, silver and copper mineralization in


the porphyry and epithermal environments. PacRim ’99 Congress Proceedings, Bali,
Indonesia.

Sillitoe, R. H. (2010) Porphyry Copper Systems. Economic Geology. 105 (1), 3-41.

Tosdal, R. M. and Richards, J. P. (2001) Magmatic and structural controls on the


development of porphyry Cu±Mo±Au deposits. Reviews in Economic Geology. (14), 157-
181.

Ulrich, T., Günther, D., and Heinrich, C. A. (2002) The evolution of a porphyry Cu-Au deposit,
based on LA-ICP-MS analysis of fluid inclusions: Bajo de la Alumbrera, Argentina. Economic
Geology (97), 1889-1920.

Uribe, L., Gutierrez, L., and Jerez, O. (2016). The Depressing Effect of Clay Minerals on the
Floatability of Chalcopyrite. Mineral Processing and Extractive Metallurgy Review, 37(4),
227-235.

27
Chapter Two: Introduction

Uzategui, A. and Ayala, E. (2012) Grade control map of clay minerals at 3820m level.
(1:2000)

Weis, P., Driesner, T. and Heinrich, C. A. (2012) Porphyry-Copper Ore Shells Form at Stable
Pressure-Temperature Fronts Within Dynamic Fluid Plumes. Science. 338 (6114), 1613-
1616.

Wilkinson, J. J. (2013) Triggers for the formation of porphyry ore deposits in magmatic
arcs. Nature Geoscience. 6 (11), 917-925.

Wyllie, D.C. and Mah, C.W. (2004) Rock slope engineering: civil and mining. Spon Press

28
Chapter Three: Regional Geology

3 Chapter Three: Regional Geology

3.1 Introduction
The principal aim of this chapter is to outline the geological setting of the Cajamarca district
the context of the tectonic and magmatic development of the Northern Peruvian margin.
This framework is developed to provide a genetic understanding of the Cerro Corona and
surrounding deposits.

The tectonic structure of northern Peruvian is dominated by NW-SE (Andean) trending


faults. These structures are related to the overall structure of the Cretaceous basin, in which
a series of horst-graben that define the Western and Eastern Peruvian Troughs (Fig. 3.1;
Scherrenberg et al. 2012). These faults not only controlled basin sedimentation, but also
are considered vital structural weaknesses that were exploited during orogenic events and
basin inversions.

Figure 3.1 Northern Peruvian margin during the late Cretaceous. The Western Peruvian Trough is defined by the
Paracas High to the West and Maranon Geanticline to the East. The Normal faults controlled sedimentation
within this miogeocline. These faults were subsequently reactivated during subsequent orogenic events. Adapted
from Scherrenberg et al. 2012.

29
Chapter Three: Regional Geology

Pre-orogenic volcanism in northern Peru is confined to the western margin of the Western
Peruvian Trough (WPT) as documented by the extensive Casma Volcanic Group (Fig. 3.1).
This group comprises ~6-9 000 m sequence of marine volcano-sedimentary sequence of
andesitic pillow lavas, hyaloclastites and tuffs, closely associated with sills and gabbros of
Lower Cretaceous age (~152-100 Ma) (Soler and Bonhomme 1990, Petford and Atherton
1994, Cobbing 1999) hosted in shales (Jacay and Sempere 2005) which crop out along the
coastal region of central Peru, near Lima. These Lower Cretaceous volcanics of calc-alkaline
affinity (Cobbing et al. 1981) suggest that a magmatic arc had developed on the margin of
the WPT by the Cretaceous (Atherton et al. 1983), whilst Cretaceous Limestones were
deposited within the Cajamarca region (Fig. 3.1).

The Cajamarca district lies between the Western and Eastern Cordillera, east of Coastal Belt
and south of the Huancabamba deflection (Fig. 3.2). The Western and Eastern Cordilleras
are separated by the Marañon Fold and Thrust Belt (MFTB), a northwest trending fold and
thrust belt developed in Inca I orogenic event (59-55 Ma) (Mégard 1984). The district has
a significant change of trend of the fold axial planes of the MFTB from northwest to a west-
northwest known as the Cajamarca curvature, developed during a second orogenic stage,
the Inca II orogeny (~44 Ma, Fig. 3.2) (Mégard 1984).

The Coastal Belt is dominated by a ~1600 km trench-parallel coastal batholith, which is


emplaced from ~106 to 60 Ma (Pitcher 1984, Mukasa and Tilton 1984, Cobbing 1999). The
Coastal batholith is comprised of several super units defined by cross-cutting contacts:
Patap (c. 100 Ma); Paccho (98-95 Ma); Santa Rosa (95-85 Ma); a major phase producing
dykes (72-68 Ma); La Mina (69-65 Ma); and San Jeronimo, Puscao, Sayan and Canas (65-60
Ma) (Pitcher 1978). The Coastal belt is separated from the Western Cordillera by low angle
thrust faults south of the Cajamarca District (Fig. 3.2).

30
Chapter Three: Regional Geology

Figure 3.2: Geological terrains of Northern Peru. The Cajamarca region lies west of the Marañon Fold Thrust Belt
(MFTB), within the Western Cordillera. The Cajamarca curvature is defined by the change of the trend of
predominant fold axial planes from northwest to a west-northwest. The Huancabamba deflection is a large-scale
change in fault trend from NW (in Peru) to NE (in Ecuador and Colombia). The inset box (expanded in Fig. 3.5)
shows the area of study. Adapted from Mégard (1984).

Petrographic, geochemical and structural evidence from the volcanic arc in Northern Peru
strongly suggests a gradational plate thickening of the continental crust from 75-24 Ma
followed by melting of the metasomatized lithospheric mantle from 24-5 Ma (Davies 2002).
The subduction of the buoyant Inca Plateau and Nazca Ridge are preceded by porphyry-
related mineralisation which is succeeded by amagmatic, shallow angle subduction.

31
Chapter Three: Regional Geology

3.2 Geological setting of the Cajamarca district

3.2.1 Stratigraphy
The basement of the Western Cordillera, similar to that for most of northern Peru, is
composed of Precambrian to Palaeozoic phyllite and schist of the Marañon Complex
(Scherrenberg et al. 2012). Outcrops of this complex are observed east of the MFTB, in the
Eastern Cordillera (Fig. 3.2). An incomplete sequence of supracrustal Ordovician
conglomerates and quartzites and Carboniferous continental sedimentary rocks overlie the
Marañon Complex. Exhumed Carboniferous plutons, hosted within the Marañon Complex,
occur in the western segment of the Western Cordillera; these were emplaced during the
Hercynian orogeny (Sánchez 2006). Extensional, NW-trending faults have dismembered
these Hercynian plutons and basement (Laubacher and Mégard 1985). These steeply-
dipping faults bound the Permian-Triassic-Jurassic extensional basin into which
conglomerates and sandstones of the Mitú Group were deposited.

The oldest exposed rock units in the Cajamarca District make up the Chicama Formation.
This consists of predominantly volcaniclastic turbidites with black shales and/or coal with
occasional conglomerate (Mégard 1984, Jacay 1992). These late Jurassic rocks acted as
décollement layers upon which the Cretaceous limestones detached during the Inca
Orogeny (Fig. 3.3, Eude 2014).

A deltaic sequence of shales, quartzites and tuffs, known as the Gollarisquizga Group,
disconformably overlies the Jurassic coals and shales (Fig. 3.3). This group thins to
approximately 1-300 m in the northeast of the district (Benavides 1956, Wilson 1963).

The Gollarisquizga Group is overlain by ferruginous sandstones and mudstones


interbedded with calcareous units and finally thin shales, marls and limestones that
comprise the Inca, Chulec and Pariatambo Formations. These units document a series of
marine transgressions during the Albian-Lower Cretaceous across the Western Peruvian
Trough (WPT, Jaillard et al. 2005; Fig. 3.1 and Fig. 3.3).

By the Late Cretaceous, a stable carbonate platform had developed in the WPT (Fig. 3.1).
The Pulluicana Group, a thick sequence of limestones, marls and shales was deposited,
indicative of a shallow sea environment. It is conformably overlain by uniformly-bedded
limestone of the Cajamarca Formation (Benavides 1956, Fig. 3.1 and Fig. 3.3).

The development of the Peruvian orogenic belt to the west of the basin resulted in an
increased supply of clastic material (Benavides 1956, Wilson 1963). By the Santonian (~83

32
Chapter Three: Regional Geology

Ma), much of the basin had filled and conglomerates of the Chota Formation were deposited
(Fig. 3.1 and Fig. 3.3), marking the emergence of the WPT.

Following this, N-NE-directed crustal shortening of the Inca I orogeny was established. This
was accommodated by NW-SE trending folds and over thrusts in the MFTB (Mégard 1984,
Scherrenberg et al. 2014). This event was associated with magmatic quiescence (McKee et
al. 1990) after which subduction-related volcanism shifted eastward with the extrusion of
~1800 metres of pyroclastic flows of the Calipuy Volcanic Group (Fig. 3.3).

Recent geochronological studies have subdivided the Calipuy Volcanic Group into six
eruptive stages separated by hiatuses, synorogenic basins and angular unconformities: the
Tantara (75-55 Ma), Llama (55-42 Ma), Pativilca (42-30 Ma), Calamarca (30-24 Ma),
Calipuy (24-10 Ma) and Negritos (10-3Ma) Formations (Chavez et al. 2012), although the
Calamarca stage is not seen within the Cajamarca district. Overall, there is an eastward
migration of the magmatic arc during this period (Navarro and Rivera 2008).

Upon cessation of magmatism, at around 8 Ma, lake and glacial sediments were deposited
on an angular unconformity. More recent fluvial and alluvial sediments occur in
topographic lows, proximal to modern rivers and streams (Wilson 1984).

33
Chapter Three: Regional Geology

Figure 3.3: Schematic stratigraphic column for the Cajamarca district. Cretaceous stratigraphy adopted from
Wilson (1984). Cenozoic volcanism is incorporated from Chavez et al. (2012). Decollement horizons are inferred
from Eude (2014)

34
Chapter Three: Regional Geology

3.2.2 Principal structures and tectonic development


The Huancabamba deflection marks the change in orientation of the Andean chain at ~6°S
from NW-SE on the southern limb to NE-SW on its northern limb and represents one of the
major Andean oroclines (Fig. 3.2). This deflection is proposed to be the consequence of the
combination of accretion of the allochthonous Amotape-Tahuin continental block onto the
Olmos massif (in the Ecuadorian and Colombian Andes) in the late Cretaceous (~140 Ma),
and the lack of accretion onto the northern Peruvian margin (Mourier et al. 1988; see Fig.
3.2).

To the south of the Huancabamba deflection, three main phases of deformation are
considered critical in the development of the Northern Peruvian Andes (Mégard 1984):

 the Mochica Phase, a Cretaceous orogenic event which terminated rifting and
subsidence in the WPT;
 the Inca Orogeny, the most significant deformation event, which is subdivided into
four pulses of crustal shortening: Inca I (59-55 Ma), Inca II (44-43 Ma), Inca III (26
Ma) and Inca IV (22Ma); and
 the Quechua orogeny, a period spanning from 26 Ma to 4 Ma with three stages of
compression, uplift and magmatism: Quechua I (~17 Ma), Quechua II (8-7 Ma,
Benavides-Cáceres 1999) and Quechua III (~7-4 Ma)

The onset of the Mochica deformation event around the Albian (~100 Ma), related to the
rifting of the South Atlantic Ocean (Steinmann 1929, Mégard 1984, Soler and Bonhomme
1990, Torsvik et al. 2009), resulted in termination of back-arc basin subsidence within the
Western Peruvian Trough (Fig. 3.1; Mégard 1984). During this event, the basin sediments
were folded into NW-SE-trending, open folds (Mégard 1984). It was accompanied by a
switch from volcanism to plutonism (Cobbing 1999) and the emplacement of the Coastal
Batholith.

The Incaic orogenic cycle is the main Andean deformation event recognised throughout the
Andes. It was responsible for tight to isoclinal folding in the MFTB (Mégard 1984) and
coincided with a dramatic increase in subduction rate of the Farallon plate beneath the
South American plate from ~50-55 mm/yr. to >100 mm/yr. as well as a change in direction
of subduction from NNE to NE, based on plate reconstruction models (Fig. 3.4; Pardo-Casas
and Molnar 1987, Pilger 1981).

35
Chapter Three: Regional Geology

Figure 3.4: Average convergence rate between the Nazca and South American plate over time and approximate
plate motion, based on plate reconstruction (Pardo-Casas and Molnar 1987). Note the periods of high
convergence coincide with deformation stages.

The Inca I Orogeny was succeeded by the eruption of thick (~1000 m) rhyolitic ash-flow
tuffs of the Llama Formation at ~55 to 44 Ma (Noble et al. 1990), which are exposed west
of the Puntre Fault (Fig. 3.5). This volcanism is coincident with the development of the
rhyolitic Yatahual caldera south of the Cajamarca district at ~54-47 Ma (Navarro 2012).

The Cajamarca curvature is characterised by a change in strike of fold axial planes in the
Cretaceous sedimentary rocks from NW to WNW (Benavides-Cáceres 1999).
Paleomagnetic studies suggest approximately half of the Cajamarca curvature occurred via
non-rotational mechanisms during the Inca I Orogeny (Mitouard et al. 1992), followed by a
25° anticlockwise rotation during the late Eocene (44-39 Ma; Mitouard et al. 1990, 1992,
Rousse et al. 2003), coincident with the Inca II Orogeny.

Following the further development of the MFTB during the Inca II Orogeny, the Pativilca
magmatic stage commenced (Fig. 3.3 and Fig. 3.5) at around 39 Ma (Noble et al. 1990) with
the extrusion of rhyolitic pyroclastics and dacitic tuffs of the Huambos volcanics, andesitic
lavas of the Yatahual, and lavas, ash and rhyolitic ignimbrites of San Lorenzo-Catan (Fig. 3.5
and Fig. 3.6).

36
Chapter Three: Regional Geology

Figure 3.5: Map of the Cajamarca district. Maps used: Wilson 1984, Noble et al.1990, Macfarlane et al. 1990,
Davies 2002, Longo 2005, Navarro 2012

After the Patlivilca magmatic event, ~32 Ma (Noble et al. 1990), there was a local lull in
magmatism in the Cajamarca district. This correlates with a decrease in subduction rate

37
Chapter Three: Regional Geology

(Fig. 3.4; Pardo-Casas and Molnar 1987, Sébrier and Soler 1991). Emplacement of gabbroic
dykes (29.4±1.4 Ma, Davies 2002) along sub-vertical faults that have NNW-N orientations
near the hinge of the El Galeno anticline may correspond to a period of regional upper
crustal relaxation and/or extension (Davies 2002, Sébrier and Soler 1991). This stage is
coincident with major plate reorganisation of the subducting Farallon Plate and spreading
between the Cocos and Nazca plate (Lonsdale 2005).

The Inca IV orogenic pulse at ~22 Ma (Benavides-Cáceres 1999) is temporally related to a


clockwise rotation of the subducting Nazca plate (Fig. 3.4; Padro-Casas and Molnar 1987)
and the onset of volcanism in the Cajamarca district. This is evidenced by the deposition of
the Chala sequence (Fig. 3.5; Noble et al. 1990) and emplacement of the Aurora Patricia
hornblende porphyry (Fig. 3.6; Davies 2002).

The Miocene-Pliocene Quechua phase of deformation was a result of the resurgence of


rapid convergence (110±8 mm/yr.) of the Nazca Plate and the South American Plate
(Padro-Casas and Molnar 1987), resulting in crustal thickening and uplift (Mégard 1984).
The precise chronology of the Quechua orogeny has been defined from dating in Central
Peru (Mégard 1984), where deformation is the most intense. The nomenclature of the
Quechua orogenic stages is adopted for local deformation events (Table 3.1; Benavides-
Cáceres 1999, Noble et al. 1990), in which it is subdivided into three stages of compression
and uplift (McKee and Noble 1982, Mégard 1984) with extensive magmatism occurring
throughout (Noble et al. 1990).

38
Chapter Three: Regional Geology

Table 3.1: Compilation of the nomenclature and ages of major deformation events in the Cenozoic of Northern
Peru. * nomenclature adopted in this thesis.

Deformation
Nomenclature Reference
age
59-55 Ma Inca I* Davies 2002, Benavides-Cáceres 1999
Davies 2002, Benavides-Cáceres 1999,
43-42 Ma Inca II*
Navarro et al. 2010
30-27 Ma Inca III* Davies 2002, Benavides-Cáceres 1999,
28 Ma Inca III Navarro et al. 2010
Jaillard et al. 2000, Sempre et al. 1990,
28-26 Ma Aymara
Sebrier et al. 1988
24 Ma Inca IV Navarro et al. 2010
~23Ma Quechua I Longo 2005
22 Ma Inca IV* Davies 2002, Benavides-Cáceres 1999
20-12.5Ma Quechua I Mégard 1984, Chiaradia et al. 2009
19.5-17 to 12-
Quechuan deformation McKee and Noble 1982
12.5 Ma
Davies 2002, Benavides-Cáceres 1999,
17 Ma Quechua I*
Navarro et al. 2010
12-11 Ma Quechua II Longo 2005
12-10 Ma Quechua II Chiaradia et al. 2009
9.5-8.5 Ma Quechua II Mégard 1984
Davies 2002, Benavides-Cáceres 1999,
7-8 Ma Quechua II*
Navarro et al. 2010
6.95 Ma Quechua III Mégard 1984
Quaternary phase
2.0 Ma Petford and Atherton 1992
deformation*

The Quechua I orogenic stage is a poorly-characterised deformation event (Megard et al.


1984) to which monoclinal folds and reverse reactivation of Incaic structures (such as the
Puntre Fault?) have been attributed. Regionally, the majority of the deformation occurred
during a pulse at 17-15 Ma (Jaillard et al. 2000). This stage of deformation is possibly the
most important in relation to intrusion-centred mineralisation and Calipuy volcanism (Fig.
3.5 and Fig. 3.6; Longo 2002).

Seismic tomography of the Andes has confirmed plate reconstruction models which
suggested that sections of the subducting Nazca plate beneath the South American plate are
flat (Fig. 3.7; Gutscher et al. 1999, 2000, Eude 2014) as initially proposed by Pilger (1981).
The Cajamarca district is situated directly above such a region of flat slab subduction,
interpreted to be due to the subduction of the young (thus buoyant) Inca Plateau. Based on
current subduction rates, the Inca Plateau is inferred to have been subducted beneath the

39
Chapter Three: Regional Geology

Cajamarca district at ~12-8 Ma (Rosenbaum et al. 2005). The subduction of this plateau is
coincident with cessation of the Calipuy Volcanic Group.

The La Zanja-Tantahuatay, Sipan-Hualgayoc and Cajamarca are “structural lineaments”


along which major economic mineralisation appears to have developed (Fig. 3.5). The
Chicama-Yanacocha “structural corridor” is a NE-trending zone, bounded by the Sipan-
Hualgayoc and Cajamarca lineaments, and is considered to be a major control in the
development of magmatic-hydrothermal deposits in the Cajamarca district (Longo et al.
2010, Davies 2002). However, the definition of the corridor from satellite imagery (Quiroz
1997) is somewhat arbitrary and its timing and controls of development remain unclear.

The Quechua II phase of deformation is prominent in the Cajamarca district and lasted from
~8-7 Ma (Table 3.1; Benavides-Cáceres 1999). NW-SE-trending faults were sheared in a
dextral sense due to N-S shortening (Megard 1984). Active deformation during this stage,
including the onset of imbricate thrusting (Shaw et al. 1999), gradually shifted eastward,
toward the Subandean Fold and Thrust Belt (SFTB; Jaillard et al. 2000). During this time,
the region underwent significant crustal thickening and uplift (Noble et al. 1990) and
magmatism ceased (Fig. 3.6).

The Quechua III deformation event at 7-4 Ma was characterised by the development of the
SFTB to the east of the Cajamarca district, a feature documented throughout the Andes
(Megard et al. 1984, Jaillard et al. 2000). This stage of E-W shortening caused gentle folding
in the Calipuy Group in southern and central Peru (McKee and Noble 1982), but very little
deformation has been attributed to this stage in the Cajamarca region.

40
Chapter Three: Regional Geology

Figure 3.6: Sequence of magmatic events in the Cajamarca district. Adapted from Davies 2002

41
Chapter Three: Regional Geology

3.3 Igneous geochemistry of Cajamarca District


Petrographic, geochemical and isotopic studies (Macfarlane et al. 1994, Davies 2002, Longo
2005, Chiaradia et al. 2009) suggest a gradational plate thickening of the continental crust
from 75-24 Ma followed by melting of the metasomatized lithospheric mantle from 24-5
Ma. Magmatism associated with porphyry and epithermal mineralisation in the district is
temporally associated with the Calipuy Volcanic Group (Fig. 3.6).

The Calipuy Volcanic Group is represented in at least 16 volcanic edifices and has been
subdivided into five temporal and spatial magmatic stages (Chavez et al. 2012): Llama (55-
42 Ma); Pativilca (42-30 Ma); Calamarca, to the south of the Cajamarca district (30-24 Ma);
Calipuy (24-10 Ma); and Negritos (10-3Ma) magmatic stages (Fig. 3.6 and Fig. 3.6).

The Llama magmatic event (55-42 Ma) is dominated by sub-alkaline mafic to intermediate
magmas at Cerro Montana, Yatahual caldera, Llama and La Carpa (Fig. 3.6, Fig. 3.8 and Fig.
3.10; INGEMMET 2008). The calc-alkaline affinity of these rocks suggests that arc
magmatism was fully established by this stage. Modelling suggests these melts were
derived from partial melting of an olivine-pyroxene mantle, with an amphibole and garnet-
poor residue (Davies 2002). Sr/Y versus Y plot show that these melts followed a “normal”
plagioclase fractionation trend (Fig. 3.9), suggesting that suppression of plagioclase
crystallisation had not occurred in these hydrous magmas.

Magmas of the Pativilica stage (~42-30 Ma) are more evolved than those of the Llama stage
with more felsic compositions represented in the Huambos and San Lorenzo-Catan
Formations. These have chemistry controlled by the fractionation of plagioclase and
pyroxene (Fig. 3.8; Davies 2002).

42
Chapter Three: Regional Geology

43
Figure 3.7: Cross section through subduction zone through the Northern Peruvian margin. Note that location of Miocene magmatism is located on the margin of a thickened continental
crust. Adapted from Eude 2014
Chapter Three: Regional Geology

There is little magmatism within the Cajamarca district during the Calamarca magmatic
event (30-24 Ma). Localised extension in the upper crust allowed for the emplacement of
gabbroic dykes, in the anticline near the El Galeno deposit (Davies 2002). These Oligocene
gabbro dykes (~29 Ma) are significantly more aluminous and Fe-rich than SiO2 equivalent
mafic rocks of the Llama calc-alkaline suite (Fig. 3.8). εNd isotope data suggest these
magmas are of mantle origin, without significant crustal contamination (Davies 2002). This
is supported by flat REE patterns, despite the high ∑REE. Modelling suggests that the melts
were derived from ~5% partial melting of a primitive mantle source with olivine- and
pyroxene-dominated residual mineralogy (Davies 2002). These magmas are inferred to be
the parental melts for Miocene magmatism (Davies 2002).

The Chala sequence marks the end of the magmatic quiescence (Noble et al. 1990), the first
eruptive unit of the Calipuy Group (24-10 Ma). This comprises of a sequence of
conglomerates, volcaniclastics and tuffs. This is contemporaneous with the emplacement
of the Minas Conga monzodiorite-diorite intrusions (Noble et al. 1990, Davies 2002). These
intermediate, calc-alkaline rocks are interpreted to have formed from ~50% melting of a
basaltic source, similar in composition to that of the gabbroic dykes (Davies 2002). Trace
element modelling indicates a residual mineralogy, or fractionation, of amphibole and
clinopyroxene, with minor garnet (Fig. 3.8 and Fig. 3.9; Davies 2002). Oscillatory zoning
and relatively shallow depth of crystallisation of amphibole suggest these hydrous calc-
alkaline magmas may have stagnated at ~6 km depth prior to final intrusion (Davies 2002).

44
Chapter Three: Regional Geology

45
Chapter Three: Regional Geology

46
Figure 3.8: Major element vs. SiO2 variation diagrams for igneous rocks the Cajamarca district. Data compiled from Davies 2002; Longo 2005; Chiaradia et al. 2009; Navarro et al. 2010;
INGEMMET 2008; and this study. Colours correlate to magmatic pulses in figure 3.6, whereby brown represents the Tantara (75-55 Ma); maroon represents the Llama (55-42 Ma); orange
represents the Pativilca (42-30 Ma); light orange (gabbroic dykes) represent the Calamarca (30-24 Ma); yellow, tan, pink represent the evolving Calipuy (24-10 Ma); and green represent the
Negritos magmatic pulse (10-3 Ma).
Chapter Three: Regional Geology

Figure 3.9: A) Sr/Y versus Y and B) La/Yb vs. Yb plots of Cenozoic igneous rocks in the Cajamarca district. Note
that the shift to adakite-like signature occurs around the time of Calipuy Volcanism (i.e. ~24 Ma). Figure 3.8
contains key of symbols and list of references for data.

47
Chapter Three: Regional Geology

Early Miocene intrusions at Aurora Patricia, Michiquillay and Salvia are similar in
composition (Fig. 3.8). They straddle the low- to medium-Fe and medium- to high-K calc-
alkaline fields (Fig. 3.8). Decreasing CaO and MgO with increasing SiO2 and increasing
incompatibles suggest magma chemistry was controlled by fractionation of pyroxene and
amphibole. Trace element models suggest a similar source and residual mineralogy and
slightly higher (~55%) degrees of melting of the basaltic source to those proposed for the
earlier Calipuy Group rocks, except with a greater garnet control (Davies 2002).

The Yanacocha pre-, syn- and post-mineralisation magmatism is the most comprehensively
studied igneous suite within the Cajamarca District (Longo 2005, Chiaradia et al. 2009,
Longo et al. 2010 and references therein). These detailed petrographic, geochronological,
geochemical and radiogenic isotope studies reveal that the magmas have a complex origin
and were originally derived from a deep, hydrous basaltic or andesitic melt that evolved
through amphibole – clinopyroxene ± garnet fractionation with increasing thickness of the
continental crust (Chiaradia et al. 2009, Longo et al. 2010). These magmas show variable
stages of recharge, assimilation and fractional crystallisation and are suggested to have
mixed with andesitic melts, derived from melting of continental crust (Longo et al. 2010).
Thermobarometric studies of amphiboles from the Yanacocha district support the
hypotheses of multiple magma chambers, in which mafic hydrous magmas supply sulphur,
fluorine, copper and gold and assimilated high level crustal rocks impart the majority of
water and chlorine required to form intrusion-centred economic mineralisation
(Chambefort et al. 2013). High-pressure fractional crystallisation followed by melting and
assimilation of lower crust is thought to drive the adakite-like signature (Castillo et al.
1999; Macpherson et al. 2006) in the Cajamarca district (Chiaradia et al. 2009).

The Negritos magmatic pulse (10-3 Ma) is dominated by explosive acidic volcanism and
marks the final magmatic stage of the Cajamarca district (Fig. 3.6). These magmas were
hydrous and highly oxidised with phenocrysts of quartz, plagioclase, biotite, hornblende
and titanite (Longo 2005). Flat angle subduction is inferred to prevent recharging of deeper
magma reservoirs and thus marks the end of magmatism by ~8 Ma.

48
Chapter Three: Regional Geology

Figure 3.10: A) Total Alkali versus silica diagram of regional igneous rocks and B) AFM diagram indicating a
linear calc-alkaline trend. Figure 3.8 contains figure key.

49
Chapter Three: Regional Geology

3.4 Uplift and erosion


A key feature that can be overlooked in studies of porphyry deposits is the rate of uplift and
erosion. However, this aspect is important because it determines the preservation potential
of deposits as well as the extent of telescoping that will occur during the evolution of a
system (Sillitoe 2010).

In the Western Cordillera of Northern Peru, the main unconformities are related to the
Incaic, Peruvian and Quechuan tectonic phases. The most significant period of uplift and
erosion in the Cerro Corona area is from the Quechua II phase to present.

Reviews of the uplift and erosion history of the Central and Northern Andean margin
(Gregory-Wodzicki 2000) and Central Peru (Garver et al. 2005) show that precise
calculation of uplift and erosion rates is complicated. However, the consensus is that by the
Pliocene the Northern Andes were <40% of modern day elevation, but by 2.7 Ma was at
modern elevation (Gregory-Wodzicki 2000). A similar history is recorded to the south,
where by the elevation at ~18 Ma was < 50 % of the current elevation (Garver et al. 2005).
The rate of uplift dramatically increased at ~10 Ma, with most uplift occurring in the last
5-6 Ma (Garver et al. 2005).

Rates of uplift appear to have increased from ~200-300 m/My in the late Miocene, to a
remarkable rate of ~600-3000 m/My during the Pliocene. This is broadly confirmed by
apatite and zircon (U-Th)/He thermochronology which suggest that no significant
exhumation of plutonic rocks in the Western Cordillera occurred until the late Miocene
(Michalak 2013, Michalak et al. 2016).

3.5 Discussion
The formation of the magmas that went on to form porphyry ore deposits in the Cajamarca
district is associated with the thickening of the continental crust. This thickening is thought
to have driven higher pressure fractional crystallisation, accompanied by melting and
assimilation of lower crust, to produce an adakite-like geochemical signature (Castillo et al.
1999; Macpherson et al. 2006; Chiaradia et al. 2009). The ascension of these melts into
shallow level magma chambers was controlled by pre-existing structures and the upper
crustal stress regime, which was predominantly compressive at the time owing to the high
convergence rate of the subducting Nazca plate. Melting and assimilation of crustal material
is thought to have contributed chlorine and water to sulphur- and metal-bearing, mantle-
derived melts (Chambefort et al. 2013) - crucial for formation of these types of deposits.

50
Chapter Three: Regional Geology

Shallow level degassing of these melts can explain the high sulphidation deposits seen in
the district, which predominantly occur along pre-existing structural weaknesses. The
emplacement of porphyry deposits was controlled by local structures, such as those
observed at Michiquillay (Hollister et al. 1974) and El Galeno (Davies et al. 2005).
Interestingly, epithermal-type mineralisation is located trenchward of the porphyry belt.
This may be due to either the state of stress in the upper crust during mineral deposition
or variable rate of post-deposition uplift.

3.6 References

Atherton, M. P., Pitcher, W. S. and Warden, V. (1983) The Mesozoic marginal basin of central
Peru. Nature. 305 (5932), 303-306.

Benavides-Cáceres, V. (1956) Cretaceous system in northern Peru. Bulletin of the American


Museum of Natural History. 104, 353-494.

Benavides-Cáceres, V. (1999) Orogenic evolution of the Peruvian Andes: the Andean


cycle. In: Skinner, B.J. (ed.) Geology and Ore Deposits of the Central Andes. Special
Publication. (7), 61-107.

Castillo, P. R., Janney, P. E. and Solidum, R. U. (1999) Petrology and geochemistry of


Camiguin Island, southern Philippines: insights to the source of adakites and other lavas in
a complex arc setting. Contributions to Mineralogy and Petrology. 134 (1), 33-51.

Chambefort, I., Dilles, J. H. and Longo, A. A. (2013) Amphibole geochemistry of the


Yanacocha Volcanics, Peru: Evidence for diverse sources of magmatic volatiles related to
gold ores. Journal of Petrology. 54 (5), 1017-1046.

Chavez, L., Navarro, P. and Mamani, M. (2012) Características geoquímicas del volcanismo
cenozoico (Grupo Calipuy) en el cuadrángulo Santiago de Chuco, Norte del Perú. Lima,
Sociedad Geológica del Perú.

Chiaradia, M., Merino, D. and Spikings, R. (2009) Rapid transition to long-lived deep crustal
magmatic maturation and the formation of giant porphyry-related mineralization
(Yanacocha, Peru). Earth and Planetary Science Letters. 288 (3), 505-515.

Cobbing, E. (1999) The Coastal Batholith and other aspects of Andean magmatism in
Peru. Geological Society, London, Special Publications. 168 (1), 111-122.

51
Chapter Three: Regional Geology

Cobbing, E., Pitcher, W. S., Wilson, J. J., Baldock, J. W., Taylor, W. P., McCourt, W. J. and
Snelling, N. J. (1981) The geology of the Western Cordillera of northern Peru. London, United
Kingdom, Overseas Memoirs, Institute of Geological Sciences.

Davies, R. C. I. (2002) Tectonic, magmatic and metallogenic evolution of the Cajamarca


mining district, Northern Peru. PhD thesis. James Cook University.

Davies, R. C. and Williams, P. J. (2005) The El Galeno and Michiquillay porphyry Cu-Au-Mo
deposits: geological descriptions and comparison of Miocene porphyry systems in the
Cajamarca district, northern Peru. Mineralium Deposita. 40 (5), 598-616.

Eude, A. (2014) La croissance des Andes centrales du nord du Pérou (5-9° S): Propagation
d’un prisme orogénique dans un contexte d’héritage tectonique et de subduction
plane. PhD thesis. Université Toulouse III Paul Sabatier.

Garver, J., Reiners, P., Walker, L., Ramage, J. and Perry, S. (2005) Implications for Timing of
Andean Uplift from Thermal Resetting of Radiation-Damaged Zircon in the Cordillera
Huayhuash, Northern Peru. The Journal of Geology. 113 (2), 117-138.

Gregory-Wodzicki, K. M. (2000) Uplift history of the Central and Northern Andes: a


review. Geological Society of America Bulletin. 112 (7), 1091-1105.

Gutscher, M. A., Olivet, J. L., Aslanian, D., Eissen, J. P. and Maury, R. (1999) The “lost inca
plateau”: cause of flat subduction beneath Peru? Earth and Planetary Science Letters. 171
(3), 335-341.

Gutscher, M. A., Spakman, W., Bijwaard, H. and Engdahl, E. R. (2000) Geodynamics of flat
subduction: Seismicity and tomographic constraints from the Andean margin. Tectonics. 19
(5), 814-833.

Hollister, V. F. and Sirvas, E. B. (1974) The Michiquillay porphyry copper


deposit. Mineralium Deposita. 9 (3), 261-269.

INGEMMET. (2008) GR-4: Volcanismo Cenozoico (Grupo Calipuy) y su Asociación a los


Yacimientos Epitermales, Norte del Perú. Lima, Peru, Instituto Geológico Minero y
Metalúrgico.

52
Chapter Three: Regional Geology

Jacay, J. (1992) Estratigrafía y sedimentología del Jurá sico curso medio del valle de Chicama
y esbozo paleogeográ fico del Jurá sico Cretá ceo del nor Perú (6ᵒ30'-8ᵒ Sur). PhD thesis.
Universidad Nacional Mayor de San Marcos.

Jacay, J. and Sempere, T. (2005) Emplacement levels of the Coastal Batholith in central
Peru. Proceedings of the 6th International Symposium of Andean Geodynamics. Barcelona,
Spain, Instituto Geológico y Minero de España. 397-399.

Jaillard, E., Hérail, G., Monfret, T., Díaz-Martínez, E., Baby, P., Lavenu, A. and Dumont, J.
(2000) Tectonic evolution of the Andes of Ecuador, Peru, Bolivia and northernmost Chile.
In: Cordani, U. G., Milani, E. J., Thomaz, A. and Campos, D. A. (eds.). Tectonic Evolution of
South America. Rio de Janeiro, Brazil, 31st International Geological Congress. 481-559.

Jaillard, E., Bengtson, P. and Dhondt, A. V. (2005) Late Cretaceous marine transgressions in
Ecuador and northern Peru: a refined stratigraphic framework. Journal of South American
Earth Sciences. 19 (3), 307-323.

Laubacher, G. and Mégard, F. (1985) The Hercynian basement: a review. In: Pitcher, W. S.,
Atherton, M. P. and Cobbing, E.J., Beckinsale, R.D. (eds.). Magmatism at a plate edge: the
Peruvian Andes. Glasgow, Blackie. 29-37.

Longo, A. A. (2005) Evolution of volcanism and hydrothermal activity in the Yanacocha


mining district, northern Peru. Oregon State University.

Longo, A. A., Dilles, J. H., Grunder, A. L. and Duncan, R. (2010) Evolution of Calc-Alkaline
Volcanism and Associated Hydrothermal Gold Deposits at Yanacocha, Peru. Economic
Geology. 105 (7), 1191-1241.

Lonsdale, P. (2005) Creation of the Cocos and Nazca plates by fission of the Farallon
plate. Tectonophysics. 404 (3), 237-264.

Macfarlane, A. W., Marcet, P., LeHuray, A. P., Petersen, U. (1990) Lead isotope provinces of
the Central Andes inferred from ores and crustal rocks. Economic Geology. 85 (8), 1857-
1880.

Macfarlane, A. W., Prol-Ledesma, R. and Conrad, M. E. (1994) Isotope and Fluid Inclusion
Studies of Geological and Hydrothermal Processes, Northern Peru. International Geology
Review. 36 (7), 645-677.

53
Chapter Three: Regional Geology

Macpherson, C. G., Dreher, S. T. and Thirlwall, M. F. (2006) Adakites without slab melting:
high pressure differentiation of island arc magma, Mindanao, the Philippines. Earth and
Planetary Science Letters. 243 (3), 581-593.

McKee, E. H. and Noble, D. C. (1982) Miocene volcanism and deformation in the western
Cordillera and high plateaus of south-central Peru. Geological Society of America Bulletin. 93
(8), 657-662.

McKee, E. H. and Noble, D. C. (1990) Cenozoic tectonic events, magmatic pulses, and base-
and precious-metal mineralization in the central Andes. In: Ericksen, G. E., Caas Pinochet,
M. T. and Reinemud, J. A. (eds.). Geology of the Andes and Its Relation to Hydrocarbon and
Mineral Resources. Circum Pacific Council Publications for Energy and Mineral Resources.
189.

Mégard, F. (1984) The Andean orogenic period and its major structures in central and
northern Peru. Journal of the Geological Society. 141 (5), 893-900.

Michalak, M. (2013). Exhumation Of The Peruvian Andes; Insights From Mineral


Chronometers. PhD Thesis, University Of Californa, Santa Cruz, USA.

Michalak, M. J., Hall, S. R., Farber, D. L., Audin, L., and Hourigan, J. K. (2016). (U-Th)/He
thermochronology records late Miocene accelerated cooling in the north-central Peruvian
Andes. Lithosphere, 8(2), 103-115.

Mitouard, P., Kissel, C. and Laj, C. (1990) Post-Oligocene rotations in southern Ecuador and
northern Peru and the formation of the Huancabamba deflection in the Andean
Cordillera. Earth and Planetary Science Letters. 98 (3), 329-339.

Mitouard, P., Laj, C., Mourier, T. and Kissel, C. (1992) Paleomagnetic study of an arcuate fold
belt developed on a marginal orogen: The Cajamarca deflection, northern Peru. Earth and
Planetary Science Letters. 112 (1-4), 41-52.

Mourier, T., Laj, C., Mégard, F., Roperch, P., Mitouard, P. and Medrano, A. F. (1988) An
accreted continental terrane in northwestern Peru. Earth and Planetary Science Letters. 88
(1-2), 182-192.

Mukasa, S. B. and Tilton, G. R. (1984) Lead Isotope Systematics in Batholithic Rocks of the
Western and Coastal Cordilleras, Peru. In: Harmon, R. S. and Barreiro, B. A. (eds.). Andean
Magmatism: Chemical and Isotopic Constraints. Boston, MA, Birkhäuser Basel. 180-189.
54
Chapter Three: Regional Geology

Navarro, P. (2012) Nuevas dataciones en el segmento sur de Cajamarca. Sociedad Geológica


del Perú.

Navarro, P. and Rivera, M. (2008) Stratigraphy of the synorogenic Cenozoic volcanic rocks
of Cajamarca and Santiago de Chuco, northern Peru. 7th International Symposium on
Andean Geodynamics. Nice, France pp.369-372.

Navarro Colque, P. A., Rivera Porras, M. A., Miguel, M., and Robert, W. (2010). Geología y
metalogenia del Grupo Calipuy (Volcanismo Cenozoico) segmento Santiago de Chuco, norte
del Perú [Boletín D 28].

Noble, D. C., McKee, E. H., Mourier, T. and Mégard, F. (1990) Cenozoic stratigraphy,
magmatic activity, compressive deformation, and uplift in northern Peru. Geological Society
of America Bulletin. 102 (8), 1105-1113.

Pardo-Casas, F. and Molnar, P. (1987) Relative motion of the Nazca (Farallon) and South
American plates since Late Cretaceous time. Tectonics. 6 (3), 233-248.

Petford, N. and Atherton, M. P. (1994) Cretaceous-Tertiary volcanism and syn-subduction


crustal extension in northern central Peru. Geological Society, London, Special
Publications. 81 (1), 233-248.

Petford, N. and Atherton, M. P. (1992) Granitoid emplacement and deformation along a


major crustal lineament: the Cordillera Blanca, Peru. Tectonophysics. 205 (1), 171-185.

Pilger, R. H. (1981) Plate reconstructions, aseismic ridges, and low-angle subduction


beneath the Andes. Geological Society of America Bulletin. 92 (7), 448-456.

Pitcher, W. S. (1978). The anatomy of a batholith President's anniversary address 1977.


Journal of the Geological Society, 135(2), 157-182.

Pitcher, W. S. (1984) Phanerozoic Plutonism in the Peruvian Andes. In: Harmon, R. S. and
Barreiro, B. A. (eds.). Andean Magmatism: Chemical and Isotopic Constraints. Boston, MA,
Birkhäuser Basel. 152-167.

Pitcher, W. S., Atherton, M. P., Cobbing, E. J., and Beckinsale, R. D. (1985). Magmatism at a
plate edge: the Peruvian Andes (p. 328). Blackie.

55
Chapter Three: Regional Geology

Quiroz, A. (1997) El corredor estructural Chicama-Yanacocha y su importancia en la


metalogenia del norte del Perú. Congreso Peruano de Geología IX. Lima, Peru, Sociedad
Geológica del Perú. 149-154.

Rosenbaum, G., Giles, D., Saxon, M., Betts, P. G., Weinberg, R. F. and Duboz, C. (2005)
Subduction of the Nazca Ridge and the Inca Plateau: Insights into the formation of ore
deposits in Peru. Earth and Planetary Science Letters. 239 (1), 18-32.

Rousse, S., Gilder, S., Farber, D., McNulty, B., Patriat, P., Torres, V. and Sempere, T. (2003)
Paleomagnetic tracking of mountain building in the Peruvian Andes since 10 Ma. Tectonics.
22 (5)

Sánchez, A. (2006) Las rocas graníticas y la mineralización aurífera asociada, en la


Cordillera Oriental del Norte del Perú, 6 30'-7 30'. Revista Del Instituto De Investigaciones
De La Facultad De Geología, Minas, Metalurgia y Ciencias Geográfica. 9 (17), 19-29.

Scherrenberg, A. F., Jacay, J., Holcombe, R. J. and Rosenbaum, G. (2012) Stratigraphic


variations across the Marañón fold-thrust belt, Peru: Implications for the basin architecture
of the West Peruvian Trough. Journal of South American Earth Sciences. 38, 147-158.

Scherrenberg, A.F., Holcombe, R.J., and Rosenbaum, G. (2014) The persistence and role of
basin structures on the 3D architecture of the Marañón Fold-Thrust Belt, Peru. Journal of
South American Earth Sciences 51, 45-58.

Sébrier, M. and Soler, P. (1991) Tectonics and magmatism in the Peruvian Andes from late
Oligocene time to the present. Geological Society of America Special Papers. 265, 259-278.

Shaw, J. H., Bilotti, F. and Brennan, P. A. (1999) Patterns of imbricate thrusting. Geological
Society of America Bulletin. 111 (8), 1140-1154.

Sillitoe, R. H. (2010) Porphyry Copper Systems. Economic Geology. 105 (1), 3-41.

Soler, P. and Bonhomme, M. G. (1990) Relation of magmatic activity to plate dynamics in


central Peru from Late Cretaceous to present. Geological Society of America Special
Papers. 241173-192.

Steinmann, G., 1929, Geologie von Peru: Heidelberg, Carl Winters Universitàts
Buchhandlung, 448.

56
Chapter Three: Regional Geology

Torsvik, T. H., Rousse, S., Labails, C., and Smethurst, M. A. (2009). A new scheme for the
opening of the South Atlantic Ocean and the dissection of an Aptian salt basin. Geophysical
Journal International, 177(3), 1315-1333.

Wilson, J. J. (1984) Geología de los cuadrángulos de Jayanca (13-d), Incahuasi (13-e),


Cutervo (13-f), Chiclayo (14-d), Chongoyape (14-e), Chota (14-f), Celendín (14-g),
Pacasmayo (15-d), Chepén (15-e). Lima, Peru, Instituto Geológico Minero y Metalúrgico.

Wilson, J. J. (1963) Cretaceous stratigraphy of central Andes of Peru. American Association


of Petroleum Geologists, Bulletin. 47 (1), 1-34.

57
Chapter Four: Geology of Cerro Corona

4 Chapter Four: Geology of Cerro Corona

4.1 Introduction

4.1.1 The Hualgayoc mining district


Cerro Corona lies 25 km north of the Yanacocha epithermal Au district in the Cajamarca
province of Perú and 1.5 km west of Hualgayoc village at an elevation of 3900 m. The
Hualgayoc district has been an active mining area since Incan times and was a centre of
silver mining during colonial times (Soule 2014). Polymetallic replacement and vein
deposits have produced a large proportion of the Peruvian gold and silver in the 18th and
19th centuries. Typical ore-grade mineralisation in the vein deposits extended for up to 1
km along strike and to ~400 m depth (Macfarlane and Petersen 1990). The vein deposits
crosscut Cretaceous sediments and Miocene intrusions (Sillitoe 1997) and are closely
associated with stratiform Pb-Zn-Ag-Cu replacement deposits, confined to the Lower
Chulec, Inca and Farrat Formations (Macfarlane et al. 1990).

The Cerro Corona porphyry-style mineralisation was first documented by BRGM in 1979
and extensive drilling by Barrick from 1994-1996 led to the first feasibility study of the
deposit. This draft feasibility study indicated an oxide reserve of 6.85 Mt at 1.34g/t Au and
84 Mt of sulphide ore at 1.02 g/t Au and 0.58% Cu. Further diamond drilling by RGC Limited
in 1998 marginally increased the oxide reserves to 6.98 Mt at 1.3 g/t and 133 Mt of sulphide
ore at 0.91g/t Au and 0.52% Cu. In December 2003, Gold Fields bought 80.7% of the Cerro
Corona deposit from Sociedad Minera Corona S.A and transferred it to the local subsidiary,
Gold Fields La Cima S.A., by July 2004 (Gold Fields 2009). By December 2008, the mine was
running at full production. The current* total resource is 146.2Mt at 0.79 g/t Au and 138.8
Mt at 0.43% Cu (*Gold Fields 2014 report). It is currently an active open pit mine.

4.1.2 Local stratigraphy and intrusive ages


Geological mapping of the Hualgayoc District in the 1980’s and 1990’s, together with a
compilation of historical mine information, enabled a coherent Cretaceous stratigraphy to
be developed (Macfarlane and Petersen 1990; Fig. 4.1).

The oldest exposed unit in the area is the Farrat Formation, characterised as an
approximately 800 m thick quartzite, siltstone and shale unit. This is overlain by a highly
reactive, ferruginous (and in part pyritiferous) calcareous sandstone unit, the Inca

58
Chapter Four: Geology of Cerro Corona

Formation, estimated to be ~30 m thick excluding the abundant sills emplaced within it.
The overlying, ~250 m thick, fossiliferous Chulec Formation is composed of calcareous
marls and shales, which is conformably overlain by the Pariatambo Formation, a ~250 m
thick sequence of thinly bedded limestone and shale. The Yumagual and Mujarrun
Formations form a 1000 m thick sequence of well-bedded limestones making up the
Pullicana Group. The Yumagual Formation contains a higher proportion of shales and marls
than the Mujarrun Formation (Macfarlane and Petersen 1990) and hosts the Cerro Corona
Diorite (Fig. 4.1 and Fig. 4.2). The Quillquiñan Group, consisting of the Romirón and Coñor
Formations, makes up a 100-200 m thick sequence of shales and marls with few beds of
limestone, which conformably overlies the Pullicana Group. The Cajamarca Formation
overlies the Quillquiñan Group and comprises of ~350 m of well-bedded uniform
limestone, similar to that of the Pullicana Formation. The concordant Celendin Formation
is composed of nodular lime-muds interbedded with marls and shales. The thickness of this
formation is variable due to subsequent erosion; locally this formation is ~150 m (Fig. 4.1).

59
Chapter Four: Geology of Cerro Corona

Figure 4.1: Stratigraphy of the Hualgayoc district. Adapted from Macfarlane and Petersen (1990) and Castillo
(2001).

The Inca orogeny produced open upright folding of the Cretaceous sediments and the
development of an erosional unconformity between the Celendin Formation and the
overlying Calipuy Volcanic Group (Chapter 3, section 3.3). The intrusions within the
60
Chapter Four: Geology of Cerro Corona

Hualgayoc district are temporally associated with the Calipuy magmatic stage (24-10 Ma),
within the larger Calipuy Volcanic Group. Quaternary gravels represent the youngest
sediments, deposited within the valleys.

Numerous porphyritic andesite sills (“Yanacocha Sills”) intrude the Inca and Chulec
Formations to the northeast of Hualgayoc village (Fig. 4.2). K-Ar ages on K-feldspar indicate
that these are the oldest intrusions in the district (16.8 ± 0.4 Ma; Table 4.1).

The “San Miguel” intrusive event comprises the Yanacocha Sill, together with several other
sills (San Miguel, Cerro Jesus, and Cerro San Jose) (Macfarlane and Petersen 1990). These
hornblende- and clinopyroxene-bearing rocks overlap in age with the Cerro Corona quartz
diorite (Table 4.1) represent a magmatic event spanning the period ~14.3-10.5 Ma
(Macfarlane and Petersen 1990). The duration of hydrothermal activity proximal to these
intrusions continued for approximately ~1.05 ± 0.37 My. No geochemical data for the San
Miguel intrusions exists in the literature.

The contact between the Corona Quartz Diorite and the host limestones is sharp (Fig. 4.3),
with marble development limited to within ~100 m from the igneous contact. Proximal to
the Corona intrusion, there are stratiform (manto) replacement deposits that trend NW-SE
and vein deposits with a NE-SW trend. These manto deposits are confined to the Chulec,
Inca and Farrat formations. Pb isotope studies of the ore minerals suggest that the ore-
forming solutions were probably sourced from Miocene intrusions and that mineralisation
formed during a single event, or multiple, closely spaced events (Macfarlane and Petersen
1990).

61
Chapter Four: Geology of Cerro Corona

Figure 4.2: Regional map of NW-SE trending Hualgayoc anticline with Miocene intrusions hosted within the
hanging wall. Map adapted from Wilson 1984, Noble et al. 1990, Macfarlane and Petersen 1990, Castillo 2001.

62
Chapter Four: Geology of Cerro Corona

Figure 4.3: Sharp intrusive contacts between the biotite quartz diorite 1 intrusion and the host limestone (A, B
and E) along with interpretations (A’, B’ and E’). Chalcopyrite and pyrite mineralisation occurs within silty
limestone (C and D) with interpretations (C’ and D’).

63
Chapter Four: Geology of Cerro Corona

The Cerro Hualgayoc biotite-bearing rhyodacite is distinct from the preceding San Miguel
intrusions in that it is significantly more felsic and contains no hornblende and/or
pyroxene phenocrysts. This is the final intrusive event in the district (Table 4.1).

Table 4.1: Radiometric ages of intrusions in the Hualgayoc district. Cerro is abbreviated to “Cᵒ”.

Sample
Intrusion Mineral Age (Ma) Reference Method
type
Yanacocha sill Andesite K-feldspar 16.8 ± 0.4 Macfarlane et al.K-Ar
(1994)
Cᵒ Corona Quartz Zircon 14.4 ± 0.1 James (1998) U-Pb
diorite
Cᵒ Jesus Andesite Whole rock 14.3 ± 0.7 Borredon (1982) K-Ar
Cᵒ Coymolache Andesite Zircon 14.3 ± 0.1 James (1998) U-Pb
Atahualpa mine Felsic sill Muscovite 13.48 ±Macfarlane et al.K-Ar
0.19 (1994)
Cᵒ Corona K-altered Biotite 13.35 ±Macfarlane et al.K-Ar
diorite (hydrothermal) 0.27 (1994)
Cᵒ Tantahuatay Andesite Zircon 13.2 ± 0.2 James (1998) U-Pb
Cᵒ San Jose Andesite Muscovite 13 ± 0.4 Macfarlane et al.K-Ar
(1994)
Cᵒ Tantahuatay Andesite? Hypogene 12.4 ± 0.4 Macfarlane et al.K-Ar
alunite (1994)
Cᵒ Coymolache Andesite Whole rock 11.8 ± 0.6 Borredon (1982) K-Ar
Cerro Jesus Andesite Muscovite 10.29 ± 0.2 Macfarlane et al.K-Ar
(1994)
Cᵒ Hualgayoc Rhyodacite Biotite 9.05 ± 0.21 Macfarlane et al.K-Ar
(magmatic) (1994)
Cᵒ Hualgayoc Biotite- Biotite 7.9 ± 0.3 Macfarlane et al.K-Ar
bearing (magmatic) (1994)
rhyodacite
Los Mantos Andesite Whole rock 7.2 ± 0.35 Borredon (1982) K-Ar

64
Chapter Four: Geology of Cerro Corona

4.1.3 Structural geology


Structural data relating to the Hualgayoc district are sparse, mostly derived from mapping
done in the 80’s and 90’s (Wilson 1984, Noble et al. 1990, Macfarlane et al. 1990). These
maps show a NW-SE trending anticline, located in the hanging wall block SW of a large
thrust fault with a similar trend (Wilson 1984, Fig. 4.2). The NW-SE trending thrust fault
has a similar trend and proximity to the Puntre Fault, a major fault of the Marañon Fold and
Thrust Belt (MFTB, Chapter 3, Fig. 3.5). The Hualgayoc Anticline is assumed to have formed
during the Inca I orogenic stage (~59-55 Ma), in which a dramatic increase in subduction
rate of the Farallon plate beneath the South American plate caused NNE-NE-directed
shortening of Cretaceous sedimentary rocks (Mégard 1984, Pardo-Casas and Molnar 1987).
The mineralised Miocene intrusions were emplaced SW of the hinge of the anticline, within
the Inca and Chulec Formations (Fig. 4.2; Macfarlane et al. 1990).

A notable an angular unconformity exists between the Llama (~44.2 Ma) and Porculla (~39
Ma) Formations, to the northwest of the district (Noble et al. 1990). This indicates there
was some degree of tilting between these volcaniclastic deposits, broadly corresponding to
the Inca II deformation event (Chapter 3, Table 3.1, ~43-42 Ma; Mégard 1984). Both these
formations also contain monoclinal folds, suggestive of further, post-Porculla NE-SW
shortening (Noble et al. 1990). This deformation is poorly dated and may be due to one of
the Inca III, IV or Quechua I deformation events, somewhere between ~30-17 Ma.

Three fracture systems crosscut Miocene intrusions: 1) Sub-vertical NW-SE-trending


fractures; 2) N80°E-trending fractures that dip 65°NW; and 3) N45°E-trending fractures
with a 65°NW dip (Macfarlane et al. 1990).

The relationship between these fracture sets has yet to be completely resolved (Macfarlane
et al. 1990). Detailed studies of the polymetallic mineralisation within the area surrounding
Hualgayoc suggested that mineralisation occurred within all fracture networks
(Macfarlane and Petersen 1990). Larger scale maps show that mineralisation is distributed
along NW- trending vertical fractures and faults (Fig. 4.5).

Regional geological maps surrounding Cerro Corona show inconsistent nomenclature,


crosscutting relationships and displacements (Castillo 2001; Uzategui and Jacay 2010;
Uzategui and Ayala 2012). Initial mapping of the Hualgayoc district recorded a set of
Andean (NW-SE) striking faults that transect the “Corona Diorite”. These Andean-trending
faults (Carlos, Jane, Enrique and Stud Faults, Fig. 4.5) show minor displacement, perhaps a
few 10’s of meters. They are variable in length, and terminate abruptly. Trans-Andean, NE-

65
Chapter Four: Geology of Cerro Corona

SW-trending faults (Manuela, Olga, Violeta and Rosi Faults) crosscut and sinistrally displace
the Andean-trending faults and the biotite quartz diorite by ~20-40 m (Uzategui pers.
comm. 2011).

4.1.4 The Cerro Corona intrusive complex


The first maps of the Cerro Corona deposit indicated a single diorite intrusion, hosted
within Cretaceous limestone (Wilson 1984). This intrusion was subsequently described as
a “disseminated porphyry Cu-Mo centre”, capped by an intense network of closely spaced
quartz veins (Macfarlane et al. 1990).

Subsequent attempts to constrain the intrusive history (Sillitoe 1997; James 1998) were
thwarted by difficulties in delineating intrusive contacts due to the obliteration of primary
igneous textures and contacts by argillic alteration at shallow depths, and the lack of deep
drill holes at their time of study. Due to this, the geological models proposed by James
(1998) and Sillitoe (1997) utilised the presence of quartz veins to distinguish intrusions.
Simplistically, early- and syn-mineralisation intrusions, termed “Quartz Diorite 1” (QD1),
contains stockwork quartz veining which is not seen in the post-mineral intrusions “Quartz
Diorite 2” (QD2) porphyries (James 1998).

Later field study of the Corona Diorite suggested multiple intrusive events were present:
two early porphyries, one syn-mineralisation and two post-mineralisation (Fig. 4.4, Sillitoe
1997). However, proof of these relationships was not presented, nor named beyond stating
their timing relative to mineralisation.

Figure 4.4: Schematic model of intrusions at Cerro Corona. Sillitoe 1997


66
Chapter Four: Geology of Cerro Corona

Figure 4.5: Surficial geology map of the Cerro Corona Diorite intrusion. The mapped “porphyry diorite” is QD1
intrusion and “Barren Core(s)” are QD2 intrusions, adapted from Uzategui and Jacay 2010.

The detailed petrographic studies of the two main intrusive events, QD1 and QD2,
determined that these porphyries are texturally and mineralogically similar and are
distinguished from one another only by the presence or absence of quartz veins (James
1998).

The pre-mineralisation QD1 contains two populations of euhedral, equant plagioclase


phenocrysts: a small (0.5-2 mm) size population, comprising up to 20% of the rock; and a
larger (5-9 mm) size population, comprising up to 10% of the rock (James 1998). Large (~8
mm), euhedral biotite phenocrysts with K-feldspar inclusions constitute ~8% of the rock.
Euhedral hornblende phenocrysts (1-4 mm) make up ~5-8% of the rock. However, very
little/no primary hornblende is preserved at shallow depths due to extensive alteration.
Minor anhedral magnetite (1-2 mm) constitutes <1% of the rock, with rare, anhedral quartz
phenocrysts (<2 mm). A biotite-rich quartzofeldspathic groundmass constitutes ~ 50 % of
the rock with amounts of zircon.

Syn- and post-mineralisation quartz diorite porphyry intrusions (QD 2) have similar
mineralogy to the pre-mineralisation quartz diorite (QD1), with similar proportions of
plagioclase, biotite and hornblende phenocrysts in a quartzofeldspathic matrix. They have
67
Chapter Four: Geology of Cerro Corona

a slightly higher proportion of magnetite than QD1 and contain apatite and epidote (James
1998). Two intrusions of QD2 were thought to exist, based on the presence/absence of
stockwork veining. However, their relationships to one another and to the QD1 were poorly
defined (James 1998).

The QD1 intrusion is approximately 800 x 850 m in diameter (Fig. 4.5) and extends beyond
the deepest drill depth of 2500 above sea level (~1400 m below surface). The contact with
the Cretaceous limestone is sharp and well defined in the south, west and east (Fig. 4.3)
with partial recrystallization to marble within ~10 m of the contact (Uzategui and Ayala
2012).

Surface mapping delineates contact between QD1 and the limestone. This is supported by
drill core intersects at depth, with the exception of the northern margin. The lack of drill
intersects between QD1 and limestone has resulted in a geological model that is open to
the north.

The QD 2 porphyries have undulatory, chilled contacts with QD1. Phenocryst sizes increase
away from this undulatory margin. The “barren” QD2 occasionally contains sub-rounded
xenoliths of QD1, which result in a marginal increase in grade. QD2 crosscuts quartz-
magnetite, K-feldspar and biotite veinlets seen within the QD1. Quartz ± pyrite veinlets
occasionally crosscut the QD2 intrusions, with no significant effect on grade. Carbonate
veins commonly crosscut QD1 and QD2, but without significant changes in ore grade
mineralisation. Two intrusive phases of QD2 (“Barren Core West” and “Barren Core East”)
are easily distinguished from the QD1 (“Corona Diorite”) in the open pit owing to their
absence of veins.

4.1.5 Hydrothermal alteration


Four alteration assemblages with associated vein sequences have been characterised: K-
silicate, sericite-chlorite-clay, quartz-sericite-pyrite, and orange clay (Table 4.2, James
1998).

The oldest K-silicate alteration observed at the deepest levels is characterised by the
presence of biotite, K-feldspar, quartz, magnetite and chalcopyrite. Up to ~ 5% of K-silicate,
rock contains veins, with variably developed vein envelopes. These veins are broadly
categorised as biotite, K-feldspar and magnetite veins which range from small (<4 mm)
discontinuous, wormy veinlets to wide (~12 mm) straight-edged veins (James 1998). A
second, less intense K-silicate alteration event is inferred to occur relating to the Quartz
Diorite 2 intrusion.
68
Chapter Four: Geology of Cerro Corona

The K-silicate alteration grades into, and is overprinted by, sericite-chlorite clay alteration.
This is volumetrically the most extensive alteration assemblage in the Cerro Corona deposit
(James 1998). Plagioclase is altered to sericite and chlorite, preserving the porphyry
texture. Shreddy biotite appears to have been altered to chlorite, whereas magmatic biotite
is generally unaltered.

Quartz-sericite-pyrite was also interpreted to be a volumetrically significant alteration


assemblage at Cerro Corona. This texture-destructive assemblage has sharp contact with
other alteration assemblages. It replaces the primary mineralogy with ~50 vol. % of quartz,
~30% sericite, ~5% pyrite, 5-15% calcite and ~10% clay (James 1998). Chalcopyrite (~2-
5%) and trace molybdenite occur in veins with associated quartz-sericite-pyrite envelopes.

The final alteration stage at Cerro Corona was loosely termed “Orange Clay” alteration,
composed of “equal” proportions of smectite, kaolinite and sericite (James 1998). However,
the relationship of this alteration to other types was unclear due to the swelling properties
of smectite, which cause the drill core to turn to rubble. The orange colour was suggested
to arise from increasing hematite and magnetite proportions due to leaching of hypogene
iron oxides (James 1998).

Table 4.2: Alteration assemblages at Cerro Corona (James 1998).

Minor
Name Key assemblage Associated Veins Style
assemblage
K-silicate Biotite, K-feldspar, Epidote, calcite, Biotite, K-feldspar, Pervasive,
quartz, magnetite, hematite, bornite, magnetite, selective
chalcopyrite pyrite specularite

Sericite- Chlorite, sericite, Hematite, pyrite Quartz-oxide- Pervasive


chlorite-clay calcite, magnetite, sulphide, quartz-
kaolinite, pyrite
chalcopyrite
Quartz-sericite- Sericite, kaolinite, Molybdenite, Quartz-pyrite, Pervasive,
pyrite quartz, pyrite, vivianite? pyrite, calcite selective
calcite
Orange clay Montmorillonite, Pervasive
hematite

69
Chapter Four: Geology of Cerro Corona

4.2 New insights into the intrusive history at Cerro


Corona
Dendritic manganese and iron oxides appear along fracture planes within the limestone,
with minor rhodochrosite. Bench mapping reveals that there is weak metasomatism along
fractures and silty (i.e. permeable) horizons within the Pariatambo Formation, associated
with stratabound precipitation of pyrite, chalcopyrite and minor bornite (Fig. 4.3).

At the time of the initial studies of the intrusive phases at Cerro Corona (see Sillitoe 1997,
James 1998), a detailed investigation of the vein types and their crosscutting relationships
had not been produced, making the preliminary conclusions rather speculative. Now, the
existence of deep drillholes (see Appendix 1 for drill core logs) allow these geological
models to be tested and updated by including observations of intrusive contacts at depth,
and by incorporation of a paragenetic vein study (Appendix 4).

The biotite-quartz diorite intrusions are crosscut and displaced by NE-SW trending faults
(Fig. 4.5). This gives the appearance of elongated intrusions. The Western biotite-quartz
diorite (“Barren core west”) appears elongated in the N-S direction, whereas the eastern
biotite-quartz diorite is elongated in an E-W direction. The western intrusion is
approximately 150 m by 180 m and the eastern biotite-quartz diorite is 370 (E-W) by 280
m (N-S).

Previously unrecognised hornblende-quartz diorite porphyries and andesitic dykes


crosscut the biotite-quartz diorites (Fig. 4.6). These intrusions are modelled as ~50 m thick,
vertical cylindrical dykes; however, the precise vertical extent of these intrusions is not
known. The hornblende-quartz diorite porphyry truncates quartz-feldspar veins within the
QD1 intrusion (Fig. 4.6) as it intrudes. The hornblende-quartz diorite has minor, sinuous K-
feldspar veinlets that are crosscut by minor magnetite-chalcopyrite veinlets. These veinlets
are not seen throughout the hornblende-quartz diorite.

The andesitic dykes have chilled margins against QD2 (Fig. 4.6) and are up to 10 m thick.
These dykes truncate all alteration and mineralisation events and are not part of the main
mineralisation event.

70
Chapter Four: Geology of Cerro Corona

Figure 4.6: Representative examples of the intrusive contacts at Cerro Corona.

71
Chapter Four: Geology of Cerro Corona

4.2.1 Petrography
The main intrusions seen at Cerro Corona have a porphyritic texture, with about 40%
phenocrysts in a quartzofeldspathic groundmass. The proportions of phenocrysts vary
between ~30 to ~50 % of the whole rock. Phenocrysts are predominantly plagioclase,
biotite, and quartz with variable amounts of hornblende. The modal proportions are
estimated from petrography. Utilising phenocrysts and groundmass proportions. these
rocks predominantly plot as granodiorites-tonalites (Fig. 4.7), whereas using the
phenocrysts, these appear to be quartz diorite to diorites.

Figure 4.7: Quartz-Alkali-feldspar-Plagioclase classification diagram with modal mineral proportions.

The groundmass is predominated by anhedral quartz and subhedral alkali feldspar and
plagioclase (Fig. 4.8) which occasionally has an intergrowth of alkali-feldspar and quartz as
a symplectite texture (Fig. 4.9). The common accessory minerals seen are apatite and
zircon, which are usually subhedral to euhedral and are between 500 μm and 30-150 μm,
respectively. Zircon commonly shows multiple generations of growth (Fig. 4.10).

72
Chapter Four: Geology of Cerro Corona

Figure 4.8: Cross-polarised image showing biotite and altered plagioclase phenocrysts in a fine-grained
groundmass of predominantly quartz and feldspar. Scale bar represents 1mm.

Figure 4.9: An example of symplectite texture in sample C12.JL.48. The dominant phenocrysts are plagioclase and
quartz. The field of view is 2mm.

73
Chapter Four: Geology of Cerro Corona

Figure 4.10: An example of zoning within zircon (blue). Sample C12.JL.185. Green mineral is apatite.

4.2.2 Mineral chemistry


Mineral chemistry of major phenocrysts was undertaken using Zeiss EVO 15LS Scanning
Electron Microscope (SEM) with an Energy Dispersive X-ray (EDX) detector as well as
Cameca SX 100 electron microprobe (Appendix 3).

4.2.2.1 Feldspar chemistry

Feldspar phenocrysts are euhedral and range up to ~2 cm in size. They are andesine in
composition, with oscillatory zones of labradorite. The zones of higher Ca content
occasionally contain apatite and magnetite inclusions (Fig. 4.11). In the least altered
porphyry samples, the fine-grained matrix is almost entirely composed of quartz and
feldspar, with minor zircon, apatite and opaque minerals. The feldspars that constitute part
of the matrix range from albite-andesine (Fig. 4.12). These are more readily altered to alkali
feldspar and/or clays than the larger phenocrysts.

74
Chapter Four: Geology of Cerro Corona

Figure 4.11: Back scattered images of plagioclase phenocrysts showing oscillatory zoning with brighter zones
having a more anorthitic composition. A) Zoned plagioclase phenocryst with labradorite zone associated with
apatite and magnetite. Scale bar is 400μm; B) The andesine plagioclase phenocryst in in a quartzofeldspathic
matrix showing brighter zones of labradorite. The sample is from a post-mineralisation hornblende porphyry
dyke. Scale bar is 3mm; C) C12.186 Zoned andesine plagioclase phenocryst within andesite dyke. Brighter zones
are more anorthitic. Scale bar is 2mm

Feldspar inclusions within amphibole are predominantly oligoclase composition, with less
common alkali feldspar (Fig. 4.12 and Fig. 4.13), whereas the inclusions occurring within
biotite are entirely alkali feldspar, ranging from albite to orthoclase (Fig. 4.12 and Fig. 4.13).

75
Chapter Four: Geology of Cerro Corona

Figure 4.12: Electron microprobe data of feldspar compositions in least altered porphyritic rocks. Phenocrysts
are predominantly andesine-labradorite in composition (blue circles), whereas oligoclase composition feldspars
occur within amphibole, and albite-orthoclase compositions occur within biotite phenocrysts.

Figure 4.13: Back scattered electron images of analysed feldspars. Red spots indicate points of microprobe
analysis. A) Poikilitic amphibole phenocrysts with inclusions of apatite, magnetite and feldspar; B) Poikilitic
biotite phenocryst with inclusions of apatite and feldspar; C) Zoned plagioclase phenocryst; D) Poikilitic
amphibole with inclusions of magnetite, apatite and feldspar in a quartzofeldspathic matrix. Scale bar for all
images is 500 μm

76
Chapter Four: Geology of Cerro Corona

4.2.2.2 Biotite chemistry

Biotite phenocrysts in the least altered biotite quartz diorites are euhedral and commonly
occur as biotite “books.” This habit allows for easy distinction from hydrothermal
“shreddy” biotite.

Magmatic biotite phenocrysts are commonly poikilitic, containing inclusions of alkali


feldspar, magnetite and apatite (± rutile and ilmenite) (Fig. 4.14 and Fig. 4.15). These
phenocrysts are commonly altered to chlorite (Fig. 4.14-D).

Figure 4.14: Biotite phenocrysts from least altered biotite quartz diorites and post ore-mineralisation andesite
dykes illustrating variable mineral inclusions and textures. A) Biotite phenocryst from post-mineralisation
andesite dyke with inclusions of plagioclase, apatite and magnetite. Scale bar 500 μm (sample C12.JL.186); B)
Biotite phenocryst from post-mineralisation andesite dyke Scale bar 200 μm (sample C12.JL.183); C) Biotite
phenocryst from hornblende quartz diorite porphyry with inclusions of plagioclase and apatite. Scale bar is 500
μm (sample C12.JL.47), and; D) Biotite phenocryst being altered to pycnochlorite from hornblende quartz diorite
porphyry with inclusions of plagioclase and apatite. Scale bar is 500 μm (sample C12.JL.48)

77
Chapter Four: Geology of Cerro Corona

Figure 4.15: Sample C11.178. Back scattered electron images of poikilitic magmatic biotite with inclusions of A)
alkali feldspar (K0.65Na0.35AlSi3O8) with two magnetite inclusions. Scale bar is 300μm; B) Rutile (darker) exsolution
from ilmenite(light) within biotite phenocryst. Scale bar is 100 μm, and; C) apatite, k-spar (K0.84Na0.16AlSi3O8)
magnetite and albite (Na0.96K0.02Ca0.02AlSi3O8). Scale bar is 300μm

Microprobe analyses of these biotite phenocrysts revealed a variable octahedral aluminium


content with consistent Fe/ (Fe+Mg) (Fig. 4.16). The biotite phenocrysts from post-
mineralisation andesite dykes are noticeably more Fe-rich than biotite phenocrysts from
the quartz diorite porphyries.

Figure 4.16: Biotite mineral chemistry projected onto phlogopite-annite-siderophyllite-eastonite quadrilateral


(Speer 1984). The circles represent biotite chemistry from the hornblende-bearing quartz diorite porphyry and
diamonds represent biotite chemistry from post-mineralisation andesite dykes.

78
Chapter Four: Geology of Cerro Corona

4.2.2.3 Quartz eyes

Quartz eyes range from 4-15 mm, with an average size of 6 mm. The distribution of quartz
phenocrysts is irregular and they are not seen in all hand samples. Where quartz is present,
it has an irregular habit and shows resorption textures (Fig. 4.17).

Figure 4.17: Examples of quartz eyes at Cerro Corona. A) Large quartz eye phenocryst in argillically altered
(kaolinite-illite) quartz biotite diorite (GFD 314, 1337m). Scale bar is 1 cm; B) Cold cathodoluminescence
electron image showing the oscillatory zoning in quartz eye. Scale bar is 1mm, and; C) Sub-rounded quartz eye
with a strongly developed embayment in a fine quartzofeldspathic groundmass. C11.55, GFD 175, 32 m. Scale is
1mm.

4.2.2.4 Amphibole chemistry

Amphiboles from the post-mineralisation andesitic dykes are euhedral, pargasitic in


composition and contain anhydrite and magnetite inclusions (Fig. 4.18), whereas
magnesio-hornblendes from the hornblende-quartz diorite porphyry intrusions are
subhedral with inclusions of feldspar, anhydrite and magnetite (Fig. 4.11).

79
Chapter Four: Geology of Cerro Corona

Figure 4.18: Back scattered images of amphiboles from post- (A and B) and syn- (C and D) mineralisation
intrusions demonstrating textures and inclusions. Red spots are location of electron microprobe analyses. A)
Pargasite amphibole with anhydrite and magnetite inclusions (sample C12.JL.183). Scale bar 200μm; B)
Pargasite amphibole with anhydrite inclusion. (sample C12.JL.186). Scale bar 200μm; C) magnesio-hornblende
with abundant anhydrite, magnetite and feldspar inclusions. (sample C12.JL.180). Scale bar is 500 μm, and; D)
Magnesio -hornblende with abundant anhydrite, magnetite and feldspar inclusions. (sample C12.JL.47). Scale bar
is 500 μm.

Electron microprobe analyses of amphiboles from 5 least altered samples (Appendix 3) of


syn- and post-mineralisation quartz diorite intrusions show that all amphiboles are part of
the calcium amphibole subgroup (as defined by Hawthorne et al. 2012, Fig. 4.19).

80
Chapter Four: Geology of Cerro Corona

Figure 4.19: Classification diagram of calcium amphibole subgroup, where B(Ca+∑M2+)/∑B≥0.75 and BCa/∑B ≥
B∑M2+/∑B. Amphiboles from syn-mineralisation hornblende-quartz diorite porphyry(circles)are predominantly
magnesio-hornblende whereas amphibole from post-mineralisation andesite dykes (diamonds) are pargasite
composition.

4.2.3 Geochemistry of igneous rocks


The Hualgayoc district is almost entirely devoid of published whole rock major or trace
element geochemical data on its igneous rocks. However, it is usually considered a northern
extension of the Cajamarca district, situated within the Chicama-Yanacocha “structural
corridor” (Chapter 3, section 3.2.2, Fig. 3.5: Davies 2002). The geochemistry of the Cerro
Corona intrusions is therefore compared with magmatic events within the context of the
broader Cajamarca district, not just the Hualgayoc district.

In this study, igneous modal mineralogy was utilised to classify intrusive rock types (Le Bas
et al. 1991) with the presence/absence of hornblende, quartz and/or biotite as potential
distinguishing features between porphyries. However, this method of logging provided
little information that helped to delineate different intrusions due to similar mineralogy
and gradational variations within intrusions. Thus, the use of igneous crosscutting
relationships was found to be the best method for distinguishing intrusions, especially at
shallow levels where alteration is intense.

81
Chapter Four: Geology of Cerro Corona

In this way, three principal intrusions were defined within the Cerro Corona deposit:
biotite-quartz diorite porphyries, hornblende-quartz diorite porphyry and minor andesitic
dykes. No samples of the late stage andesitic dykes were collected.

Fifteen samples of least-altered biotite quartz diorite were selected for whole rock analysis.
SGS Laboratory used lithium metaborate fusion and an Inductively Coupled Plasma (ICP) -
Atomic Emission Spectroscopy (AES) finish (Appendix 3). The samples selected represent
the dominant intrusions within the deposit- the “barren core west” (i.e. QD 2) and the first
biotite diorite (i.e. QD1; Fig. 4.20). Although care was taken to select only the least-altered
samples, there are a few quartz veins and occasional sulphides present. All samples have
been normalised to 100 wt. % oxide on a volatile-free basis. This aids in their comparison.

Figure 4.20: Plan section through the Cerro corona deposit at ~3600 m elevation illustrating the sample locations
for the least altered biotite diorites. Note samples occur within the “Western Barren Core” (QD 2) as well as the
initial biotite quartz diorite (QD 1).

82
Chapter Four: Geology of Cerro Corona

The least altered samples of the quartz-biotite diorite porphyry intrusions at Cerro Corona
have SiO2 ranging between 62-70 wt.% and broadly plot within the granodiorite field (Fig.
4.21). The quartz diorites follow a calc-alkaline trend, similar to the Calipuy Volcanic group
(Fig. 4.21), and fall within the medium-K to high-K calc-alkaline fields (Fig. 4.23; Le Maitre
et al. 1989). Major and trace element plots show that the different biotite-quartz diorite
intrusions identified in drillcore are geochemically indistinguishable from one another Fig.
4.21 and Fig. 4.22).

The Las Gordas rhyolite, an undated rhyolite intrusion on the margin of the tailings dam
plots within the rhyolite field and appears to be chemically equivalent to volcanics of the
Negritos Formation (8.2 Ma).

Figure 4.21. Total Alkali versus Silica classification (Le Bas et al. 1991) and tholeiitic vs. calc-alkalic trend plots
(Kuno 1968). The green symbols are samples from the early quartz-biotite diorite (QD1); purple symbols
represent samples from the “Western barren core” (QD2) and black rectangles represent samples from the Las
Gordas rhyolite. Brown symbols represent previously analysed QD1 and QD2 samples (James 1998). Error bars
represent the average analytical errors of analysed samples.

83
Chapter Four: Geology of Cerro Corona

Figure 4.22. Zr/Ti vs. Nb/Y discrimination diagram (after Winchester and Floyd 1977). The key used in this figure
is the same as Fig. 4.21. Error bars represent the average analytical errors of analysed samples. Notice that
samples straddle multiple fields and are within error of one another.

Harker plots confirm the similarity between the Cerro Corona biotite-quartz diorites and
the Calipuy Volcanic Group (Fig. 4.23), particularly TiO2 and CaO, which show linear trends
of decreasing concentration with increasing SiO2. The FeO*, MgO and P2O5 contents of the
Corona intrusions are enriched relative to the Calipuy Group (Fig. 4.23).

The Al2O3, K2O and Na2O data show a lot of scatter (Fig. 4.23), related to the complex
magmatic history proposed for the Calipuy Volcanic Group and Miocene intrusions, as well
as their subsequent hydrothermal alteration.

84
85

Figure 4.23: Harker diagrams illustrating the composition of the Cerro Corona intrusions relative to other
igneous rocks within the Cajamarca district. Key from Figure 4.22. Error bars represent the average
Chapter Four: Geology of Cerro Corona

analytical errors of analysed samples. Red lines distinguish high- med- and low-K
Chapter Four: Geology of Cerro Corona

4.2.4 General alteration facies

4.2.4.1 Potassic alteration

Potassic alteration is typically defined as an assemblage consisting predominantly of K-


feldspar and/or shreddy biotite with accessory phases of anhydrite, magnetite, rutile and
fluorite. This assemblage is commonly associated with magnetite ± chalcopyrite, K-feldspar
veins and intense quartz stockwork, in which alkali feldspar occurs as an alteration
envelope to the vein (Fig. 4.24).

Potassic alteration is the dominant alteration type at depth within the deposit, occurring as
K-feldspar alteration of the groundmass (particularly plagioclase) of porphyritic rocks
and/or as shreddy biotite alteration overprinting hornblende (Fig. 4.24). Near the surface,
this alteration is commonly overprinted by clays.

Potassic alteration is crosscut by the second biotite-quartz diorite (QD2, Fig. 4.6). Clasts of
early biotite-quartz diorite with potassic alteration are seen within the second biotite-
quartz diorite (Fig. 4.6).

Based on the observation that potassic alteration is commonly associated with quartz
stockwork veins at depth, this alteration stage is considered to have been most intense in
the now intensely argillised, early biotite-quartz diorite porphyry (QD1).

The current distribution of orthoclase and secondary biotite provides the best
representation of the preserved potassic alteration (Fig. 4.25). This alteration is pervasive
throughout the early biotite-quartz diorite. The later biotite-quartz diorites (QD2) and late
hornblende show little to no potassic alteration. No potassic alteration is seen in andesite
dykes.

86
Chapter Four: Geology of Cerro Corona

Figure 4.24: Examples of potassically altered samples. Plagioclase is readily altered to alkali-feldspar, hornblende
is altered to shreddy biotite and there is a close association with magnetite ± chalcopyrite ± bornite
mineralisation.
87
Chapter Four: Geology of Cerro Corona

Potassic alteration is crosscut by all other alteration stages. Sericitic alteration bleaches the
pink orthoclase and biotite is altered to green-grey sericite. Pervasive chlorite alteration
was not identified overprinting pervasive potassic alteration, however, at depth, there does
appear to be a gradational contact from potassic to propylitic. Veins associated with
potassic alteration (i.e. veins with alkali-feldspar selvages) are crosscut by veins with
chlorite selvages. At shallower levels, argillic alteration commonly overprints potassic
alteration, such that quartz stockwork veins occur within pervasive argillic alteration.

Figure 4.25: Plan section across the 3810 m level illustrating the logged distribution of K-feldspar. Purple line
represents open pit limit.
88
Chapter Four: Geology of Cerro Corona

4.2.4.2 Propylitic alteration

Previous studies of the propylitic alteration at Cerro Corona considered it to comprise a


sericite-chlorite-clay assemblage (James 1998), in which plagioclase phenocrysts have
been altered to clays and mafic phenocrysts (hornblende and biotite) have been altered to
chlorite. Although this assemblage of alteration minerals is better developed at shallow
levels, and is therefore simple to map out, in reality it is most likely a result of multiple
overprinting events. By disentangling the chloritisation from sericitisation and
argillisation, a more consistent alteration and fluid flow model can be developed.

Propylitic alteration is identified in hand specimen as the alteration of biotite and


hornblende to chlorite. At depth, this chloritisation is inferred to have a gradational contact
with potassic (biotite ± K-feldspar) alteration, in which biotite is altered to chlorite and
quartz (Fig. 4.26).

Mapping the distribution of chlorite maps the distribution of this alteration (Fig. 4.27),
which occurs as small pockets irregularly distributed through the deposit at shallow levels.

89
Chapter Four: Geology of Cerro Corona

Figure 4.26: An example of the gradational contact between intense potassic alteration and intense propylitic
alteration seen at depth (sample C12.JL.220; GFD-314, 1221 m). Note potassic alteration (K-feldspar and biotite)
is associated with anhydrite, fluorite and quartz. Biotite is altered to chlorite and quartz.

90
Chapter Four: Geology of Cerro Corona

Figure 4.27: Plan section across the 3810 m level showing the distribution of logged chlorite.

91
Chapter Four: Geology of Cerro Corona

4.2.4.3 Sericitic alteration

The sericitic alteration is defined as an assemblage composed of predominantly fine-


grained muscovite mica (i.e. sericite), quartz ± pyrite. Muscovite mica has replaced both
feldspar and biotite phenocrysts and groundmass (Fig. 4.28), with magnetite being altered
to pyrite. The porphyritic texture is destroyed during intense sericitisation. There is a
gradational, as well as a crosscutting, relationship with potassic alteration (Fig. 4.28), and
forms selvages to planar quartz-pyrite veins (Fig. 4.29) in which predominantly feldspar is
altered to muscovite mica. Detailed study of this alteration shows that this alteration alters
albitic composition plagioclase (Al=97) to sericite and calcite (Fig. 4.30).

Figure 4.28: Hand samples of sericitised biotite quartz diorite. A) Example of intense sericitisation of plagioclase
to white sericite and biotite to grey sericite. B) Chalcopyrite veinlet with k-feldspar halo grading into sericitic
alteration of plagioclase. Scale bar for both images is 1cm.

92
Chapter Four: Geology of Cerro Corona

Figure 4.29: An example of a planar quartz-pyrite vein associated with sericite alteration (sample C12.JL.46). A)
Sample showing the alteration is confined to the vein selvage and predominantly alters feldspar. B)
Backscattered electron image of the vein and C) the vein selvage. Note the sericitisation alters both plagioclase
and k-feldspar. Mineralogy was confirmed by EDX analyses.

93
Chapter Four: Geology of Cerro Corona

Figure 4.30: A) Hand sample of showing vein with associated sericitisation of plagioclase phenocrysts
(C12.JL.170; GFD-314; 808 m).B) Back scattered electron image of relic plagioclase phenocryst and C) higher
magnified area showing relic albite being altered to sericite and calcite.

The distribution of muscovite mica is used to infer the distribution of sericitic alteration
(Fig. 4.31). The area of highest sericitisation is the central part of the intrusion, with less
abundant sericite located along the margin of the intrusion and in the north.

94
Chapter Four: Geology of Cerro Corona

Figure 4.31: Distribution of muscovite across the 3810 m level at Cerro Corona.

95
Chapter Four: Geology of Cerro Corona

4.2.4.4 Argillic alteration

The argillic alteration is defined by the presence of illite, smectite and kaolinite clay
minerals. These clay minerals have replaced both phenocrysts and groundmass and
overprint all other alteration types, thus this alteration is texture destructive. Clay minerals
were identified in the field using handheld IR spectrometer (TerraSpec ™, Chapter 6). The
most common clays encountered are kaolinite, illite and smectite with localised areas of
dickite, predominantly as a mixed assemblage (Chapter 6). The differentiation into
subgroups is based on the dominant clay mineral present.

Smectite-dominant argillic alteration is associated with calcite and zeolite (Ca- heulandite
or clinoptilolite series; Fig. 4.32). Albite and alkali feldspar are replaced by smectite ± illite.
This alteration is abundant and pervasive at shallow levels and within the eastern margin
and central zone of the open pit (Fig. 4.33). At depth, this alteration appears to be more
closely linked to vein selvages of calcite-zeolite veins (Fig. 4.32).

Figure 4.32: Smectite-dominant argillic alteration overprinting potassic altered biotite quartz diorite (sample
C12.JL.141, GFD-314, 610 m). This smectite is associated with zeolite-calcite vein which crosscuts quartz-
magnetite-anhydrite ± chalcopyrite vein.
96
Chapter Four: Geology of Cerro Corona

Figure 4.33: Plan section across the 3810 m level at Cerro Corona showing the distribution of smectite, as
determined by IR spectroscopy.

97
Chapter Four: Geology of Cerro Corona

Illite-dominant argillic alteration typically occurs as an assemblage of illite and smectite, in


which illite is the dominant clay mineral. This assemblage has replaced feldspar, both in the
groundmass and phenocrysts (Fig. 4.34). This alteration grades into the smectite-dominant
argillic alteration and, as with smectite-dominant alteration, is associated with carbonate
veins and overprints all other alteration.

Figure 4.34: Back scattered image of illitisation of plagioclase phenocrysts and groundmass sample (C12.JL.210
(GFD 314, 1142 m).

Illite appears to be pervasive at shallow levels across the Cerro Corona deposit, although is
less intense than smectite-dominant alteration. It is most intense near igneous contacts
(Fig. 4.35).

98
Chapter Four: Geology of Cerro Corona

Figure 4.35: Plan section across the 3810 m level showing illite distribution, as determined by IR spectroscopy

Kaolinite-dominant argillic alteration is composed of an assemblage of kaolinite ± illite ±


smectite in which kaolinite is the dominant clay mineral. This alteration commonly occurs
with darker quartz veins and “vuggy” quartz, with dickite-dominant patches associated
with abundant pyrite.

Kaolinite-dominant argillic alteration is most intensely developed at shallow levels in the


deposit, within the brecciated contacts of early biotite-quartz diorite intrusion (Fig. 4.36).
99
Chapter Four: Geology of Cerro Corona

There are gradational contacts with illite-dominant and smectite-dominant argillic


alteration.

Figure 4.36: Kaolinite distribution across the 3810 m level as determined by IR spectroscopy.

100
Chapter Four: Geology of Cerro Corona

4.2.5 Vein paragenesis


A matrix of 711 crosscutting vein types (Appendix 4) provided a temporal and spatial
representation of the vein sequence. A visual representation of examples of these
relationships along with their associated alteration are shown in Fig. 4.37.

Figure 4.37 Representative examples of crosscutting relationships at Cerro Corona and their association to
alteration

Sinuous magnetite ± chalcopyrite veinlets, alkali feldspar veins and irregularly- orientated,
quartz-magnetite stockwork veins appear to be the earliest as nearly all other vein types
crosscut them (Fig. 4.37). The relationship between these vein types is generally K-feldspar
-> magnetite -> quartz stockwork, although occasionally the relationship between quartz
stockwork and magnetite is reversed. These veins are commonly associated with an alkali
feldspar selvage, which has replaced the groundmass (Fig. 4.38).

101
Chapter Four: Geology of Cerro Corona

Figure 4.38: Sample C12.JL. 161. False colour montage of quartz-magnetite-chalcopyrite vein associated with
potassic alteration. Black is predominantly quartz. Minerals are confirmed by EDX analyses. Bright red is
magnetite, darker red is chalcopyrite, and green is albite plagioclase and blue is alkali-feldspar. Note that the
matrix is almost completely altered to k-feldspar with the larger phenocrysts being relatively unaltered.

At depth, the feldspar veins are open, irregular and discontinuous, with almost miarolitic
textures (Fig. 4.39). These veins are filled with chalcopyrite and molybdenite ± anhydrite,
all of which are uncommon at shallower levels.

Figure 4.39: K-feldspar open space vein with minor infill of chalcopyrite and molybdenite. Sample C12.JL.168.
GFD 314, 780m

102
Chapter Four: Geology of Cerro Corona

Pyrite-magnetite ± chalcopyrite veins with narrow chlorite selvages crosscut quartz-


magnetite stockwork. These veins are irregular and discontinuous at depth. At shallower
levels they grade into more planar, pyrite-chalcopyrite-magnetite-hematite ± calcite veins,
and eventually hematite-dominant veins (Fig. 4.37), with a larger alteration halo
dominated by chlorite (after biotite) ± smectite (after feldspar).

Pyrite-quartz veins are very planar and associated with strong sericite selvages (Fig. 4.40).
These veins are common throughout the open pit where they crosscut all veins apart from
calcite veins. The selvage grades outward into chlorite and illite-smectite assemblages.

Pyrite-dickite veins crosscut all vein types and are most common at shallow levels. At
depth, dickite is seen in highly brecciated areas.

Black quartz and vuggy quartz are associated with dickite-kaolinite alteration and are
mostly found in the southern and western margins of the QD1 porphyry. These quartz types
crosscut all other veins (Fig. 4.37).

Calcite- and zeolite-bearing veins occur throughout the deposit, associated with smectite ±
illite alteration. The selvages of these veins are comprised of chlorite at depth and grades
outward into smectite ± illite (Fig. 4.37).

In reality, the veins do not always show this simplified, linear time sequence. A single vein
commonly has two or more sequences, because of reopening. An example of this is sample
C12.46 that, at hand specimen scale, appears to be a single quartz-pyrite vein with a sericite
vein selvage (Fig. 4.40). However, detailed SEM-cathodoluminescence imaging of this vein
shows five stages of quartz precipitation, separated by four dissolution events. The first
two stages of quartz precipitation are anhedral with very little/no-internal structure. The
third phase of quartz growth displays a colloform texture. The fourth quartz type is
euhedral and has a zoned, comb-like texture with alternating CL intensity. This is followed
by a period of internal brecciation and a final stage of euhedral quartz. Following the quartz
precipitation, there is minor fracturing and annealing of the quartz and precipitation of
calcite, fluorite and pyrite. Minor chalcopyrite is associated with the calcite. Zeolites infill
late-stage voids.

103
Chapter Four: Geology of Cerro Corona

Figure 4.40: Summary of paragenesis of sample C12.JL.46. A) Hand sample shows apparently simple quartz-
pyrite vein with an associated sericite alteration selvage. Red box illustrates location of inset E). B) Back
scattered electron image of vein illustrating a sequence of quartz - calcite ± chalcopyrite ± pyrite ± fluorite; C)
Montage of SEM-CL image of vein illustrating multiple generations of quartz growth and re-absorption; D)
interpretation of images B and C. E) Example of SEM-CL image of quartz illustrating multiple stages of
crustiform quartz growth.

Broadly speaking, the vein sequence developed here is similar to other porphyry copper
deposits (Gustafson and Hunt 1975, Sillitoe 2010).

104
Chapter Four: Geology of Cerro Corona

4.2.6 Geological modelling of Cerro Corona


To provide a framework for mass transfer analysis, geochemical and fluid flow
interpretations, 3D models of the Cerro Corona deposit were created in Leapfrog Geo 3D,
an implicit geological modelling software package. This software utilises radial basis
functions which rapidly interpolate geochemical data between drill cores without the
requirement for explicit digitisation of individual sections (i.e. wire framing). This allows
for the rapid development of 3D models based on geological observations and geochemical
data (Cowan et al. 2002). Observed crosscutting relationships outlined herein and
conventional porphyry deposit dimensions (Berger et al. 2008) are utilised to generate the
geometry of the porphyry intrusion. Fault displacements are excluded from the final
models because the crosscutting relationships between faults are inconsistent.

The Cretaceous limestones dip approximately 30°SW near the Cerro Corona deposit
(Uzategui and Jacay 2010). These sedimentary rocks are modelled in order of stratigraphic
sequence with regionally approximate thicknesses (Fig. 4.1).

The QD1 intrudes vertically into the Cretaceous limestone. The cross-sectional geometry of
QD 1 is approximately circular at surface, thus a prolate ellipsoid ratio of 5:2:2 (maximum:
intermediate: minimum) was used to model this intrusive phase. The calcium content of
assay data was utilised to define the extent of the intrusion, where calcium values of >4
wt.% were confidently interpreted as samples of the Cretaceous limestone (Fig. 4.41).

Figure 4.41: Calcium content of rocks at Cerro Corona, based on assay data. Note the drastic change of slope from
~4 wt.%. Values >4 wt.% CaO are interpreted to be within Cretaceous limestone.

105
Chapter Four: Geology of Cerro Corona

The QD2 biotite-quartz diorite porphyries are modelled as intruding into the QD1
intrusion, as observed from cross cutting relationships (Fig. 4.6). The western biotite-
quartz diorite (“Barren Core West”) intrusion is intersected at depth by CCD-146 and CCD-
101. No drill core intersecting the “Barren Core East” was logged by the author; however,
the surface expression of this intrusion is well established and post-date ore
mineralisation. These intrusions are modelled as vertical, bodies with roughly circular
cross-section, with the same prolate spheroid ratios as QD1.

All the available assay data from Gold Fields La Cima was utilised to develop grade shells of
Mo, Cu and Au (Fig. 4.42, Fig. 4.43 and Fig. 4.44). The parameters used to model these shells
are listed in Table 4.3.

The “adaptive resolution” is defined as the distance between vertices of an interpolated


triangular mesh, with large number of closely spaced data points the distance between the
vertices will decrease to optimise the best data fit (i.e. become a finer mesh). The ellipsoid
ratio and trend are measures of the trend of the calculated interpolation.

Table 4.3: Parameters for grade shell construction.

Ellipsoid ratio (max:


Grade Number of assay
Adaptive resolution int: min) and trend
shell points
(dip, azimuth, pitch)
Molybdenum 17 030 100 7:2:2; 90, 90, 90
Copper 44 157 10 5:2:2; 90, 90, 90
Gold 43 932 100 5:2:2; 90, 90, 90

Seventeen thousand and thirty assay points were utilised to generate a molybdenum ore
shell. The distribution of this ore shell correlates well with the QD1 intrusion (Fig. 4.42). It
is thus thought that the principal stage of Mo mineralisation was introduced during the
emplacement of the QD1 intrusion.

106
Chapter Four: Geology of Cerro Corona

Figure 4.42: Plan showing molybdenum grade shells across the Corona Diorite at 3810 m a.s.l.

Utilising the copper assay values of 44 157 samples across the deposit, a robust copper
shell was created (Fig. 4.43). The distribution of Cu shows a clear dilution of grade by the
later biotite- quartz diorite intrusions, implying they postdate the main copper stage. This
is supported by the observation of crosscutting relationships.

107
Chapter Four: Geology of Cerro Corona

Figure 4.43: Plan showing copper grade shells across the Corona Diorite at 3810 m a.s.l. The areas of high grade
are concentrically arranged around the “Barren Cores”: BCE - Barren Core East; BCW - Barren Core West.

A total of 43 932 assay points was used to create the gold shell (Fig. 4.44) in which most of
assayed values are less than 1 ppm Au. This gold shell appears to be internal to the copper
shell. Both copper and gold shells are crosscut by the later intrusions. Xenoliths of the early
biotite quartz-diorite within the late, western biotite-quartz diorite correlate well with
pockets of higher gold grade.

108
Chapter Four: Geology of Cerro Corona

Figure 4.44: Map illustrating distribution of gold grades across the Cerro Corona deposit at 3810 m a.s.l.

109
Chapter Four: Geology of Cerro Corona

4.3 Summary
The Hualgayoc district lacks published whole rock geochemistry, restricting the
understanding of the Cerro Corona in the context of the local evolution of the igneous
intrusions. Comparisons of the Cerro Corona to their temporal equivalents in the nearby
Yanacocha district (Chapter 3, Fig. 3.6) suggest that similar processes may have been
involved in the development of these magmas.

The current consensus is that the magmas in the Yanacocha district are a product of cooling,
fractional crystallisation, periodic recharge of deeply derived hydrous basaltic or andesitic
melts, and mixing with silicic melts derived by crustal melting (Davies 2002, Longo 2005,
Chiaradia et al. 2009, Longo et al. 2010).

The amphibole geochemistry from the Cerro Corona intrusive complex, record a general
increase in temperature, depth (Fig. 4.47) and hydration (Fig. 4.48) of the source melts with
time. This agrees with a model of recharge of hydrous basaltic melts proposed for the
nearby Yanacocha district.

The increase in hydrous basaltic melts into the upper crust is coincidental with the
subducted buoyant “Inca Plateau” beneath the Cajamarca district (Chapter 3, Rosenbaum
et al. 2005).

Locally, the metalliferous porphyry deposits appear to be emplaced along the hinge of an
NW-SE trending anticline, with replacement Pb-Zn-Ag-Cu stratabound along more
permeable and reactive Farrat and Inca Formations, perpendicular to the trend of porphyry
emplacement.

Interestingly, CaO concentrations of the intrusive rocks at Cerro Corona are slightly lower
than their Miocene equivalents elsewhere, despite the Cerro Corona magmas being
intruded into a limestone sequence. This supports the observation that there was very little
assimilation of the host limestone into the source melt and porphyry intrusions.

The observation of intrusive contacts, truncation of quartz veins associated with early
potassic alteration and truncation of grade shells confirm the multi-stage emplacement
history of the Cerro Corona system. However, the very similar mineralogy and
geochemistry of the quartz diorites suggests that they are derived from the same magmatic
source, begging the question as to why the early phase is the most intensely mineralised.
The so-called “Barren Cores” are, in fact, separate, later intrusive events so that use of the

110
Chapter Four: Geology of Cerro Corona

term “core” is a misnomer. The relative timing of the two post-ore intrusions is still
uncertain.

The calculated oxygen fugacity and temperatures from amphibole microprobe analyses
(Ridolfi et al. 2010) follow the sulphate-sulphide buffer (Fig. 4.45). This is consistent with
the observation of anhydrite saturation and titanium-oxide calculations of the magmas at
Yanacocha (Chambefort et al. 2013), which had oxidation states ranging from of NNO +1 to
+2.5 (FMQ range of +1.5 to+3.1) at temperatures between 700-820°C. This indicates the
magmas in the district were highly oxidised, a condition thought to be important for the
generation of Cu-Au (and Mo) porphyry ore deposits by limiting sulphide saturation of the
parental melts. Comparison of the two populations of amphiboles observed suggests that
the magnesio-hornblende crystallised within early hornblende quartz diorite porphyries
from magmas that were more oxidised, lower temperature and less hydrous, than pargasite
from the late andesite dykes. The presence of feldspar inclusions within magnesio-
hornblende suggests feldspar had crystallised from the more oxidised, less hydrous and
cooler parental melts forming the hornblende quartz diorites. Pargasite does not show any
feldspar inclusions, suggesting feldspar may not have crystallised prior to amphibole
within the andesite dykes.

Figure 4.45: log fO2 versus temperature diagram with calculated amphibole temperatures and oxygen fugacities
(Ridolfi et al. 2010) from hornblende-quartz diorite porphyries (circles) and post-mineralisation andesite
(diamonds). MH = magnetite-hematite buffer, FMQ = fayalite-magnetite-quartz buffer (Ming Chou 1978).

111
Chapter Four: Geology of Cerro Corona

Magnesio-hornblendes formed at lower temperature and relatively higher oxygen fugacity than pargasitic
amphibole from late-stage andesitic dykes.

Interestingly, the plagioclase microprobe analyses show that nearly all analysed
plagioclase crystals do not display an ‘excess’ of aluminium (Fig. 4.46) seen in other water-
rich “fertile” porphyry deposits (Williamson et al. 2016). The mechanism by which Al
excess is thought to occur within feldspars was inferred to be due to high water-melt
content of the melts where AlAlSiO8 and Si4O8 co-substitute into feldspars as a function of
increasing partial pressure of water (PH2O). This suggests that the melts from which
plagioclase crystallised may not have high PH2O, contradicting the amphibole hygrometer
calculations (Ridolfi et al. 2010). This brings into question the validity of the use of Al excess
in feldspars as a tool for porphyry copper deposit exploration, as this signature is not seen
here.

Figure 4.46: Al/(Ca+Na+K) vs. anorthite percentage for feldspar crystals in biotite-quartz diorite porphyry. Note
that phenocrysts appear to be more anorthitic than feldspar inclusions in amphibole and biotite. Line joins ideal
albite and anorthite end-members.

112
Chapter Four: Geology of Cerro Corona

Figure 4.47: Pressure versus temperature of syn-mineralisation hornblende quartz diorite (circles) and post-
mineralisation andesite dykes (diamonds) as calculated from amphibole thermobarometry (Ridolfi et al. 2010).

113
Chapter Four: Geology of Cerro Corona

Figure 4.48: H2O of melt versus temperature as determined from amphibole hygrometry and thermometry of
amphiboles (Ridolfi et al. 2010) from syn-mineralisation hornblende-quartz diorites (circles) and post-
mineralisation andesite dykes (diamonds).

The hydrothermal alteration observed in this study is in general agreement with historic
understanding of the Cerro Corona deposit. Early potassic-propylitic alteration is
coincidental with main copper and gold mineralisation. This alteration and mineralisation
is crosscut by later QD2 biotite quartz diorites. The planar habit of veins associated with
sericitic alteration suggests that this alteration is structurally controlled, thus a better
understanding of the timing of faults is important in modelling the distribution of this
alteration. Permeability is highest along intrusive contacts and areas of brecciation; this
permeability appears to control the distribution of the argillic alteration.

The significant rates of uplift within the district (Chapter 3, section 3.4) may cause meteoric
fluids to telescope down permeable areas. The areas of highest permeability in porphyry
deposits are the areas of brecciation and hydraulic fracturing. This fracture network may
play a vital role in the distribution of the extent of the argillic alteration. High-grade
mineralisation associated with argillic alteration is due to this overprinting.

Hydrothermal alteration transforms major minerals into new assemblages. Does an overall
mass transfer accompany these new assemblages? This is investigated in chapter five.

114
Chapter Four: Geology of Cerro Corona

4.4 References
Berger, B. R., Ayuso, R. A., Wynn, J. C. and Seal, R. R. (2008) Preliminary model of porphyry
copper deposits. United States Geological Survey open-file report. Report number: 1321.

Borredon, R. (1982) Étude géologique et métallogénique du district minier de Hualgayoc,


Pérou septentrional, à plomb-zinc-cuivre-argent. PhD thesis. University of Paris VI.

Castillo, J. C. (2001) 1:10 000 map of regional geology of Cerro Corona. Barrick.

Chambefort, I., Dilles, J. H. and Longo, A. A. (2013) Amphibole geochemistry of the


Yanacocha Volcanics, Peru: Evidence for diverse sources of magmatic volatiles related to
gold ores. Journal of Petrology. 54 (5), 1017-1046.

Chiaradia, M., Merino, D. and Spikings, R. (2009) Rapid transition to long-lived deep crustal
magmatic maturation and the formation of giant porphyry-related mineralization
(Yanacocha, Peru). Earth and Planetary Science Letters. 288 (3), 505-515.

Cowan, E. J., Beatson, R. K., Fright, W. R., McLennan, T. J. and Mitchell, T. J. (2002) Rapid
geological modelling. Applied Structural Geology for Mineral Exploration and
Mining. Kalgoorlie, Western Australia.

Davies, R. C. I. (2002) Tectonic, magmatic and metallogenic evolution of the Cajamarca


mining district, Northern Peru. James Cook University.

Gold Fields Limited. (2014) Integrated Annual Report 2014. Available


from: https://www.goldfields.co.za/reports/annual_report_2014/minerals/reg-ame-
cerro-min-res.php . [Accessed 18th September 2016]

Gold Fields Limited. (2009) Cerro Corona Mine. Available


from: https://www.goldfields.co.za/reports/rr_2009/tech_cerro.php . [Accessed 18th
September 2016]

Gustafson, L. B. and Hunt, J. P. (1975) The porphyry copper deposit at El Salvador,


Chile. Economic Geology. 70 (5), 857-912.

Hawthorne, F. C., Oberti, R., Harlow, G. E., Maresch, W. V., Martin, R. F., Schumacher, J. C. and
Welch, M. D. (2012) Nomenclature of the amphibole supergroup. American Mineralogist. 97
(11-12), 2031-2048.

115
Chapter Four: Geology of Cerro Corona

James, J. (1998) Geology, alteration, and mineralization of the Cerro Corona porphyry
copper-gold deposit, Cajamarca Province, Peru. Master’s thesis. University of British
Columbia.

Kuno, H. (1968) Differentiation of basalt magmas. In: Hess, H. H. and Poldervaart, A.


(eds.). Basalts: The Poldervaart Treatise on Rocks of Basaltic Composition. , Interscience New
York, NY. 623-688.

Le Bas, M. J. and Streckeisen, A. L. (1991) The IUGS systematics of igneous rocks. Journal of
the Geological Society. 148 (5), 825-833.

Le Maitre, Roger Walter Bateman, Dudek, P., Keller, A., Lameyre, J., Le Bas, J., Sabine, M.,
Schmid, P., Sorensen, R., Streckeisen, H. and Woolley, A. (1989) A classification of igneous
rocks and glossary of terms: Recommendations of the International Union of Geological
Sciences, Subcommission on the Systematics of Igneous Rocks. Blackwell Scientific
Publications, Oxford, International Union of Geological Sciences.

Longo, A. A. (2005) Evolution of volcanism and hydrothermal activity in the Yanacocha


mining district, northern Peru. PhD thesis. Oregon State University.

Longo, A. A., Dilles, J. H., Grunder, A. L. and Duncan, R. (2010) Evolution of Calc-Alkaline
Volcanism and Associated Hydrothermal Gold Deposits at Yanacocha, Peru. Economic
Geology. 105 (7), 1191-1241.

Macfarlane, A. W. and Petersen, U. (1990) Pb isotopes of the Hualgayoc area, northern Peru;
implications for metal provenance and genesis of a Cordilleran polymetallic mining
district. Economic Geology. 85 (7), 1303-1327.

Macfarlane, A. W., Marcet, P., LeHuray, A. P., Petersen, U. (1990) Lead isotope provinces of
the Central Andes inferred from ores and crustal rocks. Economic Geology. 85 (8), 1857-
1880.

Macfarlane, A. W., Prol-Ledesma, R. and Conrad, M. E. (1994) Isotope and Fluid Inclusion
Studies of Geological and Hydrothermal Processes, Northern Peru. International Geology
Review. 36 (7), 645-677.

Mégard, F. (1984) The Andean orogenic period and its major structures in central and
northern Peru. Journal of the Geological Society. 141 (5), 893-900.

116
Chapter Four: Geology of Cerro Corona

Ming Chou, I. (1978) Calibration of oxygen bufrers at elevated P and T using the hydrogen
fugacity sensor. American Mineralogist. 63 690-703.

Noble, D. C., McKee, E. H., Mourier, T. and Mégard, F. (1990) Cenozoic stratigraphy,
magmatic activity, compressive deformation, and uplift in northern Peru. Geological Society
of America Bulletin. 102 (8), 1105-1113.

Pardo-Casas, F. and Molnar, P. (1987) Relative motion of the Nazca (Farallon) and South
American plates since Late Cretaceous time. Tectonics. 6 (3), 233-248.

Ridolfi, F., Renzulli, A., and Puerini, M. (2010). Stability and chemical equilibrium of
amphibole in calc-alkaline magmas: an overview, new thermobarometric formulations and
application to subduction-related volcanoes. Contributions to Mineralogy and Petrology,
160(1), 45-66.

Rosenbaum, G., Giles, D., Saxon, M., Betts, P. G., Weinberg, R. F. and Duboz, C. (2005)
Subduction of the Nazca Ridge and the Inca Plateau: Insights into the formation of ore
deposits in Peru. Earth and Planetary Science Letters. 239 (1), 18-32.

Sillitoe, R. (1997) Comments on the geological model for the Cerro Corona porphyry
copper-gold prospect and exploration potential elsewhere in the Hualgayoc district,
Northern Peru. RGC Exploration Pty Limited

Sillitoe, R. H. (2010) Porphyry Copper Systems. Economic Geology. 105 (1), 3-41.

Soule, E. B. (2014) The Hualgayoc Silver Mine. In: Anonymous The Bishop's Utopia:
Envisioning Improvement in Colonial Peru. University of Pennsylvania Press. 114-145.

Speer, J. A. (1984) Micas in igneous rocks. Reviews in Mineralogy and Geochemistry. 13 (1),
299-356.

Uzategui, A. O. and Ayala, E. (2012) Alteration and structural map: Bench 3810. Internal
report, Gold Fields La Cima S.A.A.

Uzategui, A. O. and Jacay, J. (2010) Plano Geologico Local. Gold Fields La Cima S.A.A.

Williamson, B., Herrington, R. and Morris, A. (2016) Porphyry copper enrichment linked to
excess aluminium in plagioclase. Nature Geoscience. 9, 237–241.

117
Chapter Four: Geology of Cerro Corona

Wilson, J. J. (1984) Geología de los cuadrángulos de Jayanca (13-d), Incahuasi (13-e),


Cutervo (13-f), Chiclayo (14-d), Chongoyape (14-e), Chota (14-f), Celendín (14-g),
Pacasmayo (15-d), Chepén (15-e). Lima, Peru, Instituto Geológico Minero y Metalúrgico.

Winchester, J. A. and Floyd, P. A. (1977) Geochemical discrimination of different magma


series and their differentiation products using immobile elements. Chemical Geology. 203,
25-343.

118
Chapter Five: Alteration geochemistry

5 Chapter Five: Alteration geochemistry

5.1 Introduction
As hydrothermal fluids flow through and alter a rock mass, elements within the rock are
gained and lost in a metasomatic process. These modifications are reflected by the changing
mineralogy. The aim of this chapter is to establish the overall bulk mass and net element
changes associated with each alteration stage at Cerro Corona and link these to the
observed mineral transformations. These changes are then modelled in 3D to provide
insights into the nature and controls of element fluxes during porphyry-related magmatic-
hydrothermal fluid flow. The Cerro Corona deposit provides a unique opportunity to
characterise this bulk mass transfer as pre- and syn-mineralisation biotite quartz diorite
porphyry intrusions are mineralogically and geochemically very similar, despite their
temporal differences (Chapter 4). This is the first time a mass-transfer 3D model has been
produced for a porphyry deposit.

5.2 Methods

5.2.1 Geochemistry
A total of 30 samples representative of both lithologies and alteration were sent to SRK in
Lima for detailed geochemical analysis. Ten methods of analysis were utilised to provide a
full geochemical suite (Appendix 2).

5.2.2 Sample selection for geochemical mass transfer analyses


Samples of biotite-quartz diorite (both QD1 and QD2) were used for mass balance
calculations. The post-mineralisation hornblende diorite and andesite dykes do not have
extensive alteration and are thus excluded from the mass balance calculations. The Las
Gordas rhyolite intrusion is not temporally or spatially linked to the main Cerro Corona
biotite-quartz diorites and is therefore also not considered.

5.2.2.1 Unaltered samples

Two samples of the western, post-mineralisation, biotite-quartz diorite (QD2/western


“barren core”) and two samples of the least-altered, early biotite-quartz diorite porphyry
were selected as representative of the least-altered protolith (Fig. 5.1 and Fig. 5.2). These
lithologies are mineralogically and chemically very similar despite their different
119
Chapter Five: Alteration geochemistry

emplacement ages (Chapter 4, section 4.2). The geochemistry from these four samples was
averaged to incorporate lithological heterogeneity.

Figure 5.1: Hand samples of least-altered biotite-quartz diorite used for mass transfer calculations.

120
Chapter Five: Alteration geochemistry

Figure 5.2: Map of Cerro Corona showing the location of the least-altered samples (black dots). Orange lines are
the inferred location of major faults which continue through host limestone. Western and eastern meshes
delineate the post-mineralisation, biotite-quartz diorite porphyries.

5.2.2.2 Potassic samples

Four samples of potassically-altered biotite diorite were selected for mass transfer
analyses (samples C11.JL.179B, C11.JL.178, C11.JL.183 and C11.JL.194). These samples are
characterised by a potassic alteration assemblage (Chapter 4, Section 4.2.4) in which alkali
feldspar has replaced plagioclase, predominantly in the groundmass along with abundant
quartz-magnetite veinlets (Fig. 5.3). Sample C11.JL.183 is excluded from the average of the
very strong potassic alteration as this sample shows a weak transition to propylitic
alteration, as identified by weak chloritisation of biotite.

Figure 5.3: Potassically-altered samples used for mass transfer analysis with abundant quartz-magnetite
“stringer” veins and stockwork.

5.2.2.3 Propylitic samples

Four samples displaying a propylitic alteration assemblage (Chapter 4, section 4.2.4) of


chlorite ± magnetite ± calcite was used for mass balance calculations (samples C11.JL.35,
C11.JL.36, C11.JL.92 and C11.JL.128; Fig. 5.4). This alteration stage commonly overprints
pre-existing potassic alteration and results in the replacement of biotite and hornblende by
chlorite. The intensity of this alteration is reflected in the degree of preservation of primary
igneous textures. Very strong alteration has obliterated primary textures, whereas strong
alteration has left a recognisable porphyritic texture. The very strongly-altered samples in
which the precursor was potassically altered are recognised by the presence of quartz-

121
Chapter Five: Alteration geochemistry

magnetite veining (C11.JL.35, C11.JL.36 and C11.JL.128). These samples were grouped
together as a representation of very strong propylitic alteration of an average potassically-
altered protolith.

Figure 5.4: Propylitically-altered samples used for mass transfer analysis.

5.2.2.4 Sericitic samples

The sericitic alteration assemblage is defined by sericite, pyrite and quartz which have
replaced feldspars and biotite (Chapter 4, section 4.2.4). The presence of sericite was
confirmed by infrared spectroscopy (Fig. 5.5). This alteration assemblage is associated with
quartz-pyrite ± anhydrite/calcite veins. Sericitic alteration has overprinted unaltered and
potassically-altered biotite-quartz diorite. Consequently, two protolith types were
evaluated in terms of chemical mass transfer: the unaltered diorite porphyry (samples
C11.JL.24, C11.JL.25, C11.JL.138 and C11.JL.190), and an average potassically-altered
biotite-quartz diorite porphyry. Models for sericitic alteration of average potassically-
altered diorite were subdivided into moderate (sample C11.JL.193, C11.JL.199 and
C11.JL.200) and intense (C11.JL.201 and C11.JL.205) alteration styles (Fig. 5.5).

122
Chapter Five: Alteration geochemistry

Figure 5.5: Hand samples and corresponding infrared spectra showing sericitic alteration. The deep, symmetrical
absorption features at 1400, 2200 nm along with small absorption features at ~2350 and 2450 nm are
diagnostic of muscovite. A) unaltered biotite-quartz diorite protolith; B) potassically-altered biotite-quartz
diorite with moderate intensity of sericite alteration; and C) potassically-altered biotite-quartz diorite with very
strong sericitic alteration overprint.

5.2.2.5 Argillic samples

These argillic alteration is predominated by clay minerals and overprints all other
alteration types (Chapter 4, section 4.2.4). The intensity of this alteration varies
significantly. In this study, the most intensely- altered samples were selected in which both

123
Chapter Five: Alteration geochemistry

phenocrysts and matrix have been altered to clay minerals, so that the original porphyritic
texture has been obliterated and pre-existing alteration types are obscured (Fig. 5.6). The
argillic alteration was subdivided into dickite-dominant (samples C11.JL.95, C11.JL.108
and C11.JL.110), kaolinite ± smectite (samples C11.JL.78, C11.JL.96 and C11.JL.121), and
illite ± smectite (samples C11.JL.74, C11.JL.75, C11.JL.76 and C11.JL.115), based on the
dominant clay mineral type as identified by short wave infrared spectroscopy (Fig. 5.6).

Figure 5.6: Argillically-altered samples with corresponding infrared spectra.

124
Chapter Five: Alteration geochemistry

5.2.3 The isocon method


The use of conservative elements can be used to define a line of ISO-CONcentration
(“ISOCON”) when plotting an altered versus an unaltered rock. The isocon method is a
geochemical tool whereby the bulk mass change and net element gains and losses in a rock
that has undergone alteration are evaluated (Gresens 1967, Grant 1986, 2005). The basis
of the method is the recognition of elements that were neither gained nor lost (i.e.
conserved) during an alteration event.

The concentration (C) of an element (i) in an altered rock (A) is the result of altering the
concentration of the element from the protolith (O) with changing mass (M) (Grant 1986),
defined by:

𝐶 ∗ 𝑀 = 𝐶 ∗ 𝑀 + ∆𝐶

If it is assumed that the element i is conserved (i.e. ∆𝐶 =0), the equation can be rearranged
in the form:

𝐶 = 𝑀 /𝑀 (𝐶 + 0)

which defines a line (ISOCON) of slope MO/MA on a plot of CiA vs. CiO.

An ISOCON slope of 1 indicates no bulk mass gain or loss in the system, a slope <1 indicates
a bulk mass gain (i.e. dilution of the conservative elements) and a slope >1 indicates a bulk
mass loss (i.e. apparent enrichment of conservative elements).

The bulk mass change is given by:

1
𝑏𝑢𝑙𝑘 𝑚𝑎𝑠𝑠 𝑐ℎ𝑎𝑛𝑔𝑒(%) = − 1 ∗ 100
𝑚

where m = MO/MA

Once the gradient of the ISOCON has been determined, a net change in element
concentration can be calculated as the difference between the measured concentration of
the element and the product of the gradient of the ISOCON and the unaltered concentration
of the element (i.e. the offset of the measured concentration relative to the ISOCON):

∆𝐶 = 𝐶 − 𝑚 ∗ 𝐶

The identification of the least-altered protolith, and the precursor alteration type for a
given alteration step, as well as the conservative elements that define the ISOCON, are

125
Chapter Five: Alteration geochemistry

crucial to the correct interpretation of bulk mass changes and net gains and losses of
elements. The most suitable immobile elements were identified by determining correlation
coefficients of all elements with high field strength elements that are typically regarded as
conservative (i.e. relatively immobile) and which were well above the analytical detection
limit. Bivariate plots of such elements illustrate which elements show strong covariation
with a trend towards the origin; this behaviour confirms their conservative behaviour
during alteration (Fig. 5.7).

Figure 5.7: Plots of TiO2 and Zr versus Al2O3 illustrating conservative element behaviour and mass addition
results in overall dilution of these elements.

Uncertainties in the ISOCON method have been the subject of some debate (Grant 1986,
2005; Baumgartner and Olsen 1995; Mukherjee and Gupta 2008). This debate is centred
on the graphical display of data in which the elements are arbitrary scaled, thereby shifting
each component toward or away from the origin such that they lie in the arbitrary
concentration ranges of 0-30 weight percent and 0 to 300 ppm (Grant 1986). This arbitrary
scaling demonstrably causes bias when scaling is imposed on the conservative elements
which define the ISOCON (Fig. 5.8, Mukherjee and Gupta 2008) because of the increase in
leverage on the fit of data far from the origin. The shift in the ISOCON can be significant so,
in this study, arbitrary scaling is not utilised.

Here, the ISOCON is calculated from unscaled values but this means the results cannot be
easily displayed on linear graphs so that log scaled plots are shown (e.g. Fig. 5.9). The
ISOCON has been derived based on a selected set of immobile elements, utilising the least
squares method (Fig. 5.9; Baumgartner and Olsen 1995). Bar charts of the net element gains
and losses relative to the ISOCON are presented for ease of evaluation.

126
Chapter Five: Alteration geochemistry

Figure 5.8: Example of the effect of the arbitrary scaling method on definition of the ISOCON. Black line
represents the unscaled ISOCON. The red and orange lines are the resultant ISOCON from arbitrary scaling of
conservative components. Notice the drastically different slope that results from the arbitrary scaling.

Figure 5.9: Example of an ISOCON for a potassically-altered biotite-quartz diorite sample (C11.JL.178) showing
net gains (green arrows) and losses (red arrow) of major components using the least squares approach
127
Chapter Five: Alteration geochemistry

(Baumgartner and Olsen 1995). Conservative components used to define the ISOCON (red line) are TiO2, Al2O3
and Zr. The error bars indicate the analytical uncertainty for this sample and the standard deviation of the
average of the least-altered biotite-quartz diorite samples.

5.2.4 Error estimates


The uncertainties and errors involved in the ISOCON method are commonly ignored, or
overlooked (Grant 1986, 2005; Ulrich and Heinrich 2002, Idrus et al. 2009). Mass transfer
may be erroneously inferred if these uncertainties are not considered. Determination of
these uncertainties may provide a more confident determination of mass transfer. The
uncertainties considered here are:

a) Intensity of alteration. Alteration intensity was determined on a hand sample scale


(similar to Gifkins et al. 2005). Samples of similar alteration intensity were selected
to provide the best snapshot of a given alteration sequence.
b) Analytical error. This is the uncertainty in the of the geochemical analysis for the
concentration of each element within an analysed sample.
c) Lithological homogeneity. This is the geochemical variability of a group of samples.
Where multiple samples were averaged for a given lithology, the standard deviation
of a component is utilised as the uncertainty as this variance is larger (and more
representative) than the analytical uncertainty for this component in a single
sample.
d) Correlation between conservative elements. This is the error in the slope of the
ISOCON, propagated through the origin.

The combined uncertainties in the change in concentration of the element i were calculated
by incorporating the errors associated with the calculation of ∆𝐶 using the propagation of
error methods outlined in Ku (1966), where the difference between uncertainties combine
in quadrature and, for the multiplicative uncertainties, the relative uncertainties were
added in quadrature. Thus, the combined uncertainty in the change in concentration of
element i (δ∆𝐶 ) is given by:

𝛿𝑚 𝛿𝐶
δ∆𝐶 = (𝛿𝐶 ) + +
𝑚 𝐶

Where 𝛿𝐶 is the analytical uncertainty or the standard deviation from the average
(depending on whether a single sample or a group of samples were used for mass transfer
analysis) of the altered sample(s); 𝛿𝑚 is the standard error on the gradient of the ISOCON

128
Chapter Five: Alteration geochemistry

(m); 𝐶 is the concentration of element i in the protolith and 𝛿𝐶 is the standard deviation
of this element in several samples of the protolith.

This equation can be simplified to:

𝛿𝑚 𝛿𝐶
δ∆𝐶 = (𝛿𝐶 ) + +
𝑚 𝐶

5.3 Results
In this study, the bulk mass changes and net element transfers occurring within each
alteration stage were evaluated by utilising a broad range of samples which show a
spectrum of overprinting alteration types and intensities (Fig. 5.10). Samples that display
a precursor alteration stage were logically compared to the average of the preceding
alteration type. Thus, the geochemistry of potassic alteration is compared to the average
least-altered biotite-quartz diorite (Fig. 5.10-A). The geochemistry of propylitic alteration
is compared to the average potassic alteration assemblage (Fig. 5.10-B), and those samples
that show no precursor alteration were compared to the average least-altered biotite-
quartz diorite (Fig. 5.10-C). The geochemistry of the sericitic alteration stage was compared
to the average least-altered biotite-quartz diorite (Fig. 5.10-D) and to potassically-altered
biotite-quartz diorite (Fig. 5.10-E). The intense argillic alteration obliterates any pre-
existing textures, thus making the protolith challenging to determine. For simplicity, this
alteration stage was compared to the average least-altered biotite-quartz diorite (Fig. 5.10-
F).

129
Chapter Five: Alteration geochemistry

Figure 5.10: Schematic diagram of overprinting alteration relationships used in the mass transfer analysis
reflecting the spectrum of overprinting alteration. Note that each alteration type has variable intensities. A)
Potassic alteration of least-altered biotite-quartz diorite; B) propylitic alteration of previously potassically-
altered biotite-quartz diorite; C) propylitic alteration of a least-altered biotite-quartz diorite; D) sericitic
alteration of least-altered biotite-quartz diorite; E) sericitic alteration of potassically-altered biotite-quartz
diorite, with variable sericitic alteration intensity; and F) argillic alteration of an inferred least- altered biotite-
quartz diorite.

5.3.1 Potassic alteration


The potassically-altered samples were individually compared with the average least-
altered biotite-quartz diorite to evaluate the bulk mass changes and net gains and losses of
all components during potassic alteration. Samples C11.JL.179B, C11.JL.178 and C11.JL.194
were averaged in order to incorporate heterogeneity during the most intense potassic
alteration of the biotite-quartz diorite.

The slope of the ISOCON for C11.JL.179B is 0.66±0.02, for C11.JL.178 is 0.60±0.02, for
C11.JL.183 is 0.68±0.03 and for C11.JL.194 is 0.77±0.04. These slopes indicate that there
was a dilution in the concentration of the conservative elements by bulk mass gains of
52±4%, 66±5%, 47±5% and 29±6% respectively. The slope of the ISOCON for average
potassic alteration is 0.68±0.02, equating to a bulk mass gain of 46±3%.

5.3.1.1 The major oxides

130
Chapter Five: Alteration geochemistry

Additions of the major element oxides most likely account for most of the dilution of the
conservative elements as they make up the largest proportions of the samples. During this
stage of alteration, SiO2, K2O and Fe2O3 were added, on average, by 24%, 3% and 2%,
respectively (Fig. 5.11). SiO2 addition was primarily added through the emplacement of
quartz veins and account for most of the bulk mass gain. Na2O was within error of no net
change except for samples C11.JL.179B and C11.JL.183, in which it was added by 0.7±0.5%
and 0.9±0.5%, respectively (Fig. 5.11). On average, sulphur was within error of no net
change (1±2%), except for sample C11.JL.194 in which it was added by 3.3±0.4% (Fig. 5.11).

CaO was lost on average by 0.9 ±0.4%, with most net CaO loss in sample C11.JL.194 with a
loss of 1.3±0.3%. In average potassically-altered samples, MgO was within error of no net
change (0.4±0.5%), except for sample C11.JL.194, which had lost 1.1±0.1 % (Fig. 5.11).
Cr2O3 and P2O5 were within error of no net change. The Loss On Ignition (LOI), assumed in
this case to be primarily controlled by water, was within error of no change except for
samples C11.JL.183 and C11.JL.194 in which it increased by ~1.3±0.5% and 0.9±0.5%
respectively (Fig. 5.11).

Figure 5.11: Bar chart of net changes of major components during potassic alteration. Uncertainties have been
calculated through propagation of errors (see Section 5.2.4).

5.3.1.2 High Field Strength Elements

The High Field Strength Elements (HFSE) were broadly within error of no net change as
might be anticipated due to their generally immobile behaviour (Fig. 5.12). In several cases,
the calculated gains and losses appear to show mobility beyond the calculated uncertainty;

131
Chapter Five: Alteration geochemistry

however, the absolute concentrations of these elements are close to analytical detection
limits and thus caution must be exercised in making inferences from these results.

Thorium appeared to be gained in samples C11.JL.179B and possibly C11.JL.183 by 2.2±1.6


and 1.6±1.4 ppm, respectively, and lost in samples C11.JL.178 and C11.JL.194 by 1.0±0.4
and 1.4±0.4 ppm, respectively. Tantalum appeared to be added in sample C11.JL.194 by
~2.1±1.1 ppm (Fig. 5.12).

Figure 5.12: Bar chart of net changes in HFSE based on comparison of potassically- altered biotite-quartz diorite
with least-altered biotite-quartz diorite.

5.3.1.1 Large Ion Lithophile Elements

Barium had the strongest addition, with an average gain of ~238±54 ppm (Fig. 5.13).
Addition of Ba occurred in all samples: 79±44 ppm (C11.JL.179B), 262±48 ppm
(C11.JL.178), 968±108 ppm (C11.JL.183) and 384±65 ppm (C11.JL.194). Rubidium
behaved similarly to potassium and was enriched by an average of 37±6 ppm. All samples
showed Rb addition: 30±2 ppm (C11.JL.179B), 44±1 ppm (C11.JL.178), 39±3 ppm
(C11.JL.183) and 37±1 ppm (C11.JL.194), (Fig. 5.13). Strontium was generally lost, except
for sample C11.JL.183. The average potassically-altered biotite-quartz diorite displayed Sr
loss of 181±177 ppm, with samples C11.JL.179B, C11.JL.178 and C11.JL.194 having lost
145±35, 134±32 and 257±29 ppm, respectively. Sample C11.JL.183 had a 172±57 ppm gain
in Sr.

132
Chapter Five: Alteration geochemistry

Figure 5.13: Net changes of LILE are based on comparison of potassically-altered biotite-quartz diorite with
least-altered biotite-quartz diorite.

5.3.1.2 Rare Earth Elements

The LREE were broadly lost with Ce being the most depleted, followed by La, Nd and Pr
(Fig. 5.14). Samarium, Eu and Gd were generally within error of no net change. The samples
C11.JL.179B and C11.JL.183 were within error of no net change of the REE (Figs. 5.14 and
Fig. 5.15). The average potassically-altered biotite-quartz diorite lost ~6±5, 11±10, 1.3±0.7
and 5±4 ppm of La, Ce, Pr and Nd respectively (Fig. 5.14). Sample C11.JL.178 lost La, Ce, Pr
and Nd by 6±2, 11±8, 1.3±0.3 and 5±2 ppm, respectively. Likewise, sample C11.JL.194 lost
La, Ce, Pr and Nd by 8±2, 17±7, 1.9±0.3 and 8±2 ppm respectively (Fig. 5.14). Sample
C11.JL.194 also shows losses of Sm and Gd by 1.1±0.4 and 1.0±0.4 ppm, respectively (Fig.
5.15).

133
Chapter Five: Alteration geochemistry

Figure 5.14: Net changes of LREE are based on comparison of potassically-altered biotite-quartz diorite with
least-altered biotite-quartz diorite.

The HREE were within error of no change, except for Dy, Er and Yb, which were lost in some
samples (Fig. 5.15). Lutetium was below detection in the altered samples, and is therefore
not shown. Dysprosium was lost by 0.54±0.47 ppm in average potassically-altered biotite-
quartz diorite and by 0.7±0.2 and 0.7±0.3 ppm in samples C11.JL.178 and C11.JL.194,
respectively. Erbium was lost by 0.4±0.3 ppm in average potassically-altered biotite-quartz
diorite and by 0.4±0.3 and 0.5±0.3 ppm in samples C11.JL.178 and C11.JL.194, respectively.
Ytterbium was lost by 0.3±0.2 ppm in average potassically-altered biotite-quartz diorite
and by 0.4±0.2 and 0.4±0.2 ppm in samples C11.JL.178 and C11.JL.194, respectively.

134
Chapter Five: Alteration geochemistry

Figure 5.15: Net changes of HREE are based on comparison of potassically-altered biotite-quartz diorite with
least-altered biotite-quartz diorite.

Scandium was added to samples C11.JL.179B, C11.JL.178 and C11.JL.183 by 3.5±0.2,


3.5±0.2 and 1.7±0.3 ppm, respectively. It was removed from sample C11.JL.194 by 4.6±0.2
ppm and was within error of no net change for the average potassically-altered biotite-
quartz diorite (Fig. 5.16). Yttrium was within error of no change for average potassically-
altered biotite-quartz diorite, sample C11.JL.179B and C11.JL.183 (Fig. 5.16). In both
C11.JL.178 and C11.JL.194, Y was removed by 4±1 ppm.

Figure 5.16: Net change of REE-affinity elements Sc and Y based on comparison of potassically-altered biotite-
quartz diorite with least-altered biotite-quartz diorite.

5.3.1.3 Metals

The most significant metal gains were in Cu and Au. Copper was added by ~3000 ppm in
average potassically-altered biotite-quartz diorite and Au by ~420 ppb (Fig. 5.17). Copper
was added by ~2134±174, 1459±130 and 5506±320 ppm in samples C11.JL.179B,
C11.JL.178 and C11.JL.194 respectively, whereas gold was added by 296±33, 217±26 and
736±61 ppb in the same samples, respectively (Fig. 5.17). Sample C11.JL.183 has lost
~330±37 ppm and 58±7 ppb of Cu and Au, respectively (Fig. 5.17), probably due to an
overprinting alteration.

135
Chapter Five: Alteration geochemistry

Figure 5.17: Net change of copper and gold based on comparison of potassically-altered biotite-quartz diorite
with least-altered biotite-quartz diorite.

Chromium and V were lost by a significant amount, by ~66 and ~29 ppm respectively (Fig.
5.18). Caesium was consistently lost in all potassically-altered samples averaging 2.3 ± 0.4
ppm (Fig. 5.18). Vanadium was removed in samples C11.JL.179B, C11.JL.178 and
C11.JL.194 by 23±5, 15±5 and 47± 4 ppm, respectively, but added in sample C11.JL.183 by
14±7 ppm (Fig. 5.18). The average potassically-altered biotite-quartz diorite lost V by
~30±10 ppm. Lead was gained in the samples C11.JL.183 and C11.JL.194 by 7±3 and 5±3
ppm, but was within error of no change in samples C11.JL.179B and C11.JL.178. The
average potassically-altered biotite-quartz diorite had a Pb gain of 4±2 ppm (Fig. 5.18).
Molybdenum was predominantly within error of no change, except for sample C11.JL.179B,
where it was added by 8±7ppm (Fig. 5.18). Zinc was strongly added in samples
C11.JL.179B, C11.JL.178 and C11.JL.183 by 18±9, 26±10 and 28±8 ppm respectively; but
lost in sample C11.JL.194 by 30±6 ppm.

136
Chapter Five: Alteration geochemistry

Figure 5.18: Net change of metals based on comparison of potassically-altered biotite-quartz diorite with least-
altered biotite-quartz diorite.

The other metals were, on average, within error of no change, or inconsistently added in
some samples, but removed in others, thus little systematic behaviour can be deduced for
these metals.

5.3.2 Propylitic alteration


The ISOCON gradients for samples C11.JL.35 (Fig. 5.19-A), C11.JL.36, C11.JL.92 and
C11.JL.128 assuming an average least-altered biotite diorite protolith was 0.580±0.002,
0.525±0.004, 0.93±0.02 and 0.579±0.002, respectively. The ISOCON of the average very
strong propylitic alteration assuming an average least-altered biotite-quartz diorite
protolith was 0.56±0.02, equating to a bulk mass gain of 80±5%. However, altered samples
C11.JL.35, C11.JL.36 and C11.JL.128 contain magnetite and/or quartz-magnetite stockwork
veins- indicative of a preceding potassic alteration stage (Chapter 4, Section 4.2.4), thus a
more realistic ISOCON plot for these samples was calculated utilizing an average
potassically-altered biotite diorite protolith. These more realistic ISOCON values were
0.85±0.02 (C11.JL.35), 0.77±0.01 (C11.JL.36), and 0.85±0.02 (C11.JL.128), equating to
smaller bulk mass gains in the propylitic alteration step of 18±3, 30±2 and 18±3%,
respectively (Fig. 5.19-B). The ISOCON gradient of sample C11.JL.92 compared with the
least-altered biotite-quartz diorite equated to a mass gain of 8±2%. The gradient of the
ISOCON for the average very strong propylitically altered sample overprinting the average
potassically-altered biotite diorite was 0.81±0.02. This ISOCON equated to a bulk mass gain
of 23±4%. Estimates of bulk mass gain are significantly lower than those for potassically-

137
Chapter Five: Alteration geochemistry

altered samples; nonetheless, they clearly demonstrate that propylitic alteration was not
an isochemical process and involved substantive mass transfer.

Figure 5.19: Example of an ISOCON for propylitically-altered sample (C11.JL.35) compared to (A) the average
least-altered biotite-quartz diorite, and (B) to the average potassically-altered biotite-quartz diorite showing net
gains and losses of major components using the least squares approach (Baumgartner and Olsen 1995). Error
bars are the relative standard deviation of the analyses for the propylitically-altered sample and the standard
deviation of the average least-altered biotite-quartz diorite and average potassically-altered biotite-quartz
diorite.

138
Chapter Five: Alteration geochemistry

5.3.2.1 The major oxides

Net gains of SiO2>Fe2O3>LOI were seen in all samples, suggesting an unequivocal addition
of these components (Fig. 5.20). Silica was gained by 13±2, 16±2, 12±2 and 27±2% for
samples C11.JL.35, C11.JL.36, C11.JL.128 and C11.JL.92, respectively. Ferric oxide was
added by 2.9±0.8, 6.9±0.9, 6.2±0.9, 3.6±0.7% to samples C11.JL.35, C11.JL.36, C11.JL.128
and C11.JL.92, respectively. The loss on ignition, assumed in this case to be primarily
controlled by water, was gained by 2.0±0.6, 2.5±0.6, 2.1±0.6 and 2.0±0.6% to samples
C11.JL.35, C11.JL.36, C11.JL.128 and C11.JL.92, respectively.

Sodium oxide was unequivocally removed from all propylitically-altered samples (Fig.
5.20) as seen in net losses of 1.9±0.2, 1.6±0.2, 2.2±0.1 and 2.1±0.2% in samples C11.JL.35,
C11.JL.36, C11.JL.128 and C11.JL.92, respectively. Calcium oxide appeared to have been
gained in samples C11.JL.35 and C11.JL.36 by 1.9±0.5 and 1.2±0.6%, but removed in
samples C11.JL.128 and C11.JL.92 by 0.8±0.3 and 0.7±0.5% (Fig. 5.20). Potassium oxide
was removed from samples C11.JL.36 and C11.JL.128 by 1.0±0.5 and 1.1±0.5%, but added
to the sample with an inferred unaltered protolith (sample C11.JL.92) by 1.3 ±0.5% and
was within error of no change for sample C11.JL.35. Sulphur was gained in C11.JL.36,
C11.JL.128 and C11.JL.92 by 3±1, 2±1 and 0.6±0.3% respectively, but was within no change
in sample C11.JL.35 (Fig. 5.20). Magnesium oxide and P2O5 were within error of no change
in all samples (Fig. 5.20).

The average very strongly propylitically-altered samples had net gains of SiO2, Fe2O3 and
LOI by 14±1, 5±2, 2.2±0.3% a loss of Na2O by 1.9±0.2% and no net change of CaO, K2O, MgO,
P2O5 and S.

Figure 5.20: Bar chart of net changes of major components during propylitic alteration compared with an
average potassically-altered biotite-quartz diorite (C11.JL.35, C11.JL.36 and C11.JL.128) or the average least-
139
Chapter Five: Alteration geochemistry

altered biotite-quartz diorite (C11.JL.92). “VS Prop” is the average of very strongly propylitically-altered samples
compared with the average potassically-altered biotite-quartz diorite. Error bars were calculated through
propagation of error (see section 4.2.4)

5.3.2.1 High Field Strength Elements

As expected, the high field strength elements were within error of no change (Fig. 5.21).
Thus, during the propylitic alteration the HFSE were conserved.

Figure 5.21 Bar chart of net changes of HFSE based on comparison of propylitically-altered biotite-quartz diorite
to potassically-altered protolith (samples C11.JL.35, C11.JL.36 and C11.JL.128) or least-altered biotite-quartz
diorite (sample C11.JL.92). “VS Prop” is the average of very strongly propylitically-altered samples compared
with the average potassically-altered biotite-quartz diorite.

5.3.2.1 Large Ion Lithophile Elements

Propylitic alteration unequivocally removed Ba and Sr by an order of magnitude more than


Rb (Fig. 5.22). Barium was removed by 178±34, 296±35, 226±30 and 95±29 ppm, whereas
Sr was removed by 95±17, 107±11, 132±12, and 308±19 ppm from samples C11.JL.35,
C11.JL.36, C11.JL.128 and C11.JL.92. Rubidium, as expected, mimicked that of K2O. It was
removed from samples C11.JL.36, C11.JL.128 by 17±1 and 11±1ppm; was added to the
sample with an unaltered protolith (sample C11.JL.92) by 7±1 ppm; and was within error
of no change in sample C11.JL.35 (Fig. 5.22).

140
Chapter Five: Alteration geochemistry

Figure 5.22: Bar chart of net changes of LILE based on comparison of propylitically-altered biotite-quartz diorite
with potassically-altered protolith (samples C11.JL.35, C11.JL.36 and C11.JL.128) or least-altered biotite-quartz
diorite (sample C11.JL.92). “VS Prop” is the average of very strongly propylitically- altered samples compared
with the average potassically-altered biotite-quartz diorite.

5.3.2.1 Rare Earth Elements

The light rare earth elements La, Ce, Pr and Nd were added to sample C11.JL.35 by 7±4,
16±5, 2±1 and 8±4ppm, respectively, with smaller gains in sample C11.JL.36 (Fig. 5.23).
Samples C11.JL.128 and C11.JL.92 lost these LREE, where La was removed by 2±1 and 4±2
ppm, Ce by 4±2 and 7±3 ppm, and Nd by 2±1 and 3±2 ppm, respectively. Samarium, Eu and
Gd were within error of no net change as was Pr in samples C11.JL.128 and C11.JL.92 (Fig.
5.23).

141
Chapter Five: Alteration geochemistry

Figure 5.23: Bar chart of net changes of LREE based on comparison of propylitically-altered biotite-quartz
diorite with potassically-altered protolith (samples C11.JL.35, C11.JL.36 and C11.JL.128) or least-altered biotite-
quartz diorite (sample C11.JL.92). “VS Prop” is the average of very strongly propylitically-altered samples
compared with the average potassically-altered biotite-quartz diorite.

The HREE were within error of no net change, indicating that they were conserved during
propylitic alteration (Fig. 5.24).

Figure 5.24: Bar chart of net changes of LREE based on comparison of propylitically altered biotite-quartz diorite
to potassic altered protolith (samples C11.JL.35, C11.JL.36 and C11.JL.128) or least-altered biotite-quartz diorite
(sample C11.JL.92). “VS Prop” is the average very strong propylitically altered samples compared with the
average potassically-altered biotite-quartz diorite.

Scandium was added in all propylitically-altered samples, whether overprinting


potassically-altered or an unaltered protolith (Fig. 5.25). It was gained by 10.2±0.4,

142
Chapter Five: Alteration geochemistry

13.8±0.4, 4.9±0.3 and 2.6±0.2 ppm in samples C11.JL.35, C11.JL.36, C11.JL.128 and
C11.JL.92, respectively. Yttrium was within error of no net change, apart from sample
C11.JL.35 in which it gained was added by 4±2 ppm (Fig. 5.25).

Figure 5.25: Bar chart of net change of REE-affinity elements Sc and Y based on comparison of propylitically-
altered biotite-quartz diorite with potassically- altered protolith (samples C11.JL.35, C11.JL.36 and C11.JL.128)
or least-altered biotite-quartz diorite (sample C11.JL.92). “VS Prop” is the average of very strongly propylitically-
altered samples compared with the average potassically-altered biotite-quartz diorite.

5.3.2.2 Metals

Copper, Au, Cr, Zn and V were all added during propylitic alteration of potassically-altered
biotite diorite (Fig. 5.26). Also gained, in decreasing order of net addition in the average of
the very strongly propylitically-altered samples were Cs, Co, Mo, Ga, As, and Ag, with Pb, Ni
and Th within error of no net change (Fig. 5.27). In samples C11.JL.35, C11.JL.36 and
C11.JL.128, Cs, Co, As and Ag were added whereas Pb, Mo, Ga and Ni were variable in gains
and/or losses (Fig. 5.27) indicating somewhat erratic behaviour.

The metals Sb, Bi, Tl, Th, Ge, Hg, Te, Sn and W were within error of no change in all samples,
indicating that these metals were conserved under the conditions of propylitic alteration
at Cerro Corona (Fig. 5.27).

143
Chapter Five: Alteration geochemistry

Figure 5.26: Bar chart of largest net changes of metals based on comparison of propylitically-altered biotite-
quartz diorite with potassically-altered protolith (samples C11.JL.35, C11.JL.36 and C11.JL.128) or least-altered
biotite-quartz diorite (sample C11.JL.92). “VS Prop” is the average of very strongly propylitically-altered samples
compared with the average potassically-altered biotite-quartz diorite. All changes are in ppm except for Au,
which is in ppb.

Figure 5.27: Bar chart of net changes of metals based on comparison of propylitically-altered biotite-quartz
diorite with potassically-altered protolith (samples C11.JL.35, C11.JL.36 and C11.JL.128) or least-altered biotite-
quartz diorite (sample C11.JL.92). “VS Prop” is the average of very strongly propylitically-altered samples
compared with the average potassically-altered biotite-quartz diorite.
144
Chapter Five: Alteration geochemistry

5.3.3 Sericitic alteration


The sericitised whole rock samples were grouped into:

a) very strongly sericitised, previously unaltered/weakly altered biotite diorite


porphyry;
b) previously potassically-altered samples with two subsets reflecting moderate and
very strong sericitisation.

The ISOCON gradients of these groups were thus calculated against their respective
protoliths. The ISOCON slopes of the sericitised, previously unaltered samples (n=4),
moderate (n=3) and very strongly (n=2) sericitised, previously potassically-altered
samples were 0.87±0.02 (Fig. 5.28), 0.61±0.01 (Fig. 5.29) and 0.26±0.01 (Fig. 5.30),
respectively. This equated to a bulk mass gain of 15±3% for the sericitisation of a
previously unaltered lithology and mass gains of 63±3%, and 288±13% with increasing
sericitic alteration intensity of the average, potassically-altered protolith.

Figure 5.28: An ISOCON of the average sericitised biotite-quartz diorite (samples C11.JL.24, C11.JL.25, C11.JL.138
and C11.JL.190) compared with the average least-altered biotite-quartz diorite using the least squares approach
(Baumgartner and Olsen 1995). Error bars are the standard deviation of average sericitised biotite-quartz
diorite and the average least-altered biotite-quartz diorite.

145
Chapter Five: Alteration geochemistry

Figure 5.29: An ISOCON of the average moderately-sericitised biotite-quartz diorite (samples C11.JL.193,
C11.JL.199 and C11.JL.200) compared with the average potassically-altered biotite-quartz diorite using the least
squares approach (Baumgartner and Olsen 1995). Error bars are the standard deviation of average very
strongly-sericitised biotite-quartz diorite and the average potassically-altered biotite-quartz diorite.

146
Chapter Five: Alteration geochemistry

Figure 5.30: An ISOCON of the average very strongly-sericitised biotite-quartz diorite (samples C11.JL.201 and
C11.JL.205) compared with the average potassically-altered biotite-quartz diorite using the least squares
approach (Baumgartner and Olsen 1995). Error bars are the standard deviation of average moderately-
sericitised biotite-quartz diorite and the average potassically-altered biotite-quartz diorite.

5.3.3.1 The major oxides

Sodium oxide was removed in all stages of sericitisation, regardless of the protolith (Fig
5.31). Average sericitisation of an unaltered biotite-quartz diorite resulted in a net loss of
2.6±0.3 %, moderate sericitisation of a potassically-altered biotite-quartz diorite resulted
in the removal of 1.6±0.1% and very strong sericitisation of a potassically-altered biotite-
quartz diorite resulted in a net loss of 0.7±0.1 % (Fig 5.31). Removal of CaO probably
occurred during sericitisation of a previously unaltered protolith (loss of ~1.1±1.0%), but
results were within error of no change when considering a potassically-altered protolith
(Fig. 5.31). K2O was added to previously unaltered biotite-quartz diorite (2±1%), but was
within error of no change in very intensely-sericitised samples with an average
potassically-altered protolith, or was possibly removed from moderately-sericitised
samples that had an average potassically-altered protolith (0.5±0.3%). These apparent
losses may have been due to various intensities of the preceding potassic alteration, which
was not taken into consideration. Magnesium oxide and P2O5 appeared to be within error
of no change in all sericitic alteration stages (Fig. 5.31).

Ferric oxide was not added to the previously unaltered protolith, but was added with
increasing intensity of sericitic alteration after previously potassically-altered rocks, with
~ 7±3% gain in the most intense sericitisation (Fig. 5.31).

Sulphur and SiO2 were gained in all sericitised samples, with the largest net increases
observed in the most intensely-sericitised samples, with gains of 5±3 and 65±5%,
respectively (Fig. 5.31). The LOI, assumed in this case to be primarily controlled by water,
increased by ~4% regardless of the state of the protolith (Fig. 5.31).

147
Chapter Five: Alteration geochemistry

Figure 5.31: Bar chart of net changes of major components during sericitic alteration compared with the average
least-altered biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an average potassically-altered
biotite-quartz diorite (“Sericite_M”) or very strong sericitisation of an average potassically-altered biotite-quartz
diorite (“Sericite_VS”). Error bars were calculated through propagation of error (section 4.2.4).

5.3.3.1 High Field Strength Elements

Of the HFSE, Nb appeared to be mobile (Fig. 5.32). It was lost in samples modelled to
overprint an unaltered protolith (2±1 ppm), within error of no change in moderately-
sericitised, previously potassically-altered samples, but added in very strong sericitic
alteration by ~0.7±0.4 ppm (Fig. 5.32). The other HFSE appeared to be within error of no
net change (Fig. 5.32).

148
Chapter Five: Alteration geochemistry

Figure 5.32: Bar chart of net changes of HFSE during sericitic alteration compared with the average least-altered
biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an average potassically-altered biotite-quartz
diorite (“Sericite_M”) or very strong sericitisation of an average potassically-altered biotite-quartz diorite
(“Sericite_VS”). Error bars were calculated through propagation of error (section 4.2.4).

5.3.3.1 Large Ion Lithophile Elements

Strontium and Ba were the most drastically affected elements within the LILE, Sr was lost
by ~443±63 ppm from sericitised samples relative to an unaltered protolith (Fig. 5.33).
Strontium was removed by ~105±23 ppm in moderately- sericitised samples with a
potassically-altered protolith (Fig. 5.33) but was within error of no net change in very
intensely-sericitised samples with a potassically-altered protolith. Barium was broadly
removed during sericitisation with the largest losses seen on previously unaltered
protolith (~278±106 ppm) and lesser removal from very intensely-sericitised, previously
potassically-altered samples (~120±3 ppm) (Fig. 5.33). Rubidium appeared to follow the
trend of K2O in that it was added during sericitic alteration of the unaltered protolith (12±6
ppm), but was within error of no change for sericitisation of a potassically-altered protolith
(Figs. 5.31 and 5.33).

149
Chapter Five: Alteration geochemistry

Figure 5.33: Bar chart of net changes of LILE during sericitic alteration compared with the average least-altered
biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an average potassically-altered biotite-quartz
diorite (“Sericite_M”) or very strong sericitisation of an average potassically-altered biotite-quartz diorite
(“Sericite_VS”). Error bars were calculated through propagation of error (section 4.2.4).

5.3.3.1 Rare Earth Elements

The REE were within error of no net change during sericitisation, regardless of protolith
composition or intensity of alteration, perhaps apart from Eu (Fig. 5.34). This suggests that
most of the REE were conserved during sericitisation. Europium may have been weakly
removed during sericitisation of an unaltered biotite-quartz diorite by 0.3±0.2 ppm.

150
Chapter Five: Alteration geochemistry

Figure 5.34: Bar charts of net changes of REE during sericitic alteration compared with the average least-altered
biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an average potassically-altered biotite-quartz
diorite (“Sericite_M”) or very strong sericitisation of an average potassically-altered biotite-quartz diorite
(“Sericite_VS”). Error bars were calculated through propagation of error (section 4.2.4).

Scandium was within error of no net change in sericitised, previously unaltered biotite-
quartz diorite and in moderately sericitised, previously potassically-altered biotite-quartz
diorite, but was gained by 5.2±0.5 ppm in very intensely-sericitised, previously
potassically-altered biotite-quartz diorite (Fig. 5.35). Yttrium was nearly always conserved
during sericitisation of potassically-altered biotite-quartz diorite, but may have been
removed during sericitisation of unaltered biotite-quartz diorite by 2.2±2.0 ppm (Fig. 5.35).

Figure 5.35: Bar chart of net change of REE-affinity elements Sc and Y during sericitic alteration compared with
the average least-altered biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an average potassically-
altered biotite-quartz diorite (“Sericite_M”) or very strong sericitisation of an average potassically-altered
biotite-quartz diorite (“Sericite_VS”). Error bars were calculated through propagation of error (section 4.2.4).
151
Chapter Five: Alteration geochemistry

5.3.3.2 Metals

The most important economic metals at Cerro Corona, Cu and Au, were added during very
intense sericitisation of potassically-altered protolith (Cu by ~9000 ppm and Au by ~1250
ppb, Fig. 5.36). However, during sericitisation of previously unaltered and moderate
sericitisation of potassically-altered biotite-quartz diorite these metals were within error
of no net change (Fig. 5.36).

Silver, Pb, Ga, V, Zn, and Cr showed similar behaviour during sericitisation. They were
within error of no change during sericitisation of a least-altered biotite-quartz diorite, but
were added during moderate and very strong sericitisation of previously potassically-
altered biotite-quartz diorite (Fig. 5.36). Molybdenum was added by ~18±7 ppm in very
strong sericitisation of potassically-altered biotite-quartz diorite, but was within error of
no change when either this protolith was already moderately sericitised or where the least-
altered biotite-quartz diorite was sericitically altered (Fig. 5.36). Thorium and Cs were
within error of no change in all sericitised samples (Fig. 5.36).

152
Chapter Five: Alteration geochemistry

Figure 5.36: Bar charts of net changes of metals during sericitic alteration compared with the average least-
altered biotite-quartz diorite (“Sericite_U”), moderate sericitisation of an average potassically-altered biotite-
quartz diorite (“Sericite_M”) or very strong sericitisation of an average potassically-altered biotite-quartz diorite
(“Sericite_VS”). Error bars were calculated through propagation of error (section 4.2.4). All changes are in ppm
except for Au, where net changes are in ppb.

5.3.4 Argillic alteration


The intense alteration of this alteration facies has led to textural destruction, thus
preventing the confident recognition of pre-existing alteration types (or even lithology).
Therefore, the mass transfer calculations were done with respect to the average of the
least-altered biotite diorite porphyry. This approach calculated the total net changes in
bulk mass and element concentrations from the fresh protolith to the argillically-altered
end-point, ignoring any intermediary alteration steps. The argillic alteration was separated
into groups based on the dominant clay mineral assemblage: dickite-dominated (n=3),
kaolinite-smectite (n=3) and illite-smectite (n=4). The ISOCON slopes for these
assemblages were 0.41±0.01 (Fig. 5.37), 0.78±0.05 and 0.79±0.01, reflecting bulk mass
gains of 143±5%, 28±8% and 26±2%, respectively. The most significant changes were
during the generation of dickite-dominated assemblages, with less extensive modification
of the rock during the formation of kaolinite and illite-smectite assemblages.

153
Chapter Five: Alteration geochemistry

Figure 5.37: An ISOCON of the average dickite-dominant argillically-altered biotite-quartz diorite (samples
C11.JL.95, C11.JL.108 and C11.JL.110) compared with the average least-altered biotite-quartz diorite using the
least squares approach (Baumgartner and Olsen 1995). Error bars are the standard deviation from these
averages.

5.3.4.1 The Major oxides

The dickite-dominated assemblage showed net gains of SiO2, Fe2O3, LOI and S by 22±14,
19±5, 13±1 and 0.8±0.3%, respectively (Fig. 5.38). There was a notable addition of P2O5, by
2±1%. Potassium oxide, MgO and Na2O were lost by 0.7±0.2, 0.6±0.1 and 1.3±0.2%,
respectively (Fig. 5.38). Calcium oxide appeared to be within error of no net gain or loss
(Fig. 5.38).

The kaolinite-smectite assemblage showed similar major component changes to the


dickite-dominant assemblage, albeit not as extensive (Fig. 5.38). SiO2, Fe2O3, LOI and S were
added by 20±3, 4±2, 4±2 and 3±2%, respectively. Calcium oxide and Na2O were removed
by 2.2±0.3 and 2.3±0.5 %, respectively. Unlike the dickite-dominant assemblage, P2O5, MgO
and K2O were within error of no net gain or loss (Fig. 5.38).

The illite-smectite-dominated assemblage had similar net changes of major components to


the other argillic alteration stages (Fig. 5.38). There was a net addition of SiO2, LOI and S by
15±4, 6±2 and 3±1% and a loss of Na2O by 2.4±0.3%. Calcium oxide, Fe2O3, K2O, MgO and
P2O5 were within error of no change.
154
Chapter Five: Alteration geochemistry

Figure 5.38: Bar chart of net changes of major components during argillic alteration compared with an average
least-altered biotite-quartz diorite. “Dickite ave” is the average of dickite-dominant argillically-altered samples,
“K-S ave” is the average of kaolinite-smectite-dominant argillically-altered samples and “I-S ave” is the average
of illite-smectite-dominant argillically-altered samples compared with the average least-altered biotite-quartz
diorite. Error bars were calculated through propagation of error (section 4.2.4).

5.3.4.1 High Field Strength Elements

The results for the HFSE in argillically-altered biotite-quartz diorite showed that Nb was
consistently removed, Th was added in the illite-smectite assemblage and Ta was added in
dickite- and kaolinite-smectite-dominant argillic alteration, but otherwise these elements
were within error of no net change (Fig. 5.39).

Figure 5.39: Bar chart of net changes of HFSE during argillic alteration compared with the average least-altered
biotite-quartz diorite compared with an average least-altered biotite-quartz diorite. “Dickite ave” is the average
of dickite-dominant argillically altered samples, “K-S ave” is the average of kaolinite-smectite dominant
argillically altered samples and “I-S ave” is the average of illite-smectite dominant argillically altered samples
compared with the average least-altered biotite-quartz diorite. Error bars are calculated through propagation of
error (section 4.2.4)
155
Chapter Five: Alteration geochemistry

5.3.4.2 Large Ion Lithophile Elements

Dickite-bearing argillic alteration showed loss of Rb and Ba by 16±4 and 139±5 ppm, but
no net change of Sr, whereas the kaolinite-smectite and illite-smectite argillic alteration
types were within error of no net change for Ba, and lost Sr by 414±48 and 356±38 ppm
respectively (Fig. 5.40). Rubidium was within error of no net change in the kaolinite-
smectite assemblage, but was added by 15±10 ppm in the illite-smectite assemblage.

Figure 5.40: Bar chart of net changes of LILE in argillically altered samples compared with an average least-
altered biotite-quartz diorite. “Dickite ave” is the average of dickite-dominant argillically altered samples, “K-S
ave” is the average of kaolinite-smectite dominant argillically altered samples and “I-S ave” is the average of
illite-smectite dominant argillically altered samples compared with the average least-altered biotite-quartz
diorite. Error bars were calculated through propagation of error (section 4.2.4).

5.3.4.1 Rare Earth Elements

The rare earth element gains in dickite-altered samples are noteworthy, being an order of
magnitude higher than any other alteration stage encountered (Fig. 5.41). These samples
were enriched in all REE, particularly the LREE and Y (Fig. 5.41). The addition of HREE
within the dickite alteration was particularly anomalous. Kaolinite-smectite samples
broadly showed net loss of REE whereas illite-smectite facies samples showed no net
change in the REE (Fig. 5.41).

156
Chapter Five: Alteration geochemistry

Figure 5.41: Bar charts of net changes of REE during argillic alteration compared with an average least-altered
biotite-quartz diorite. “Dickite ave” is the average of dickite-dominant argillically altered samples, “K-S ave” is
the average of kaolinite-smectite dominant argillically altered samples and “I-S ave” is the average of illite-
smectite dominant argillically altered samples compared with the average least-altered biotite-quartz diorite.
Error bars were calculated through propagation of error (section 4.2.4.)

Scandium and Y were added to dickite argillic alteration by 5±1 ppm and 42±1 ppm,
respectively (Fig. 5.42). Kaolinite-smectite-altered samples had an addition of Sc and
removal of Y, by 3±1 and 4±2 ppm, respectively. The illite-smectite alteration appeared to
be within error of no net change of either Sc or Y (Fig. 5.42).

157
Chapter Five: Alteration geochemistry

Figure 5.42: Bar chart of net changes of the REE-affinity elements Sc and Y in argillic alteration compared with
the average least-altered biotite-quartz diorite.

5.3.4.2 Metals

Metals appear to have been added during the argillic alteration stages, especially the dickite
stable samples. The exception to this was Cr which was lost during kaolinite-smectite and
illite-smectite alteration types (Fig. 5.43).

Dickite-altered samples show massive gains of Au, Cu, Zn Pb and Cr, by 204±0.2 ppb,
~6.3±0.3%, ~930±40 ppm, 320±10 ppm and 150±20 ppm, respectively (Fig. 5.43).
Significant net gains were seen in As, Co, Se, Ni, Sn and Th during this alteration stage.
Molybdenum, Li, Ga, Cd and Sb were added by minor amounts (12±7, 8±2, 5.0±0.6, 5±1 and
1.3±0.7 ppm, respectively, Fig. 5.43).

Kaolinite-smectite-altered samples had similar ore metal gains as dickite-altered samples


(Fig. 5.43), except the additions were not as great. Chromium was lost and Co, Ni, Sn, Mo,
Li, Ga, Cd and Sb were within error of no change.

Illite-smectite alteration showed gains similar to kaolinite-smectite altered samples, in that


Cu was gained more so than Au and Pb (Fig. 5.43). Zinc and Cr are within error of no net
change as are Ni, Sn, Mo, Li, Cd and Sb. Gallium was added by 4±1ppm (Fig. 5.43).

158
Chapter Five: Alteration geochemistry

Figure 5.42: Bar chart of net changes of elements in argillic alteration compared with the average least-altered
biotite-quartz diorite. A.) Major contributions to metals are seen in predominantly dickite-dominant alteration
assemblages, with less significant gains in kaolinite-illite and smectite-illite assemblages. B.) Notice the
significant gains in metals in dickite-dominant argillic alteration.

5.3.5 Model of chemical mass transfer


The ISOCON method was applied to mine assay data from 11755 ~2m composite drill core
samples. This was done to understand the main features of major component transfer
during multiple stages of alteration at Cerro Corona. The goal was to identify notable zones
of net gain or loss of elements within the deposit.

Zirconium was not determined in most whole rock assays, thus was not used for mass
transfer analyses. The entire dataset shows a weak correlation between Al2O3 and TiO2 (Fig.
159
Chapter Five: Alteration geochemistry

5.44) but with two main trends. The majority samples have a TiO2/Al2O3 ratio to the
potassically-altered protolith, between 0.01 and 0.035 (Fig. 5.44), and so these were
selected for the mass transfer analysis on the assumption that they reflect variable degrees
of overprinting of potassically-altered biotite-quartz diorite. This assumption is supported
by core observations, in which most biotite-quartz diorite porphyry contained quartz
stockwork, quartz-magnetite veinlets, shreddy biotite (and/or) relic alkali feldspar
indicative of initial potassic alteration. The exclusion of samples with higher TiO2/Al2O3
ratios and the host limestone results in a ~20% reduction of the data but drastic
improvement in the correlation coefficient (Fig. 5.44).

Figure 5.43: TiO2 versus Al2O3 for assay data (n = 11755) showing the least-altered average biotite-quartz diorite
(grey) and potassically-altered biotite-quartz diorite (pink) with all assay data (A) and with assay data for
samples attributed just to a potassically-altered biotite-quartz diorite protolith (B).

160
Chapter Five: Alteration geochemistry

The ISOCON method was applied to the reduced dataset (n=11755) using the average of
four samples of potassically-altered biotite-quartz diorite (section 5.2.2) as the protolith
composition. Samples with errors in bulk mass change exceeding the bulk mass change
values themselves were excluded. These samples (n=515) have TiO2/Al2O3 ratio that are
tangential to the overall TiO2/Al2O3 trend toward to the origin.

5.3.5.1 Bulk mass change

The bulk mass change results show that the data have an overall skew toward mass gain
(i.e. conservative elements are diluted, toward the origin) as expected by conservative
element trends (Fig. 5.44). A significant proportion of samples exceed 100% mass gain,
potentially due to significant dilution and low concentrations of the conservative elements
(Fig. 5.45).

Figure 5.44: Histogram of bulk mass change of altered samples (n=11240).

The calculated bulk mass gain values were plotted in 3 dimensions using Leapfrog Geo and
then contoured using an adaptive resolution of 100 m, with overall ellipsoid ratio of 5:2:2
(maximum: intermediate: minimum). The areas of lowest mass change appear to coincide
with areas where potassic alteration is preserved, as well as with the locations of the
“Barren Cores” (Fig. 5.45), consistent with these being largely unaltered, late intrusions. At
depth, the southern margin of the intrusive complex appears to have little mass gain (Fig.
5.45), as expected due to the lack of extensive overprinting in this area.
161
Chapter Five: Alteration geochemistry

Areas of very high (>100%) bulk mass gain are located on the northern and south-eastern
margins (Figs. 5.45 and 5.46). These areas appear to be zones of extensive quartz veining
(Gold Fields Limited 2005) and orthoclase and magnetite precipitation, suggesting a
secondary potassic alteration overprint in this area. The zones of less drastic bulk mass
gain are sub-horizontal and occur within the centre of the host intrusion (Fig. 5.46).

Figure 5.45: Plan view at 3810 m elevation showing contours of bulk mass change. Black represents bulk mass
change of +10%, orange represents bulk mass change of +60% and red represents a bulk mass change of +100%.
162
Chapter Five: Alteration geochemistry

Figure 5.46: 763300E cross section across the Cerro Corona deposit contoured by bulk mass change. The black,
near horizontal line represents ground surface, and thick dark blue line is proposed level of the final open pit.

163
Chapter Five: Alteration geochemistry

The net change of major components is considered within error of no change if the
propagated error exceeds the calculated net change. For the major components Fe2O3, K2O,
Na2O, CaO, P2O5, MgO and S, approximately 21, 32, 16, 19, 92, 52 and 67 percent of data are
within error of no net change, respectively. These specific results were excluded from the
mass transfer modelling.

The calculated net transfer of major components provides an insight into their
redistribution by overprinting alteration stages, taking into account bulk mass change. This
provides a more realistic understanding of the metasomatic processes that occurred within
the deposit and is described in the following sections.

5.3.5.2 K2O

The calculated net changes in K2O have an asymmetrical, negative skew (Fig. 5.47) with
almost all results being negative, indicating pervasive K2O removal during overprinting
alteration. There is a small outlying population at 1.5% gain; this K2O addition occurs as a
plug in the northern area of the host intrusion (Figs. 5.48 and 5.49), coincidental with the
area of >100% bulk mass gain (Fig. 5.46).

Figure 5.47: Histogram of net change in K2O relative to precursor potassic alteration (n=8025).

164
Chapter Five: Alteration geochemistry

Figure 5.48: Plan view at 3810 m elevation with contours of net change in K2O.

165
Chapter Five: Alteration geochemistry

Figure 5.49: 763300E cross-section across the Cerro Corona deposit contoured for net change in K2O. The black,
near horizontal line represents topography, and thick dark blue line is proposed level of the final open pit.

166
Chapter Five: Alteration geochemistry

5.3.5.3 Na2O

Na2O is removed from the majority of samples and broadly appears to have a normal
distribution around ~1% loss, with a smaller population of Na2O gain round +1% (Fig.
5.50).

Figure 5.50: Histogram of net change in Na2O of drill core samples relative to precursor potassic alteration
(n=9867).

The small population of Na2O net gain appears to be in areas of no bulk mass change (Figs.
5.46 and 5.52), suggesting that this may reflect an area of less intense original potassic
alteration or a weak, previously unidentified, sodic alteration. At shallow levels, Na2O is
removed by around 2 to 1% (Fig. 5.51). In cross section, the Na2O loss appears to be
deflected upwards in the north (Fig. 5.52), coincidental with areas of bulk mass gain >100%
(Fig. 5.46) and K2O addition (Fig. 5.49). This suggests that Na2O is less readily removed
during the inferred second stage of potassic alteration. The removal of Na2O is deflected
downward in the centre and south of the host intrusion (Fig. 5.52).

167
Chapter Five: Alteration geochemistry

Figure 5.51: Plan view at 3810 m elevation with contours of net change in Na2O.

168
Chapter Five: Alteration geochemistry

Figure 5.52: 763300E cross-section across the Cerro Corona deposit contoured by Na2O net change. The black,
near horizontal line represents topography, and thick dark blue line is proposed level of the final open pit.

169
Chapter Five: Alteration geochemistry

5.3.5.4 CaO

CaO net change is bimodal in distribution (Fig. 5.53), with the largest proportion of samples
having loss of CaO by around 1.0%. A second population appears to be distributed around
1.5% net gain.

Figure 5.53: Histogram of net change of CaO in drill core samples relative to precursor potassic alteration
(n=9491).

Calcium oxide removal forms a concentric zone around the post-mineralisation intrusions
(Fig. 5.54), especially at shallow levels (Fig. 5.55). Addition is coincidental with post-
mineral intrusions and the limestone host. At depth, it is generally added by more than 1%.
The largest net CaO gain is proximal to the Pariatambo Formation at depth, in the south
section of the deposit (Fig. 5.55), suggesting some transfer of calcium from the host
sequence into the intrusive complex, either by xenolith entrainment of my metasomatism.

170
Chapter Five: Alteration geochemistry

Figure 5.54: Plan view at 3810 m elevation with contours of CaO net change. Note the areas of CaO net loss are
concentric around the post-mineral intrusions.

171
Chapter Five: Alteration geochemistry

Figure 5.55: 763300E cross section across the Cerro Corona deposit contoured by CaO net change. The black,
near horizontal line represents topography, and thick dark blue line is proposed level of the final open pit.

172
Chapter Five: Alteration geochemistry

5.3.5.5 Fe2O3

The net changes in Fe2O3 are highly variable (Fig. 5.56). It is worth noting that the detailed
mass transfer analysis showed that Fe2O3 is generally added during all alteration stages,
with the largest additions occurring in the propylitic and sericitic alteration stages. Given
this, the relatively frequent occurrence of small net losses in the drillcore assay data is
unexpected.

Figure 5.56: Histogram of net changes in Fe2O3 in drill core samples relative to precursor potassic alteration
(n=9319).

In 3 dimensions, the major net gains of Fe2O3 occur in a concentric zone centred on the
western post-mineral intrusion (Fig. 5.57), with a secondary addition in the north and
central parts of the deposit to a depth of around 3500 m (Fig. 5.58). The northern gains
appear coincidental with the region of high bulk mass gain (Fig. 5.46), net gains in K2O (Fig.
5.49), upward deflection of net Na2O (Fig. 5.52) loss and net CaO loss (Fig. 5.55).

173
Chapter Five: Alteration geochemistry

Figure 5.57: Plan view at 3810 m elevation with contours of net change in Fe2O3. Note all the Fe addition is
concentrated in the western half of the intrusive complex.

174
Chapter Five: Alteration geochemistry

Figure 5.58: 763300E cross-section across the Cerro Corona deposit contoured by Fe2O3 net change. The black,
near horizontal line represents topography, and thick dark blue line is proposed level of the final open pit.

175
Chapter Five: Alteration geochemistry

5.3.5.6 P2O5

The net change in P2O5 is asymmetric, with most samples showing no net change (Fig. 5.59).
Absolute concentrations of P2O5 in whole rock samples are relatively low, an average of
0.16 wt. % (n=11755), thus the calculated error in net change commonly exceeds the
absolute net change. Nevertheless, the reduced dataset after these analyses have been
removed broadly mimics that of the whole dataset in that most of the data are within error
of no net change (Fig. 5.59).

Figure 5.59: Histogram of net change of P2O5 in drill core samples relative to precursor potassic alteration. The
darker fill with red outline represents the reduced dataset, after removal of results in which the calculated net
change is larger than the error (n=940). The bars with a light blue fill represent the entire dataset for
comparison (n=11755).

AS expected, P2O5 broadly shows no net change throughout the host intrusive (Figs. 5.60
and 5.61). Areas of net gain are proximal to the limestone contact at shallow levels in the
south (Fig. 5.61). Minor phosphate addition is noted at depth in the north of the intrusive.
Minor P2O5 loss appears to coincide with the two “barren cores”, although this could reflect
slightly lower P2O5 as a fundamental characteristic of the inferred late intrusions.

176
Chapter Five: Alteration geochemistry

Figure 5.60: Plan view at 3810 m elevation with contours of net change of P2O5.

177
Chapter Five: Alteration geochemistry

Figure 5.61: 763300E cross section across the Cerro Corona deposit contoured by P2O5 net change. The black,
near horizontal line represents topography, and thick dark blue line is proposed level of the final open pit.

178
Chapter Five: Alteration geochemistry

5.3.5.7 S

The calculated net changes in sulphur of the reduced dataset show a relatively broad
distribution with a positive skew (Fig. 5.62). Most samples show an addition of >2.5%
relative to the average potassic alteration. Sulphur addition is seen in the overprinting
propylitic (Fig. 5.20) and more drastically in sericitic alteration (Fig. 5.31), in which pyrite
is ubiquitous.

Figure 5.62: Histogram of calculated net change in S relative to an average potassic altered protolith (n=3651).

In 3 dimensions, the zone of greatest net gain in sulphur is within the southwestern area of
the host intrusive complex (Fig. 5.63) but it is broadly gained throughout the centre of the
system (Fig. 5.63 and Fig. 5.64).

179
Chapter Five: Alteration geochemistry

Figure 5.63: Plan view at 3810 m elevation with contours of net change in S. Note similar pattern to net Fe2O3
addition.

180
Chapter Five: Alteration geochemistry

Figure 5.64: 763300E cross section across the Cerro Corona deposit contoured by net change in S. The black, near
horizontal line represents topography, and thick dark blue line is proposed level of the final open pit.

181
Chapter Five: Alteration geochemistry

5.4 References
Baumgartner, L. P. and Olsen, S. N. (1995) A least-squares approach to mass transport
calculations using the isocon method. Economic Geology. 90 (5), 1261-1270.

Departamento de geologia. (2005) Distribucion de intensidad de venillas. Internal map, Gold


Fields Limited.

Grant, J. A. (2005) Isocon analysis: A brief review of the method and applications. Physics
and Chemistry of the Earth, Parts A/B/C. 30 (17–18), 997-1004.

Grant, J. A. (1986) The Isocon Diagram - a Simple Solution to Gresens Equation for
Metasomatic Alteration. Economic Geology. 81 (8), 1976-1982.

Gresens, R. L. (1967) Composition-volume relationships of metasomatism. Chemical


Geology. 247-65.

Gifkins, C., Herrmann, W. and Large, R. R. (2005) Altered Volcanic Rocks: A Guide to
Description and Interpretation. 1st edition. University of Tasmania, Centre for Ore Deposit
Research.

Idrus, A., Kolb, J. and Meyer, F. M. (2009) Mineralogy, Lithogeochemistry and Elemental
Mass Balance of the Hydrothermal Alteration Associated with the Gold-rich Batu Hijau
Porphyry Copper Deposit, Sumbawa Island, Indonesia. Resource Geology. 59 (3), 215-230.

Ku, H. H. (1966) Notes on the use of propagation of error formulas. Journal of Research of
the National Bureau of Standards. 70 (4), 263-273.

Mukherjee, P. K. and Gupta, P. K. (2008) Arbitrary scaling in ISOCON method of geochemical


mass balance: An evaluation of the graphical approach. Geochemical Journal. 42 (3), 247-
253.

Ulrich, T. and Heinrich, C. A. (2002) Geology and Alteration Geochemistry of the Porphyry
Cu-Au Deposit at Bajo de la Alumbrera, Argentina. Economic Geology. 97 (8), 1865-1888.

182
Chapter Six: Infrared spectroscopy of clay minerals

6 Chapter Six: Infrared spectroscopy of clay


minerals

6.1 Introduction
Clay minerals are challenging to distinguish owing to their small grain size (usually <2 μm)
without the aid of x-ray diffraction of IR spectroscopic measurements. This chapter aims to
differentiate between clay mineral assemblages and their distribution utilising IR
spectroscopy (outlined in Chapter 1).

6.2 Methods

6.2.1 IR spectroscopy
The ASD TerraSpec™ 4 is a handheld infrared spectroradiometer with spectral accuracy
and reproducibility of 0.5 and 0.1 nm respectively (Table 6.1). Three detectors allow for
rapid collection (10 spectra per second) across the spectral range of the instrument
(350nm-2500 nm). The three detectors are configured to detect the reflectance at three
electromagnetic ranges (Table 6.1). The spectral resolution is 3 nm for VNIR and 6 nm for
SWIR. The quoted signal-to-noise ratio for all detectors is <9000:1.

The fibre optic cable is tested prior to the start of scans and the instrument was run for 30
minutes prior to analysis, which allowed it to “warm up” and the instrument to stabilise. A
“dark current” measurement was taken by collecting 200 spectra without exposing the
instrument to light. This allows for internal calibration and removal of the electromagnetic
contribution from the instrument. A measurement of white reference (Spectralon ®) is also
acquired every 30 minutes, to provide a white reference material as this material has a high
diffuse reflectance and is chemically inert.

Table 6.1: The technical specifications of ASD TerraSpec™ 4.

Wavelength Range 350-2500 nm


SWIR 1(1001-1800 nm) and
VNIR (350-1000 nm) SWIR 2(1801-2500 nm) are both
Detectors 512 element silicon graded index indium gallium
photodiode array arsenide photodiodes which are
thermoelectrically cooled
Resolution 3 nm @700nm 6 nm @ 1400 and 2100 nm ranges
Wavelength reproducibility 0.1 nm
Wavelength accuracy 0.5 nm
183
Chapter Six: Infrared spectroscopy of clay minerals

6.2.1.1 Data collection

A total of 4803 spectra were collected over two field seasons: 3011 in 2011 and a further
1792 in 2012. A spectral reading was taken approximately every metre, with a bias towards
clay minerals. The areas that were selected for analysis were prepared by scratching the
surface of the drillcore with a tungsten carbide tipped pen to ensure a fresh surface is
exposed, minerals are randomly orientated and to maximise the surface area. The spectral
number, drillcore number and depth of measurement along with preliminary mineral
identification was recorded.

6.2.1.2 Data processing

Raw spectra were “spliced” utilising ViewSpec Pro Version 6.0.11. This function calculates a
bias for the VNIR and SWIR2 detectors, such that they match the SWIR 1 spectrum without
any offset at the point of connection. This splice correction does not affect the quality of the
data and is primarily for cosmetic appearance and comparison between spectra.

Following the splice correction, a concave hull is fitted to the general shape of the spectra
utilising the local spectral highs (Green and Craig 1985). This hull fitting is primarily to
attribute broad spectral suppression commonly seen in multimineral surfaces that may
suppress a larger range of the spectrum and that mathematically shift the absorption
features. This fitted hull is subsequently straightened by dividing the depth of the spectral
feature by the hull, a process known as “hull removal” or “hull quotient method”. This
removes wavelength-dependent scattering, enhances and standardizes the absorption
features and aids in comparison between spectra. This processing was done utilising
SpecWin version 2.0. The importance of this step is commonly disregarded, but is crucial in
identifying spectral shifts (Clark and Roush 1984). The band depth (DB) is defined as:

𝑅 −𝑅
𝐷 =
𝑅

where: RC is the reflectance of the continuum band centre and RB is the reflectance at the
band centre (Clark and Roush 1984).

The depth and position of absorption features are extracted using the “show absorption”
function of SpecWin. This function extracts the depth and position of absorption features
from a hull-removed spectrum (Fig. 6.1). A macro was run which exported all absorption
positions and depths for each spectrum into a workable .xls file format and combined into
a single file (Appendix 5).

184
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.1: Example of the full infrared spectrum (red line) and absorption features extracted from the hull-
removed spectrum (dashed black line). Note that the relative depths of the absorption features have changed in
the hull-removed data.

6.2.1.3 Clay mineral identification

Kaolin group

The kaolin group is identified by distinctive asymmetrical doublets (i.e. two minima in close
proximity) 1400 nm and 2190 nm. The 1400 nm doublet is attributed to the first overtone
of OH- stretching mode (2𝑣 ) (Hunt and Salisbury 1970), whereas the 2190 nm doublet is
attributed to combination of OH- stretching and bending of Al2-OH. The presence of a
doublet in these positions suggests that the OH- groups are in slightly different sites as the
energy required for the bonds to oscillate are not equal (Hunt 1977).

The 1415 nm absorption feature is attributed to the first order overtone of the “inner
hydroxyl” of the kaolin group (Fig. 6.2), with the shorter wavelength absorption between
1360-1400 nm being related to the hydrogen bonded OH- group on the exterior of the
octahedral sheet (Crowley and Vergo 1988). The absorption near 2200 nm is attributed to
a combination of OH- stretching with bending of xAlOH (where x=Al, Ga, Fe, or Cr). The
precise locations of the absorptions at 1415 nm and 2200 nm in the kaolin group are
dependent on the cation to which the hydroxyl is bonded in the octahedral site due to the
slightly different energies required to bond with these cations (Delineau et al. 1994, Petit
et al. 1999, Table 6.2).

185
Chapter Six: Infrared spectroscopy of clay minerals

Table 6.2: The position of the minima of absorption features as a function of cationic pairs in octahedral sites to
which the hydroxyl is bonded in kaolinites (Petit et al. 1999).

Bond vibration x=Al x=Ga3+ x=Fe3+ x=Cr3+


2vOH 1415 1425 1425 1431
vOH+δAl-x OH 2209 2223 2239 2235

The position of the first overtone of the inner hydroxyl bond stretch at 1415 nm is similar
in all kaolin groups, as the location of the inner hydroxyl is directed toward the empty
octahedral sites in all kaolin polymorphs (Giese and Datta 1973). This similar location of
the inner hydroxyl allows similar frequency of vibration (Frost and Johannson 1998)
provided that the cations to which the hydroxyls are bonded are the same in all polymorphs
(Petit et al. 1999).

In dickite, the hydrogen bonding between the outer hydroxyl group (Fig. 6.2) and the apical
oxygen of the silica tetrahedron is stronger than in kaolinite (Giese and Datta 1973), due to
the difference in layer stacking, which is slightly off-centred compared to that of kaolinite
(and halloysite?). Thus, there is a more consistent orientation of “inner surface” hydroxyls
in dickite than in kaolinites (Benco et al. 2001; Giese and Datta 1973). The fundamental
stretching of the “inner surface” hydroxyl groups is therefore different between dickite and
kaolinite (Balan et al. 2010), occurring at shorter wavelengths in dickite than in kaolinite
(Fig. 6.3 and Fig. 6.4).

Figure 6.2: Profile view (100) of kaolinite with single Al-octahedral sheet bonded to a Si-tetrahedral sheet, thus a
1:1 phyllosilicate. Two cavities are occupied by Al ions. Two hydroxyl groups are seen: 1) the outer hydroxyl
groups or “inner surface” hydroxyls; and 2) the inner hydroxyl groups (termed by Frost and Johannson 1998).
Si=khaki, Al=blue, O=red and H=white. Adapted from Geysermans and Noguera (2009).

186
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.3 Examples of infrared spectra of kaolin group minerals, arbitrarily offset for clarity. The dickite
spectrum is displayed in red, kaolinite in orange and halloysite in green. Note the doublets are the distinguishing
features of this group of minerals, and are most intense for dickite. The 1900 nm feature increases in depth from
dickite->kaolinite->halloysite

Figure 6.4: (A) Location of the first overtone stretch (2vOH) bands related to inner surface and inner hydroxyl
vibrations in the kaolin group minerals. Note the position of the inner hydroxyl absorption band is similar in all
kaolin groups, whereas the position of absorption bands related to inner surface hydroxyl is slightly shorter for
dickite (red), than for kaolinite (blue). (B)Absorption bands related to the combination mode of OH- stretching
(vOH) of inner surface and inner hydroxyl bands with bending of Al2OH groups.

The 1900 nm feature is related to vibration of the O-H bonds in the water molecule (vw+
δw). This can be used as a proxy for the hydration of the kaolin (Hunt and Salisbury 1970)
and is removed when samples are placed under vacuum (Delineau et al. 1994). The

187
Chapter Six: Infrared spectroscopy of clay minerals

intensity of this water molecule absorption band increases with increasing adsorption onto
the clay minerals. The presence of this hydration feature is commonly utilised to
distinguish halloysite-(10Å) from other kaolin groups (Joussein et al. 2005), but is seen in
other mineral phases so is not wholly diagnostic. In this thesis, “halloysite” refers to
hydrated halloysite-(10Å) as well as highly disordered kaolinite.

Distinction between kaolinite and halloysite in SWIR spectroscopy is challenging owing to


the similar position of adsorption features (Fig. 6.4). The intensity of the absorption feature
of the inner surface hydroxyl stretch relative to the inner hydroxyl is considered a proxy
for order in kaolinite (Parker 1969). Subsequent research on the disorder of kaolinite (or
“kaolinite crystallinity”) suggested this disorder is a result of octahedral site vacancy
(Plançon and Tchoubar 1977) or Fe3+ substitution into the octahedral site (Brindley et al.
1986). This octahedral discrepancy is reflected in the modification of OH- stretching band
(Barrios et al. 1977). The intensity ratio of kaolinite is demonstrably linked to the

kaolinite crystallinity, measured using the Hinckley Index, whereby the ratio increases

with increasing ‘kaolinite crystallinity’ (Crowley and Vergo 1988, Zhang et al. 2001). To
avoid confusion, the ‘kaolinite crystallinity index’ here refers to the disorder occurring from
modification of OH- stretching band, whether by Al vacancy or Fe3+ substitution
(Guggenheim et al. 2006). This ratio measures kaolinite disorder, thus high-disorder
kaolinites are grouped together with halloysite and both are referred to "halloysite" in this
thesis.

The primary distinction between the kaolin group minerals is the position and relative
intensities of the inner surface hydroxyl and the inner hydroxyl (Table 6.3). Further
distinction is made using the adsorption of water into the interlayer site between layers
(Table 6.3).

Table 6.3: Major absorption feature positions of kaolin group minerals which have been used for distinction.

Kaolin Kaolinite
1400 nm doublet 2200 nm doublet
group "crystallinity" index
mineral 2vOH-inner 2vOH-inner 2vOH-inner surface / vOH-inner surfacevOH-inner hydroxyl

surface hydroxyl 2vOH-inner hydroxyl + δAl2OH + δAl2OH


Dickite 1375-1390 1405-1420 N/A 2175-2185 2200-2215
Kaolinite 1385-1404 1405-1420 >0.8 2155-2175 2200-2215
Halloysite 1385-1404 1405-1420 <0.8 2155-2175 2200-2215

188
Chapter Six: Infrared spectroscopy of clay minerals

Illite

A profile view of the illite clay structure indicates that the hydroxyl is located within the Al
octahedral layer (Fig. 6.5). The Al-OH bond strength in illite is comparable to that seen in
kaolinite, thus the illite spectrum overlaps with that of the kaolin group minerals in the
1400 nm and 2200 nm regions, making identification of mixtures of more than two
minerals challenging (Kerr et al. 2011).

Figure 6.5: Profile view (100) of illite with single Al-octahedral sheet sandwiched between two Si-tetrahedral
sheets, thus a 2:1 phyllosilicate. The charge imbalance across the layer is partially balanced by interlayer K+.
Modified from Grim (1962).

The reflectance spectrum of illite is also very similar to that of muscovite, due to their
similar crystal structure. The characteristic features of this dioctahedral phyllosilicate are
in the 1400 and 2200nm regions (Fig. 6.6). Absorption features in these regions correspond
to the overtone stretch of the hydroxyl group within the octahedral layer (2vOH) at ~1400
nm and the combination band of the hydroxyl stretch and bend of octahedral cation-
hydroxyl (vOH+δXAlOH) at ~2200 nm (Hunt et al. 1977).

189
Chapter Six: Infrared spectroscopy of clay minerals

Minor features at ~ 2340 nm and 2440 nm are present in muscovite and illite that
distinguish them from smectite (Fig. 6.6; Clark et al. 1990). The 2340 nm feature occurs at
a shorter wavelength and the 2450 nm feature is less pronounced in illite than in muscovite
(Clark et. al. 1990).

The illite “crystallinity” here is calculated as a ratio between the depth of the combination
band of hydroxyl stretch and bend of the octahedral cation-hydroxyl (vOH+δXAlOH) at 2200
nm and the depth of the 2340nm feature.

𝐷𝑒𝑝𝑡ℎ 2200 𝑛𝑚
𝐼𝑙𝑙𝑖𝑡𝑒 crystallinity 𝑖𝑛𝑑𝑒𝑥 =
𝐷𝑒𝑝𝑡ℎ 2340 𝑛𝑚

The location of the vOH+δXAlOH absorption band is a function of the abundance of aluminium
in the octahedral site. The wavelength decreases (from 2225 nm to 2194 nm) with
increasing Al content (from 2.0-2.8 Al per 11O) (Duke and Lewis 2010). This characteristic
shift is commonly utilised as an indicator of the temperature of formation, because a higher
Al content correlates to a higher formation temperature.

In this study illite is thus identified by the relative depths of the continuum removed
spectrum (Fig. 6.6), where:

𝐷𝑒𝑝𝑡ℎ > 𝐷𝑒𝑝𝑡ℎ > 100

𝐷𝑒𝑝𝑡ℎ > 𝐷𝑒𝑝𝑡ℎ

𝐷𝑒𝑝𝑡ℎ 𝑎𝑛𝑑 𝐷𝑒𝑝𝑡ℎ > 100

Illite comprises of interlayer water, unlike micas, thus the depth of the 1900 nm (vw+ δw),
is utilised to distinguish between illite and mica, where the depth of the continuum
removed 1900 nm feature is <100.

190
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.6: Example of an illite reflectance spectrum with distinguishing feature ranges highlighted in red. The
dashed red lines are the exact position and depth of these features in the continuum removed spectrum. The
absorption features located at 2340 and 2440 nm are used to distinguish from smectite, which do not have these
features.

Smectite

The most obvious features of the smectite mineral group are the absorption bands related
to the hydration. The diagnostic absorption features of this group are the combination of
stretching (v) and bending (δ) of the water molecule within the clay mineral structure. The
relative depth of the absorption at ~1900 nm due to bound water and the bonding and
stretching hydroxyl group at ~2200 nm is utilised to distinguish between smectite and
illite. When the depth of the 1900 nm is greater than the depth of the 2200 nm feature, the
sample is identified as primarily smectite. If an absorption band is present at ~2350 nm
then the illite-smectite mixture is inferred.

The smectites are a group of 2:1 phyllosilicates with a negative charge imbalance of 0.2-0.6
per formula unit within the phyllosilicate layers. This imbalance is caused by the
substitution of cations within the octahedral site and/or in the tetrahedral sites. The
resulting negative charge on one of the layers is balanced by interlayer exchangeable
cations, primarily Ca2+, Na+ and polar molecule H2O. (Fig. 6.7).

191
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.7: Profile view (100) of smectite with single Al-octahedral sheet sandwiched between two Si-tetrahedral
sheets, thus a 2:1 phyllosilicate. The charge imbalance across the layer is partially balanced by interlayer Ca2+,
Na+ and H2O.Modified from Grim (1962).

The dioctahedral group comprises of two trivalent cations (primarily Al3+ and Fe3+),
whereas the trioctahedral group is defined by three divalent cations (primarily Fe2+ and
Mg2+) within the octahedral site. This can be utilised in infrared spectroscopy to
differentiate between these two subgroups as there is a distinguishable shift between the
combination band of the hydroxyl stretch and bending of the X-OH band, where X=Al, Fe,
Mg (vOH + δXOH), depending on the cation to which the hydroxyl is bonded (Fig. 6.8).

Furthermore, the metal cation which bonds with the inner hydroxyl causes a shift from
2200 to 2300 nm. This shift is utilised for distinction between beidellite (Al-Al-OH),
montmorillonite (Al-Mg-OH), nontronite (Al-Fe-OH) and hectorite (Li-Mg-OH) (Fig. 6.8).

192
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.8: Combination bands of stretching (vOH) and bending (δ XAlOH) modes associated with the smectite group
minerals. The position of absorption bands is indicated by dashed lines and labelled with the band modes. The
band assignments are adopted from Sposito et al. (1983), Post and Noble (1993), Cariati et al. (1983), Bishop et
al. (2002) and Madejova and Komadel (2001).

A linear relationship between the position vOH+δX-OH and the aluminium content of the
smectite (Post and Noble 1993), allows the content the aluminium content to be calculated
from the wavelength position (𝑥):

𝐴𝑙 𝑂 (𝑤𝑡%) = −0.3803𝑥 + 861.39

The calculated error of this linear regression is ± 0.14% and the analytical error is ±0.2%,
based on the 0.5 nm accuracy of the ASD TerraSpec™.

The negatively charged layer of smectite is partially balanced by the interlayer water and
cations. This polarises and bends the water molecule within the smectite crystal (Cariati et
al. 1983). The polarised water molecules bond via weak hydrogen bonds to the inner
surface (v’w) and with stronger hydrogen bonds to adjacent water molecules (vw). The
difference in bond strength of these hydrogen bonds influences their wavelength positions,
such that the combination band stretching of the weakly bonded inner surface water (or
inner sphere) and the bending of the water molecule (v’w + δw) is located at around 1910
nm and the molecular water bonded to adjacent water molecules (adsorbed water) and the
bending of the water molecule (vw + δw) is located at around 1970 nm (Cariati et al. 1983).
The overtone fundamental combination bands behave similarly, with v’w + 2δw at ~1410 nm
and vw + 2δw at ~1460 nm (Fig. 6.9). The precise location of these bands is influenced by the
interlayer cation and the amount of adsorbed water (Bishop et al. 2002). An uneven

193
Chapter Six: Infrared spectroscopy of clay minerals

distribution of interlayer cations (Iwasaki et al. 1988) may influence the water molecule
residence. As there was no control of humidity during the collection of the spectra, the
degree of hydration is unknown and any further investigation of the impact of hydration is
beyond the scope of this study.

Figure 6.9: Spectral assignment of overtone and combination bands encountered in smectite. Sample spectra are
show for Ca-montmorillonite (blue), Na-montmorillonite (pink), nontronite (light blue) and hectorite (green)
(Clark et al. 1990). Dashed lines indicate stretching (v) and bending (δ) combination band positions and the
subscript indicates the bonds of hydroxyl (OH) or water (w). For water molecule absorptions, v’ indicates the
stretching of water bonded to the inner layer and v denotes the stretching of adsorbed water. Assignment of
positions of combination bands are from Cariati et al. (1983), Bishop et al. (2002). The spotted grey and blue
lines are unassigned features within the infrared spectrum.

194
Chapter Six: Infrared spectroscopy of clay minerals

Smectite is identified by the presence of absorption bands with:

𝐷𝑒𝑝𝑡ℎ > 𝐷𝑒𝑝𝑡ℎ

𝐷𝑒𝑝𝑡ℎ > 𝐷𝑒𝑝𝑡ℎ 𝑜𝑟 𝐷

The two dominant smectite group minerals found at Cerro Corona are montmorillonite and
beidellite. These are identified by a deeper 1900 nm feature than the 1400 and 2200 nm
absorptions related to the OH- stretch. Montmorillonite is distinguished by the higher
wavelength position of the 2200 nm feature (Fig. 6.9). The shift in the position of this
feature from ~2204 to ~2214 nm in montmorillonite is disputed to be either a function of
increasing calcium content (Clark et al. 1990) or decreasing aluminium content (Post and
Noble 1993). Beidellite is identified by the 2200 nm absorption feature between 2170-
2195 nm. No tri-octahedral subgroups smectites are identified at Cerro Corona.

Illite-smectite

Illite-smectite interlayers or samples containing admixtures of both illite and smectite are
subdivided by utilising the relative depths of the 1400 and 2300 nm features to the 1900nm
feature (Fig. 6.10). Illite-dominant illite-smectite samples are identified using the following
absorption feature characteristics:

𝐷𝑒𝑝𝑡ℎ > 𝐷𝑒𝑝𝑡ℎ

𝐷𝑒𝑝𝑡ℎ > 𝐷𝑒𝑝𝑡ℎ

𝐷𝑒𝑝𝑡ℎ > 200

Smectite dominant illite-smectite samples are characterised by:

𝐷𝑒𝑝𝑡ℎ < 𝐷𝑒𝑝𝑡ℎ

𝐷𝑒𝑝𝑡ℎ > 𝐷𝑒𝑝𝑡ℎ

𝐷𝑒𝑝𝑡ℎ > 100

195
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.10: Spectrum of an illite-smectite mixture in which smectite is the dominant clay. Dashed red lines
indicate the depth and position of the absorption features. Refer to text for further explanation of the
characterisation of illite-smectite mixtures.

6.3 Results
Based on core log scanning conducted by SpecCam, the total content of clays is evaluated
semi-quantitatively by analysis of the frequency of clay mineral identification by line scans
spaced at ~4cm intervals. These results show that the periphery of the host quartz biotite
quartz diorite intrusion is the locus of high clay mineral content (Fig. 6.11 and Fig. 6.12).

196
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.11: Cross section across 3810m level of the Cerro Corona intrusion with the contours of the percentage
of total clay content. Dots represent sample locations.

197
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.12: Cross section along 76300 across the deposit contoured for total clay mineral content based on
SpecCam spectral logging.

198
Chapter Six: Infrared spectroscopy of clay minerals

6.3.1 Major minerals in infrared spectroscopy


Approximately 43% of spectra were unidentified using a search and match function in excel
in which the identification of absorption features as described in the methodology identify
the presence of the mineral. The lack of identification is largely attributed to the lack of
access to specialised software (such as The Spectral Geologist). Access to this software was
denied as the cost of this software was beyond the budget of this project (as determined by
the supervisors).

The dominant clays in samples where only one mineral was identified were smectite
(15%), illite (6%) and kaolin group (5%) (Fig. 6.13). Perhaps unsurprisingly, a large
proportion of the spectra were classified as mixtures: smectite-illite (27%), kaolinite-illite
(2%) and kaolinite-illite-smectite (2%) and kaolinite-smectite (<1%) (Fig. 6.13).

Figure 6.13: Pie chart showing the percentages of dominant mineral groups identified in samples from Cerro
Corona using infrared spectroscopy.

6.3.2 Mineral variance

6.3.2.1 Kaolin group

Abundance

The kaolin group and mixtures of kaolinite and clays contribute ~ 16% of the identified
clay minerals. The dominant kaolin group mineral is kaolinite, with lesser amounts of
halloysite and dickite (Fig. 6.14-A). Within the kaolinite group mixtures (Fig. 6.14-B),

199
Chapter Six: Infrared spectroscopy of clay minerals

kaolinite-illite is the most dominant mixture (55%) with lesser amounts of kaolinite-illite-
smectite (43%) and minor kaolinite-smectite (2%) (Fig. 6.14).

Figure 6.14: Pie charts of the abundance of the kaolinite group minerals (A) and their mixtures (B).

200
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.15: Map of distribution of kaolinite, based on frequency of kaolinite identification

201
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.16:The kaolinite content, based on semi-quantitative methods by SpecCam.

202
Chapter Six: Infrared spectroscopy of clay minerals

Crystallinity

The crystallinity index for dickite is asymmetrically centred on 0.6 (Fig. 6.17). There is a
normal distribution of crystallinity in the kaolin subgroup (i.e. kaolinite and halloysite),
centred on 0.9. As expected, the halloysite subgroup overlaps with the least ordered
kaolinite (Fig. 6.18).

Figure 6.17: Histogram of the crystallinity index for dickite (a) and for kaolin group minerals (b). There is a slight
asymmetry toward a higher ratio in dickite-dominant spectra and a more normal distribution for the rest of the
kaolin group. “All” means all samples with kaolinite and dickite.

The crystalline kaolinites appear to predominate at depth, with lower kaolinite crystallinity
at and near the current surface (Fig. 6.18 and Fig. 6.19).

203
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.18:Plan section across 3810 contoured with the kaolinite crystallinity.

204
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.19: Cross section across 763300 with contours of kaolinite crystallinity index.

The bimodal distribution of vOHinner surface+δAlAlOH versus vOHinner inner hydroxyl+δAlAlOH (Fig. 6.20.a)
is due to dickite having a higher ratio than kaolinite-halloysite subgroup (Figs. 6.20 a and

205
Chapter Six: Infrared spectroscopy of clay minerals

b). Halloysite and kaolinite may not be a truly separate group, as their 2180/2210 ratio
show a normal distribution (Fig. 6.20).

Figure 6.20: Histogram showing the 2180/2210 absorption band ratio. A) The kaolin group minerals have a
bimodal distribution (grey), which can be separated into dickite (orange) and kaolinite (maroon). B) The
subgroup of the kaolin into kaolinite (maroon) and halloysite (purple) appears artificial and does not distinguish
two mineralogical groups.

Hydration

The 1900 nm H2O feature appears predominantly centred on 1912 nm, with overlapping
peak at ~1920nm (Fig. 6.21).

Figure 6.21: The location of the 1900 nm absorption feature. There are possibly two overlapping peaks in this
region, the major adsorption feature at ~1912 nm and a secondary feature at ~1920 nm.
206
Chapter Six: Infrared spectroscopy of clay minerals

The assessment of dickite crystallinity (i.e. the ratio between the ~1380 nm and 1415 nm
absorption feature) appears to have a positive correlation with the 2180/2210 ratio (Fig.
6.22). The most crystalline dickite appears to have high ratios in both and appear to have a
normal distribution in these ratios (Fig. 6.23).

Figure 6.22: Plot of the absorption depth ratios between dickite doublets. The dickite crystallinity index is the
ratio of the depth of 1380 nm feature versus the depth of the 1415 nm feature, whereas D2180/D2210 is the ratio
of the depth of the 2180 nm feature versus the depth of the 2210 nm feature. The most crystalline dickite samples
have the highest values in both doublets.

Figure 6.23: Histogram of the distribution of dickite crystallinity index (~1390/1415 absorption depths) and
~2170/2210 nm absorption depths.

207
Chapter Six: Infrared spectroscopy of clay minerals

6.3.2.2 Illite group

Abundance

Illite minerals and mixtures constitute about 65% of the identified clay minerals. Of this
proportion, pure illite makes up ~15%. By far the largest proportion is the smectite-illite
mixture (~71%, Fig. 6.24). Other mixtures include kaolin-illite (~7%) and kaolinite-illite-
smectite (~5%). Well crystallised micas are included in this group and make up a minor
proportion (~2%).

Figure 6.24: Relative proportions of illite-bearing mineral assemblages. The dominant mixture observed is
smectite-illite.

208
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.25: Plan section across 3810 contoured for the illite content.

209
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.26: Cross section across 763300 with contours of calculated illite content.

The position of the vOH+δXAlOH absorption for illite is bimodal and is centred on 2198 and
2208 nm (Fig. 6.27). These two illite groups are split into <2204 and >2204 nm illite for the

210
Chapter Six: Infrared spectroscopy of clay minerals

following discussion (Fig. 6.27). Kaolinite-illite and kaolinite-illite-smectite mixtures show


a single absorption mode, normally distributed around 2208 nm.

Figure 6.27: Histogram of illite end-members, based on the 2200 nm absorption feature position. Illite with
absorption feature <2204 nm is plotted on the secondary axis for clarity.

Crystallinity

The crystallinity index of illite appears to be broadly bimodal (Fig. 6.28), with a more
crystalline illite (centred on 4) and a lower crystallinity end-member, centred on 7. Clay
mixtures have a lower illite crystallinity index than illite, with illite-smectite mixtures
asymmetrically centred on 4 and kaolinite-illite mixtures centred on 6 and 8. Kaolinite-
illite-smectite mixtures appear to have the broadest range of crystallinity.

211
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.28: Histogram of the ratio of the depths of the 2200 and 2340 nm features as a proxy for crystallinity.
The smaller the value, the higher the “crystallinity”.

Illite with the lower wavelength vOH+δXAlOH absorption feature is highly crystalline (Fig.
6.29), whereas the illite with a higher wavelength vOH+δXAlOH absorption has a more variable
crystallinity. Mixtures of illite-smectite and kaolinite-illite display a broad range of
crystallinity values (Fig. 6.29).

212
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.29: Plot of illite “crystallinity” index versus 2200 nm wavelength position for kaolinite-illite-smectite,
illite-smectite and illite clays. A) The red arrow indicates increasing “crystallinity” with decreasing absorption
position within the illite mixtures. B) The subcategories of pure illite show that illite with < 2204 nm wavelength
have a consistently higher crystallinity (i.e. lower “crystallinity” index) than those with absorption feature >2204
nm, which have a variable “crystallinity”.

213
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.30: Plan section of illite crystallinity index.

214
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.31: Cross section of illite crystallinity index show that more crystalline illite is centred in the north and
central parts of the host intrusion

215
Chapter Six: Infrared spectroscopy of clay minerals

6.3.2.3 Smectite group

Abundance

The smectite group minerals contribute to about 76% of the identified clays at Cerro
Corona (Fig. 6.13). Of this proportion, ~35% are identified pure smectite, ~61% are illite-
smectite mixture, ~4% are kaolinite-illite-smectite and <1% are kaolinite-smectite mixture
(Fig. 6.32). The dominant smectite is montmorillonite, with lesser amounts of
montmorillonite-beidellite and trace amounts of pure beidellite (Fig. 6.32).

Figure 6.32: (A) Identified smectite mineral proportions and (B) the most commonly identified pure smectite.

216
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.33:Plan section of overall smectite content across Cerro Corona.

217
Chapter Six: Infrared spectroscopy of clay minerals

Figure 6.34:Cross section across 763300 with contours of overall smectite content.

At Cerro Corona, the smectite mineral group appears to be dominated by the


montmorillonite-beidellite dioctahedral endmembers, as indicated by the vOH+δAlAlOH
combination band in the 2200 nm region. The distinction between beidellite and

218
Chapter Six: Infrared spectroscopy of clay minerals

montmorillonite in the SWIR has not been convincingly demonstrated. However, the use of
the position of the vOH+δAlAlOH absorption band may be used as a proxy for the Al2O3 content
of the smectite (Post and Noble 1993), thus beidellitic (Al2O3>30wt.%) smectite are
distinguished by lower wavelength positions of the 2200 nm absorption feature than that
of montmorillonitic smectite (Al2O3~20-25wt. %).

6.3.3 Future work


Control of the humidity during data collection may aid in the distinction between interlayer
cations within the smectites (Bishop et al. 1994). The presence of interlayer cations is
known to have an impact on the absorptive ability (and expandability) of the smectite (from
most absorptive Mg>Ca>Fe>Na to least absorptive). This knowledge would be beneficial to
metallurgy and tailing construction.

Improvements to the fitting algorithm of the absorption features in SWIR spectra would
improve the detection (such as the fundamental stretching and bending mode of the
absorbed water molecule located in the ~1960 nm region). This improvement may also aid
in the determination of more subtle absorption features and their shifts. Thus, future
recommendations include an improvement in the IR spectral processing software. Future
research strongly depends on access to reliable software packages which would enable
feature positions and integrate mineral databases.

Strong debates still exist about the (meta) stability of natural clay mineral assemblages (Aja
and Rosenberg 1992; Essene and Peacor 1997). The primary cause for the debate is the
sluggish reaction rates involved in the clay mineral assemblage. Clay minerals do form in a
variety of environments and the attainment of assemblage stability is difficult to
demonstrate, as the clay minerals are <2μm, thus textural relationships (and thus stability)
of clay minerals is challenging to observe. This is an important note, as the identified clay
minerals at Cerro Corona can change over the scale of meters, centimetres and millimetres.

6.4 References
Aja, S. U. and Rosenberg, P. E. (1992), The thermodynamic status of compositionally
variable clay minerals, Clays and Clay Mineralogy, 40 (3), 292-299.

Balan, E., Delattre, S., Guillaumet, M. and Salje, E. K. (2010) Low-temperature infrared
spectroscopic study of OH-stretching modes in kaolinite and dickite. American
Mineralogist. 95 (8-9), 1257-1266.

219
Chapter Six: Infrared spectroscopy of clay minerals

Barrios, J., Plançon, A., Cruz, M. and Tchoubar, C. (1977) Qualitative and quantitative study
of stacking faults in a hydrazine treated kaolinite-relationship with the infrared
spectra. Clays and Clay Minerals. 25 (6), 422-429.

Benco, L., Tunega, D., Hafner, J. and Lischka, H. (2001) Ab initio density functional theory
applied to the structure and proton dynamics of clays. Chemical Physics Letters. 333 (6),
479-484.

Bishop, J. L., Pieters, C. M. and Edwards, J. O. (1994) Infrared spectroscopic analyses on the
nature of water in montmorillonite. Clays and Clay Minerals. 42 (6), 702-716.

Bishop, J.L., Madejova, J., Komadel, P. and Fröschl, H. (2002) The influence of structural Fe,
Al and Mg on the infrared OH- bands in spectra of dioctahedral smectites. Clays and Clay
Minerals. 37 (4), 607-616.

Brindley, G. W., Kao, C., Harrison, J. L., Lipsicas, M. and Raythatha, R. (1986) Relation
between structural disorder and other characteristics of kaolinites and dickites. Clays and
Clay Minerals. 34 (3), 239-249.

Cariati, F., Erre, L., Micera, G., Piu, P. and Gessa, C. (1983) Polarization of water molecules in
phyllosilicates in relation to exchange cations as studied by near infrared
spectroscopy. Clays and Clay Minerals. 31 (2), 155-157.

Clark, R. N. and Roush, T. L. (1984) Reflectance spectroscopy: Quantitative analysis


techniques for remote sensing applications. Journal of Geophysical Research: Solid Earth. 89
(B7), 6329-6340.

Clark, R. N., King, T. V. V., Klejwa, M., Swayze, G. A. and Vergo, N. (1990) High spectral
resolution reflectance spectroscopy of minerals. Journal of Geophysical Research: Solid
Earth. 95 (B8), 12653-12680.

Crowley, J. K. and Vergo, N. (1988) Near-infra-red reflectance spectra of mixtures of kaolin-


group minerals: use in clay mineral studies. Clays and Clay Minerals. 36 (4), 310.

Delineau, T., Allard, T., Muller, J. P., Barges, O., Yvon, J. and Cases, J. M. (1994) FTIR
reflectance vs. EPR studies of structural iron in kaolinites. Clays and Clay Minerals. 42 (3),
308-320.

Duke, E. F. and Lewis, R., S. (2010) Near infrared spectra of white mica in the Belt
Supergroup and implications for metamorphism. American Mineralogist. 95 (7), 908-920.

220
Chapter Six: Infrared spectroscopy of clay minerals

Essene, E. J. and Peacor, D. R. (1997) Illite and smectite: metastable, stable or unstable?
Further Discussion and a Correction. Clays and Clay Minerals. 45 (1), 116.

Frost, R. L. and Johansson, U. (1998) Combination bands in the infrared spectroscopy of


kaolins--a DRIFT spectroscopic study. Clays and Clay Minerals. 46 (4), 466-477.

Geysermans, P. and Noguera, C. (2009) Advances in atomistic simulations of mineral


surfaces. Journal of Materials Chemistry. 19 (42), 7807-7821.

Giese, R. F. and Datta, P. (1973) Hydroxyl orientation in kaolinite, dickite, and


nacrite. American Mineralogist. 58 (5-6), 471-479.

Green, A. and Craig, M. (1985) Analysis of aircraft spectrometer data with logarithmic
residuals. JPL Proceedings of the Airborne Imaging Spectrometer Data Analysis
Workshop. United States.

Grim, R. E. (1962) Clay Mineralogy: The clay mineral composition of soils and clays is
providing an understanding of their properties. Science. 135 (3507), 890-898.

Guggenheim, S., Adams, J., Bain, D., Bergaya, F., Brigatti, M. F., Drits, V., Formoso, M. L., Galán,
E., Kogure, T. and Stanjek, H. (2006) Summary of recommendations of nomenclature
committees relevant to clay mineralogy: report of the Association Internationale pour
l’Etude des Argiles (AIPEA) Nomenclature Committee for 2006. Clay Minerals. 41 (4), 863-
877.

Hunt, G. R. and Salisbury, J. W. (1970) Visible and near-infrared spectra of minerals and
rocks: I silicate minerals. Modern Geology. (1), 283-300.

Hunt, G. (1977) Spectral signatures of particulate minerals in the visable and near
infrared. Geophysics. 42 (3), 501-513.

Iwasaki, T. and Watanabe, T. (1988) Distribution of Ca and Na ions in dioctahedral


smectites and interstratified dioctahedral mica/smectites. Clays and Clay Minerals. 36 (1),
73-82.

Joussein, E., Petit, S., Churchman, J., Theng, B., Righi, D. and Delvaux, B. (2005) Halloysite
clay minerals–a review. Clay Minerals. 40 (4), 383-426.

Kerr, A., Rafuse, H., Sparkes, G., Hinchey, J. and Sandeman, H. (2011) Visible/infrared
spectroscopy (VIRS) as a research tool in economic geology: Background and pilot studies
from Newfoundland and Labrador. Geological Survey, Report. 11, 145-166.

221
Chapter Six: Infrared spectroscopy of clay minerals

Madejova, J. and Komadel, P. (2001) Baseline studies of the clay minerals society source
clays: infrared methods. Clays and Clay Minerals. 49 (5), 410-432.

Parker, T. W. (1969) A classification of kaolinites by infrared spectroscopy. Clay Minerals. 8


(2), 135-141.

Petit, S., Madejová, J., Decarreau, A. and Martin, F. (1999) Characterization of octahedral
substitutions in kaolinites using near infrared spectroscopy. Clays and Clay Minerals. 47 (1),
103-108.

Plançon, A. and Tchoubar, C. (1977) Determination of structural defects in phyllosilicates


by X-ray powder diffraction; II, Nature and proportion of defects in natural kaolinites. Clays
and Clay Minerals. 25 (6), 436-450.

Post, J. L. and Noble, P. N. (1993) The near-infrared combination band frequencies of


dioctahedral smectites, micas, and illites. Clays and Clay Minerals. 41639-639.

Sposito, G., Prost, R. and Gaultier, J. (1983) Infrared spectroscopic study of adsorbed water
on reduced-charge Na/Li-montmorillonites. Clays and Clay Minerals. 31 (1), 9.

Zhang, G., Wasyliuk, K. and Pan, Y. (2001) The characterization and quantitative analysis of
clay minerals in the Athabasca Basin, Saskatchewan: Application of shortwave infrared
reflectance spectroscopy. The Canadian Mineralogist. 39 (5), 1347-1363.

222
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

7 Chapter Seven: Oxygen and hydrogen isotope


analyses of argillic alteration

7.1 Introduction
In this chapter, the primary objective is to evaluate the proportions of magmatic and
meteoric waters involved in the formation of argillic alteration. Argillic alteration typically
forms at shallow levels in porphyry systems and is the result of lower temperature
hydrogen metasomatism of silicate rocks (Giggenbach 1992). This alteration can be formed
during both hypogene and supergene processes and thus can span the duration of porphyry
formation (Chapter 2, section 2.1).

7.2 Methods
The primary isotopic composition of the Cerro Corona biotite quartz diorites, prior to
alteration, is unknown. In inferring the most probably bulk isotopic composition of the host
intrusion, the dominant igneous minerals are considered. The average δD values of
magmatic hornblende is -70 ‰ (Sheppard and Gustafson 1976), average δD of biotite
associated with porphyry deposits typically ranger between -71 to -79 ‰ (Sheppard et al.
1971). The range of δD values of shallow-level hydrous dacites within the nearby
Yanacocha district is between -55 and -75‰ (Chambefort et al. 2013). It thus seems most
likely that the Cerro Corona biotite quartz diorite had a δD of ~ -70‰. The δ18O value of
the Cerro Corona diorite is inferred to mimic that of the average δ18O of plagioclase (~ 7‰)
in porphyry copper deposits (Sheppard and Gustafson 1976). Thus, the inferred
composition of the Cerro Corona biotite quartz diorite is -70‰ and 7‰, for δD and δ18O
respectively, which will be used in modelling the impact of alteration in this chapter.

7.2.1 Clay mineral separation and X-ray Diffraction analysis


Prior to isotope study, 57 clay mineral separates were produced from whole rock drill core
samples (Appendix 6). The predominant clay mineral assemblages in these samples had
been previously identified in the field by hand-held SWIR spectrometer (TerraSpec™)
(Chapter 6).

X-ray diffraction produces a diffraction pattern based on the atomic arrangement of the
crystalline material, based on Bragg’s Law:

𝑛λ = 2𝑑 sin θ
223
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Where λ is the wavelength of the incident x-ray, d is the spacing between the atomic planes,
and θ is the angle between the incident ray and the scattering planes. As the wavelength for
the Cu K-alpha incident wave is known and the θ angle is controlled during the run, the
spacing between the atomic planes can be determined. This spacing between atomic layers
is used to distinguish between clay minerals by comparison of crystallographic changes
from air-dried, glycol and heating treatments (Moore and Reynolds 1989).

6.4.1.1 Preparation of clay samples

The liberation of clay minerals involved jaw crushing of the whole rock sample, placing the
crushed sample into an ultrasound bath and several stages of centrifuging to separate out
the 1-2 µm and <1 µm size fractions (Fig. 7.1). In the jaw crusher, whole rock samples were
crushed to approximately <2 cm. Jaw crushed samples were placed in a large beaker filled
with de-ionised Milli-Q water in an ultrasound bath for approximately 12 hours. This
liberates the clay minerals from the remaining particles and places them in suspension. The
suspension was pipetted off and centrifuged in an Eppendorf 5810R Centrifuge at 1000
rpm for 4 minutes. Clay-sized particles (particles less than two microns) remained in
suspension. The second suspension was run at 2200 rpm for 3 minutes which separated
out particles greater than one micron and kept particles less than one micron in suspension.
The one to two micron separate is given a syntax “-A”. The third suspension, with particles
less than one micron, was centrifuged at 4000 rpm for 20 minutes and given a syntax “-F”.
This process was repeated until sufficient separate was generated for XRD and isotope
analysis.

Figure 7.1: Schematic diagram illustrating how clay minerals were separated from whole rock samples.

224
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

The suspensions are pipetted onto porcelain tiles and air dried overnight. This causes the
clay minerals to preferentially orientate, such that the basal 001 plane aligns parallel to the
porcelain surface. This preferential orientation produces a more definitive XRD pattern,
and similar alignment aids in a more realistic quantitative analysis of clay mineral
proportions (Moore and Reynolds 1989).

6.4.1.2 Glycol treatment

Following XRD analysis of air-dried orientated samples, liquid ethylene glycol was pipetted
onto the orientated sample. This glycol treatment results in an expansion of the smectite
lattice, detectable in the resultant XRD pattern. Interlayered illite-smectites causes a shift
and broadening of the illite-smectite peaks for approximately 3 hours, until the clay dries
out again. When no smectite is present, the ethylene glycol has no effect (Moore and
Reynolds 1989). Glycolated XRD patterns were labelled JLx.xxxG.

6.4.1.3 Heat treatment

Samples were heated in a furnace to 400 °C for 1 hour. At this temperature, the smectites
lattice collapses and the diffraction pattern resembles pure illite with the 001 reflection at
about 10 Å. Samples are labelled JLx.xxxH

Samples are heated to 550 °C for 1 hour, which results in the kaolin phases to become
amorphous and thus their XRD pattern disappears. This allows for distinction between
chlorite and the kaolin group (Moore and Reynolds 1989). Samples are labelled JLx.xxxI.

6.4.1.4 X-Ray diffraction analysis

X-ray diffraction measurements were made on a Philips PW1820 operated at 40 kV and 40


mA using Cu K-alpha radiation at angles between 2.5 to 40° with step size 0.015° at 2
seconds per step.

The response of clay minerals separates to the variety of treatments is indicative of the clay
mineralogy (Table 7.1). Thus, the identification of clay minerals can confidently be
established.

225
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Table 7.1: Summary of treatments and their effects on the basal 001 reflection of clay minerals. *peak intensities
change (001 increase, higher-order peaks 002, 003, 004 decrease) **better comparison of 001 reflections of
Kaolinite ~14 Å, Chlorite ~14 Å

Mineral name Air dried Glycolated Heated (400 °C) Heated (500 °C)
Smectite ~15 Å ~14-17 Å 10 Å 10 Å
Illite ~10 Å Unchanged Unchanged Unchanged
Kaolinite ~7 Å** Unchanged Unchanged Disappear
Chlorite ~7 Å** Unchanged Unchanged Unchanged*

This confirmed that the sub-1 micron fractions generally contained less quartz then their
sub-2 micron counterparts. Due to the limited amounts of sub-1 micron material, there can
be a contribution to the XRD pattern of the porcelain tile. However, this can be easily
identified and eliminated during phase identification.

6.4.1.5 Clay mineral quantification

Mineral intensity factors and sum to 100 % utilised in conjunction to quantify the clay
mineral proportions (Kahle et al. 2002). In this approach, the weight percent of the clay
fraction (𝑊 ) is calculated:

𝐼
𝑀𝐼𝐹
𝑊 =
∑ 𝐼 𝑀𝐼𝐹

Where the subscript 𝛼 denotes the mineral and n denotes any mineral phase in the mixture,
I is the area of the peak and MIF is the mineral intensity (area) factor for that phase. The
mineral intensity factors for each mineral phase is calculated from known mineral mixtures
of binary mixtures of these clay minerals.

The quantification is based on the “sum to 100 %” approach whereby the area of selected
peaks is normalised to sum to 100 %. The peaks areas used for smectite, illite, chlorite and
kaolinite are their 003, 002, 003 and 002 peaks respectively (Fig. 7.2).

226
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.2: Characteristic X-ray diffraction pattern of dominant clay minerals, showing location of diagnostic
peak positions.

7.2.2 Sample homogeneity


The Quanta FEI 650 was utilised to confirm sample homogeneity in clay mineral separates.
This technique provided higher levels of magnification than conventional SEM, and semi-
quantitative element distribution maps confirmed clay mineral separates were indeed
homogeneous clays (Fig. 7.3).

A B

Figure 7.3: Back scattered electron images of clay


C
mineral separates taken under the FEI Quanta 650
scanning electron microscope. These images show
sample homogeneity in: A) Kaolinite-dominant
sample (88 % kaolinite) JL2.216A. Field of view: 35
μm; B) Kaolinite-dominant sample (97 % kaolinite)
JL2.230F. Field of view: 35 μm, and; C) Illite-
dominant sample (81 % illite) JL2.98F. Field of view:
60 μm

227
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

7.2.3 Single mineral dehydroxylation temperatures


Thermogravimetric Analysis (TA) of clay mineral separates was done to identify the
temperature at which dehydration (i.e. loss of adsorbed water) and dehydroxylation (i.e.
removal of –OH hydroxyl) occurs. Knowledge of the temperatures at which dehydration
and dehydroxylation occurs allows for temperature targeting of the dehydroxylation for
accurate isotope analyses.

7-15 mg of a clay sample (<1µm particle size) was placed into the sample crucible of the TA
instrument model SDT Q600. The sample is heated at 10 °C/min from room temperature
(approximately 20 °C) to 950 °C. A nitrogen purge rate of 100ml/min was used to maintain
a consistently dry environment.

Illite shows the smallest bulk loss with general mass loss occurring at temperature <180 °C
and between 400-800 °C (Fig. 7.4, orange). Kaolinite shows a larger bulk mass loss (Fig. 7.4,
blue), the majority of which is lost between 400-700 °C. The largest bulk loss is seen in
smectite (Fig. 7.4, green), the majority of the weight loss occurs <200 °C, with a smaller
mass loss between 500-700 °C. These results are comparable to baseline thermal analysis
of standard clay minerals (Guggenheim and van Groos 2001). This analysis demonstrates
that most of the adsorbed water is removed by ~200 °C and dehydroxylation occurs
between 400-800 °C.

Figure. 7.4: Thermal analysis of predominant clay mineral separates of kaolinite (blue), illite (orange) and
smectite (green) with the specific temperatures at which dehydration (lower temperatures) and dehydroxylation
indicated.

228
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

7.2.4 Oxygen isotope analysis

Pre-analysis treatments

Three treatments were applied prior to measuring the oxygen isotope composition of both
the structural oxygen and the hydroxyl in the clay mineral separates:

1. No treatment (“Pumped”) – the clay mineral separate was not heated (n=38);
2. Mild heating (“Heated”) – the clay mineral separate was heated to ~75 °C for >24
hours to remove adsorbed atmospheric water, without de-hydroxylating the clay
(n=13);
3. Heating under vacuum (“Vacuum furnace”) - the clay mineral separate was pumped
down to 1.02x10-5Pa at 200 °C for >12 hours (n=40).

To measure the structural oxygen, as opposed to both structural and hydroxyl oxygen, a
fourth pre-heating treatment was adopted. The clay mineral separate was heated to >1000
C under high vacuum to remove the hydroxyl layer (n=13, Fig. 7.5). The remaining sample
was then released into a dry atmosphere, by exposure to liquid nitrogen. The sample was
then loaded into the sample holder and analysed using the standard laser fluorination
method (see below). This method was used as opposed to the partial fluorination technique
adopted by Hamza and Epstein (1980) as the partial fluorination technique is shown to be
unreproducible (Girard and Savin 1996).

Figure 7.5: Clay mineral proportions of samples selected for OH-removed oxygen isotope analyses.

229
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Following the pre-treatment, oxygen was extracted from the clays. The procedure adopted
for the extraction of oxygen from phyllosilicates mimics that of the laser fluorination
method for the liberation of oxygen from silicates (Sharp 1990, 1992), with a few minor
modifications as outlined below.

Two samples were placed in a nickel sample holder for each run, evacuated, and
approximately 0.25 bar (2.5x105 Pa) of chlorine trifluoride (ClF3) vapour was then added
to the sample chamber. Although ClF3 is less stable and the vapour pressure is higher than
BrF5, it was preferred over BrF5 and F2 due to 3 main advantages: it freezes completely in
liquid nitrogen; leaves the system within 30 minutes of pumping (i.e. less sticky) and the
CO2 produced is halogen free (Borthwick and Harmon 1982). The clay separate was
analysed first, followed by a standard. The samples were heated with a CO2 laser, which
was admitted through a BaF2 window (Akagi et al. 1993). This window is transparent to
both infrared and visible radiation and does not react with the reagent (Sharp 1990).

The ClF3 reacts with the clay minerals and liberates oxygen, as illustrated by the reactions:

dickite and kaolinite:

6𝐴𝑙 𝑆𝑖 𝑂 (𝑂𝐻) + 28𝐶𝑙𝐹 → 12𝐴𝑙𝐹 + 12𝑆𝑖𝐹 + 14𝐶𝑙 + 27𝑂 + 12𝐻

illite:

13𝐾 . 𝐴𝑙 (𝐴𝑙 . 𝑆𝑖 . 𝑂 )(𝑂𝐻) + 146.6𝐶𝑙𝐹


→ 13𝐾𝐹 + 53𝐴𝑙𝐹 + 67𝑆𝑖𝐹 + 73.3𝐶𝑙 + 120𝑂 + 20𝐻

smectite (Na-Al end member):

𝑁𝑎 . 𝐴𝑙 (𝑆𝑖 𝑂 )(𝑂𝐻) ∗ 4𝐻 𝑂 + 7.4𝐶𝑙𝐹


→ 0.3𝑁𝑎𝐹 + 2𝐴𝑙𝐹 + 4𝑆𝑖𝐹 + 3.7𝐶𝑙 + 8𝑂 + 4𝐻

Liberated oxygen was passed successively over two nitrogen traps and through a hot
mercury fluorine-getter. This removes any halogen contaminants which may be present.
The oxygen was converted to CO2 by reaction with hot carbon rod, heated by conduction. A
liquid nitrogen trap causes deposition of CO2 with the non-condensable fraction discarded.
The CO2 volume was measured and passed into a sample tube for mass spectrometry using
a VG SIRA II dual inlet mass spectrometer

Three in-house and two international standards are used for data quality control purposes
(Table 7.2). The standards were selected on the basis that they have similar isotope
compositions to the samples. The in-house standards are pure, homogeneous quartz (SES
230
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

and TOR1) and olivine (SCXO). The international standards are homogeneous garnet (GP
147) and the Gore Mountain garnet standard (UGW-2; Valley et al. 1995). A standard was
run after each clay analysis and an end-of-day calibration was calculated utilising the raw
and true standard data.

Isotope compositions are reported as per mil (‰) deviations of the 18O/16O and D/H ratios
relative to Vienna Standard Mean Ocean Water (V-SMOW), expressed using standard δ
notation (Craig 1961):

𝑅
δ= − 1 ∗ 1000
𝑅

Thus, a δ value of +5 ‰ is 5 parts per thousand (or 0.5 %) more enriched in 18O or D than
V-SMOW. The analytical error is ± 0.3 ‰ based on one standard deviation of the analysed
standards (Table 7.2).

Table 7.2 Standards used for oxygen isotope analyses. See text for descriptions of standards.

Standard name In-house/International δ18O SMOW ( ‰)


GP 147 International 7.25
TOR 1 In-house 9.60
SCXO In-house 5.20
SES In-house 10.20
UGW2 International 5.80

The standard deviation on the measurement of the international standards, including


fluorination and mass spectrometry, was 0.3 ‰ for GP147 and 0.2 ‰ for UWG2. The in-
house standards have a standard deviation of 0.5 ‰ for SES, 0.37 ‰ for TOR and 0.15 ‰
for SCXO. Thus, the error of reproducibility was consistently equal to or better than ±0.5
‰: the average standard error of all analysed standards (n=115) was ±0.3 ‰ (Fig. 7.6).

231
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.6: Measured versus true δ18O of all analysed standards (n=115) throughout isotope study. The
coefficient of determination is based on all standards.

Duplicates of the same natural clay samples measured with the various pre-treatment steps
display a variable standard deviation. The non-treated separates have a standard deviation
(σ) of 0.75 ‰ (n=7). The mild heating pre-treatment showed little improvement in this
standard deviation (σ=0.83 ‰, n=3) whereas the vacuum furnace-treated clay separates
showed a significant improvement (σ=0.20 ‰, n=4), similar to the standard deviation of
the standards. Thus, these experiments show that the most precise method for determining
the structural and/or the hydroxyl oxygen isotope values for clays is by pre-heating the
sample in a vacuum furnace to ~200 °C for an extended period, thus removing the adsorbed
water component. This is in accord with previous studies (Savin and Epstein 1970;
Lawrence and Taylor 1972; Sheppard and Gilg 1996; Cuadros et al. 1994).

7.2.5 Hydrogen isotope analysis


A total of 44 samples were selected for hydrogen isotope analyses, derived from a range of
drillholes, depths and clay mineral fractions (Appendix 7).

Hydrogen was removed from the clay mineral separates following a method modified from
Godfrey (1962). All labile adsorbed and interlayer water was removed from the separate
by heating the sample in a thoroughly cleaned and degassed Pt crucible. This was
232
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

accomplished by heating the sample within the Pt crucible while under vacuum for
approximately 12 hours. The glass vessel and crucible were both heated to remove any
water that may have condensed and thus could potentially contaminate the sample
(Suzuoki and Epstein 1976).

Subsequently, the crucible was incrementally heated to >1000 °C using an induction coil
and then left at temperature for 40 minutes to ensure complete dehydration and liberation
of gas. During heating, a liquid nitrogen trap was used to condense H2O gas. After holding
the sample at >1000 °C for 15 minutes, the liquid nitrogen trap was replaced by an acetone-
dry ice “slush” trap, which keeps H2O frozen and the volume of non-condensable gases was
measured.

The trapped H2O was then reduced by reaction in a hot Cr furnace (Donnelly et al. 2001),
without isotopic fractionation, following the equation:

2𝐶𝑟 + 3𝐻 𝑂 → 𝐶𝑟 𝑂 + 3𝐻

The H2 gas was then moved by a Toepler pump into a sample tube and analysed using a VG
Optima dual inlet mass spectrometer.

Three in-house standards are used for calibration and data quality control purposes (Table
7.3). These standards were selected as their isotopic values are similar to the sample
values. An international standard of Greenland Ice Sheet Precipitation (GISP) is utilised as
a well-established reference (Hut 1987). A standard was run approximately every week
and calibration was calculated utilising the values of these standards. This calibration was
used to calculate the true isotopic compositions of the samples, relative to V-SMOW from
the raw mass spectrometer values, as well as the overall analytical error for the samples.

Table 7.3: Standards used for hydrogen isotope analyses.

Standard name In-house/International δDSMOW ( ‰)


GISP International -189.5
SNOWMELT In-house -123

SEAWATER In-house -6

The standard error calculated was ±2 ‰, based on the standard deviation of all measured
standards, including mass spectrometry and processing through the vacuum line.

233
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

7.2.6 Local meteoric waters


For H isotope analysis, sample and standard waters were directly injected into a chromium
furnace at 800°C (Donnelly et al. 2001), with the evolved H2 gas analysed on-line on a VG
Optima mass spectrometer. Replicate analyses of water standards (international standards
V-SMOW and GISP, and internal standards) gave a reproducibility of ±3‰.

For O isotope analysis of water, 200 µl of the water samples analysed by standard
techniques on a Thermo Scientific Delta V mass spectrometer, set at 25°C. Each sample was
over-gassed with a 1% CO2-in-He mixture for five minutes and left to equilibrate for a
further 24 hours before mass spectrometric analysis. Oxygen isotope data were then
produced using the method established by Nelson (2000). As with H isotopes, the data are
reported as per mille variations from the V-SMOW standard. During the analyses of these
samples, reproducibility of the data, based on within-run repeat analyses of standards, was
around ±0.3‰.

7.3 Results

7.3.1 Bulk mineral oxygen isotope results


The bulk clay mineral oxygen isotope compositions appear to have a bimodal distribution
centred around +4 ‰ and +5.5 ‰ (Fig. 7.7 and Appendix 7), with a few outlying samples.
Samples are sub-divided into their predominant clay minerals and plotted against present
day depth (Fig. 7.8).

Figure 7.7: Histogram of bulk mineral δ¹⁸O values (i.e. structural and hydroxyl oxygen).
234
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Smectite-dominant clay mineral samples have δ18O values ranging from+1.2 ‰ to +11.8
‰ (Fig. 7.8). This large spread in composition is reflected in all samples, regardless of pre-
treatment. Illite-dominant clays appear to have a smaller range of δ18O values ranging
between +2.2 to +9.6 ‰, with the samples that are pre-heated under vacuum having a
smaller range of δ18O values (+2.9 to +6.8 ‰, Fig. 7.8). The δ18O values for kaolinite-
dominant mineral separates range from +1.0 ‰ to +6.4 ‰ if the structural δ18O values are
excluded.

Figure 7.8 Depth versus bulk oxygen isotope values for A) kaolinite; B) illite, and; C) smectite minerals. Note these
are samples where the minerals are >50 wt. %.

7.3.2 De-hydroxylated oxygen isotope results


The δ18O values of the dehydroxylated (i.e. hydroxyl removed) residue have range from
+4.0 ‰ to +11.9 ‰ (Fig. 7.9) and have higher δ18O values than the bulk samples (Fig. 7.7
235
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

and Fig. 7.8) indicating that isotopically light 16O is preferentially incorporated in the
hydroxyl site.

Figure 7.9: Histogram of the δ18O values of dehydroxylated clay mineral separates.

O’Neil and Pickthron (1988) noted that the isotopic composition at different sites within
minerals should have different isotopic values as vibrational frequencies, frequency shifts
and bond strengths are different. Furthermore, oxygen bonded to hydrogen should be
isotopically lighter than that bonded to Si, Al etc., and dramatically so at low temperatures.
The difference in internal oxygen isotope composition within kaolinite and illite was
subsequently quantitatively demonstrated to be a function of temperature (Zheng 1993).
This theory and experimental work was further supported by the calculated first-principles
density-functional theory between kaolinite and water (Méheut et al. 2007).

Kaolinite internal fractionation and temperature calculation

Dehydroxylation of kaolinite results in 2/7 of the oxygen atoms in the residue being derived
from the hydroxyl group (Gilg pers. comm. 2017) following the reaction:

Al Si O (OH) → 2 H O + Al Si O

Assuming equilibrium fractionation occurs during dehydroxylation, the known isotopic


composition of bulk kaolinite (denoted the subscript kaolinite (bulk)) and residual kaolinite
(denoted the subscript kaolinite (residual)), the isotopic composition of the isotopic composition of
the hydroxyl oxygen (denoted the subscript kaolinite (OH)) can be calculated as:

236
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

7 2
δ O ( ) = ∗δ O ( ) + ∗δ O ( )
9 9

9 7
δ O ( ) = (δ O ( ) − δ O ( ))
2 9

The isotopic composition of the structural non-OH kaolinite (denoted the subscript kaolinite
(non-OH)) can this be calculated as the residual kaolinite is the sum of the non-OH and the and
water, thus:

5 2
δ O ( ) = ∗δ O ( ) + ∗δ O ( )
7 7

7 2
δ O ( ) = (δ O ( ) − ∗δ O ( )
5 7

Fractionation of oxygen between the hydroxyl group and the structural oxygen sites in
kaolinite is a function of the temperature of crystallisation (Méheut et al. 2007). The
temperature of kaolinite formation can thus be determined by the oxygen isotope
fractionation between the non-hydroxyl group (denoted the subscript kaolinite (non-OH)) and
hydroxyl group (denoted the subscript kaolinite (OH)), using the equation:

10 𝑙𝑛𝛼 ( ) ( )

= −2.81 + 1.192 ∗ 10 ∗ 𝑇 + 3.43 ∗ 10 ∗𝑇 − 1.998 ∗ 10 ∗𝑇

Where T is the temperature in °K and ln𝛼Ox is the logarithmic fractionation factor of oxygen
isotopes between the two sites. This equation assumes complete isotopic equilibrium
between fluid and kaolinite that is not modified post-crystallisation.

Zheng (1993) calculated that oxygen isotope fractionation between the whole kaolinite and
hydroxyl group within kaolinite is a function of the temperature, determined by the
equation:

10 𝑙𝑛𝛼 ( ) ( ) = 2.71 ∗ 10 ∗ 𝑇 + 2.64 ∗ 10 ∗ 𝑇 − 1.3

The calculated formational temperature using these two methods yielded markedly
different temperatures (Table 7.4). Nevertheless, the kaolinite that occurs at depth appears
to have a higher calculated formational temperature than kaolinite occurring at shallower
depths (Fig. 7.10). The formational temperatures are expected to be between 150-200 °C-
which will be adopted in modelling in fluid compositions.

237
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Table 7.4: Calculation of oxygen isotope composition within kaolinite structure and calculated temperature (1=Méheut et al. 2007 method;
2=Zheng 1993 method).
δ18Obulk - δ18Onon- Temp. Temp.
Samples δ18OBULK δ18OResidue δ18OOH δ18Onon-OH
δ18OOH OH-δ
18O
OH (°C)1 (°C)2
JL2.154 6.4 11.9 -12.9 21.8 19.3 34.7 -103 160

238
JL2.216 4.5 8.5 -9.5 15.7 14.0 25.2 -74 243
JL1.68 5.9 10.0 -8.5 17.4 14.4 25.8 -76 236
JL2.230 3.1 6.9 -10.2 13.7 13.3 23.9 -69 257
JL1.111 5.3 9.8 -10.5 17.9 15.8 28.4 -85 210
JL2.236 1.0 4.0 -9.5 9.4 10.5 18.9 -43 331
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.10: Calculated formational temperatures of kaolinite using non-hydroxyl group - hydroxyl fractionation
method (1=Méheut et al. 2007) and hydroxyl-bulk oxygen isotope method (2=Zheng. 1993) versus depth.

Illite internal fractionation and temperature calculation

During dehydroxylation of illite one water molecule is created for each illite, with 1/12 of
the oxygen in the residue being derived from the OH group, following the reaction:

K . Al . 𝑆𝑖 . 𝑂 (OH) → K . Al . 𝑆𝑖 . 𝑂 +𝐻 𝑂

The hydroxyl site (illite(OH)) isotope composition can thus be calculated as the difference
between the bulk illite oxygen isotope (illite(bulk)) and the residual oxygen isotope (illite(residual)):

10 2
δ O ( ) = ∗δ O ( ) + ∗δ O ( )
12 12

12 10
δ O ( ) = (δ O ( ) − δ O ( ))
2 12

The fractionation between the bulk illite and the hydroxyl group site in illite is theorised to
be temperature dependant (Zheng 1993), following the equation:

10 𝑙𝑛𝛼 ( ) ( ) = 0.93 ∗ 10 ∗ 𝑇 + 5.91 ∗ 10 ∗ 𝑇 − 2.47

Where T is the temperature in °K and ln𝛼Ox is the logarithmic fractionation factor of oxygen
isotopes between the two sites.

239
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Experimental studies of the internal fractionation between structural and hydroxyl oxygen
in illite appeared to be linearly dependent on the temperature of the fluid from which it
formed (Bechtel and Hoernes 1990):

1000 𝑙𝑛𝛼( ) = −0.076𝑇 + 30.42

Where T is the temperature in °C and lnα (illite-OH) is the fractionation between the illite bulk
oxygen and the hydroxyl oxygen.

Interestingly sample JL.98F appears to have a larger fractionation between sites and
consequently a lower inferred formational temperature. This is presumed to be due to
contamination from other clay minerals (Fig. 7.5).

Despite the differences in calculated temperatures using linear and non-linear isotope
fractionation (Table 7.5), they generally equate to higher formational temperatures than
kaolinite.

Table 7.5: Calculation of oxygen isotope composition within illite structure and calculated temperature using
logarithmic isotope fractionation method (1=Zheng 1993) and linear isotope fractionation (2= Bechtel and
Hoernes 1990).

Samples δ18OBULK δ18OResidual δ18OOH Temp (°C)1 Temp (°C)2


JL1.196 3.6 5.7 -6.9 306 262
JL2.98F 4.5 9.6 -21 43 65
JL1.25 6.7 8.8 -3.8 306 262

Smectite internal fractionation

The temperature dependence of fractionation between the oxygen sites in smectites has
yet to be determined. Here, a dehydroxylated smectite sample has a δ18O of 11.8±0.1 ‰
and the difference between the bulk mineral δ18O and the hydroxyl δ18O is 5.5±0.3 ‰. Thus,
the fractionation between the sites is marked, significantly higher than the analytical
uncertainty, and is therefore also likely to be temperature dependent. However, further
research is required to better define this relationship.

7.3.3 Hydrogen isotope results


The results of the hydrogen isotope analyses of the clay separates give an average δD of -96
‰, with a standard deviation of 12 ‰ (Fig. 7.11). This indicates a real natural variability,
beyond analytical error which is consistently around or better than 2 ‰. Most δD values
are between -111 ‰ and -82 ‰, with a few isotopically “heavier” samples (Fig. 7.11).
240
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.11 Histogram of hydrogen isotope results of clay minerals from Cerro Corona.

7.3.4 Local meteoric waters


The results yielded from isotope analyses of the rainwater samples are δD values of -135
‰ and -142 ‰, and δ18O values of -15.2 ‰ and -15.4 ‰ which plot off the global meteoric
water line (GMWL). Unfortunately, these samples had leaked in transit and are thus
potentially isotopically fractionated. They can be inferred to be residual equivalents of
slightly evaporated local meteoric water and thus considered to be a maximum value of
local meteoric water around the period they were sampled. Assuming a local evaporation
line slope of ~4 for high-latitude fluids along the Peruvian Andes (Gibson et al. 2008), these
fluids presumably have a higher δD and δ18O values than the original meteoric waters,
which would plot on the GMWL at around -145 ‰ and -20 ‰ δD and δ18O, respectively.

241
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

As little confidence can be placed on the analysed meteoric waters, a similar elevation
(~4000m) water sample (δD -130 ‰, δ18O -17.5 ‰) was selected as a representative local
meteoric water value from the International Atomic Energy Agency dataset (IAEA 2016).

7.4 Discussion

7.4.1 Estimation of clay formation temperatures


Estimation of the temperature and isotopic composition of fluids that deposited a mineral
can be done in a variety of ways, each of which has its own set of assumptions, strengths
and weaknesses.

The most common methods for estimation of the temperature of mineral precipitation from
a hydrothermal fluid are:

 Utilising the temperature stability fields for mineral assemblages, often themselves
a function of fluid speciation, pH and sometimes pressure (Fig. 7.12, Seedorf et al.
2005). Hydrothermal biotite and K-feldspar typically form at temperatures
between 500-600 °C; chlorite and sericite typically form at around 300-400 °C; illite
typically ~200-300 °C, and; kaolinites and smectites ~200 °C or less (Hemley and
Jones 1964);

242
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.12 Phase diagram for the system K2O-Al2O3-SiO2-H2O-KCl-HCl as a function of temperature and K+/H+
activity ratio at a fluid pressure of 1 kbar (Seedorf et al. 2005). This shows the stability ranges of hydrothermal
minerals that typically form in the porphyry environment from the alteration of aluminosilicate rock.

 Using fluid inclusion homogenisation temperatures (minimum formation


temperature unless fluids can be shown to be boiling) for primary inclusions that
formed at the time of mineral precipitation. Fluid inclusions associated with
potassic and sericitic alteration have similar salinities (Bodnar et al. 2014, Fig.
7.13). However, fluid inclusions associated with potassic alteration homogenise at
>600 to 250 °C, whereas the fluid inclusions associated with sericitic alteration
typically homogenise between 200-400 °C, with few homogenising at >400 °C
(Bodnar et al. 2014, Fig. 7.13). Significantly fewer fluid inclusions associated with
propylitic and argillic alteration have been analysed (Bodnar et al. 2014). The
sparse data suggest fluid inclusion homogenisation temperatures between 100-
350 °C for propylitic alteration and ~200-350 °C for argillic alteration and similar,
low salinities (Fig. 7.13, Bodnar et al. 2014). The fluid inclusions associated with
argillic alteration at Cerro Corona are predominantly vapour-dominated (Fig. 7.14),
thus like other fluid inclusion studies, the homogenisation temperatures for argillic
alteration could not be accurately determined.
243
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.13: Fluid inclusion homogenisation temperatures and salinities for typical alteration assemblages from
13 porphyry copper deposits (Bodnar et al. 2014). Propylitic and argillic alteration are grouped together as their
composition and homogenisation temperatures are similar.

244
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.14:Vapour-dominant fluid inclusions associated with argillic alteration from Cerro Corona. The large
vapour phase makes identification of homogenisation temperatures challenging.

 Using isotope geothermometry, based on the fractionation of isotopes of an element


between two co-precipitated mineral phases (Sheppard et al. 1971). This requires
convincing textural evidence of two minerals precipitating from a fluid (such as
quartz-sericite), which was not undertaken in this study;
 Or a combination of these techniques (Hedenquist et al. 1998).

The application of internal oxygen single-mineral geothermometry of kaolinite and illite in


the porphyry environment provides sensible results that are broadly conformable with
current understanding of argillic alteration (Table 7.4 and 7.5). This relatively new
approach limits the uncertainty to the oxygen isotope composition of a single mineral and
assumes no post-formation exchange.

7.4.2 Calculation of fluid temperatures and compositions


Of the six samples of kaolinite that have well-constrained formation temperatures, only five
samples have both bulk oxygen and hydrogen isotope analyses. The isotopic composition
of the water was calculated using the fractionation between kaolinite-water for oxygen and

245
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

hydrogen by substituting inferred temperatures, in the stability field of kaolinite (T) and
measured δ18O and δD values of kaolinite into equations (Sheppard and Gilg 1996):

Hydrogen: 1000𝑙𝑛𝛼 = −2.2 ∗ 10 ∗ 𝑇 − 7.7

𝛿𝐷 = 𝛿𝐷 + 2.2 ∗ 10 ∗ 𝑇 + 7.7

Oxygen: 1000𝑙𝑛𝛼 = 2.76 ∗ 10 ∗ 𝑇 − 6.75

𝛿 𝑂 =𝛿 𝑂 − 2.76 ∗ 10 ∗ 𝑇 + 6.75

The calculated δ18O and δD values suggest a formational fluid with a composition between
meteoric and magmatic composition (Table 7.6 and Fig. 7.16).

Table 7.6: The calculated isotope composition of fluids in equilibrium with kaolinite at temperatures inferred
from kaolinite stability field (Seedorf et al. 2005)

Sample Formation
δ18O Kaolinite δD Kaolinite δ18OH2O δDH2O
number temperature (°C)

JL1.68 6.0 -103 166 -1.6 -84

JL2.154 6.5 -89 127 -4.0 -68

JL2.216 4.5 -82 165 -3.1 -63

JL2.230 3.1 -111 178 -3.7 -92

JL2.236 1.0 -105 199 -4.6 -88

Average 167

The hydrogen isotope fractionation factor for illite - water is adapted from the muscovite-
water hydrogen isotope fractionation factors (Suzuoki and Epstein 1976), in that it is
extrapolated to down to lower temperature (Fig. 7.15). Likewise, the oxygen isotope
fractionation for illite-water is extrapolated to up to higher temperature (Sheppard and
Gilg 1996; Fig. 7.15). The isotope composition of the fluid exchanging with illite is
calculated by utilising temperatures (T) inferred from stability field of illite-muscovite and
the measured isotope composition into the equations:

246
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Hydrogen: 1000𝑙𝑛𝛼 = −22.1 ∗ 10 ∗ 𝑇 + 19.1

𝛿𝐷 = 𝛿𝐷 + 22.1 ∗ 10 ∗ 𝑇 − 19.1

Oxygen: 1000𝑙𝑛𝛼 = 2.39 ∗ 10 ∗ 𝑇 − 3.76

𝛿 𝑂 =𝛿 𝑂 − 2.39 ∗ 10 ∗ 𝑇 + 3.76

Figure 7.15: Fractionation between muscovite and water for hydrogen isotopes, solid green line, (Suzuoki and
Epstein 1976) extended to lower temperatures (dashed green) and oxygen isotope fractionation between illite
and water (solid golden line, Sheppard and Gilg 1996) extended to higher temperatures (dashed golden line). The
calculated formation temperatures are derived from internal oxygen isotope values and are used to calculate the
isotopic differences between the mineral phases and the fluids from which they precipitated.

The isotope composition of the formational waters of illite alteration appears to be closer
to that of a magmatic fluid than a meteoric sourced fluid (Fig. 7.16).

The isotope composition of fluids forming smectite was calculated using fractionation
factors for oxygen (Savin and Lee 1988) and hydrogen (Capuano 1992):

Oxygen: 1000𝑙𝑛𝛼 / = (2.58 − 0.19 ∗ 𝐼) ∗ 10 ∗ 𝑇 − 4.19

𝛿 𝑂 =𝛿 𝑂 / − ((2.58 − 0.19 ∗ 𝐼) ∗ 10 ∗ 𝑇 − 4.19)

Hydrogen: 1000𝑙𝑛𝛼 = −45.3 ∗ 10 ∗ 𝑇 + 94.7

𝛿𝐷 = 𝛿𝐷 − (−45.3 ∗ 10 ∗ 𝑇 + 94.7)

247
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Where T is the temperature of formation and I is illite fraction in an illite-smectite mix


(Savin and Lee 1988). The temperatures used to calculate the fluid composition are within
the stability of field of smectite, 150-200 °C (Hemley and Jones 1964, Fig. 7.12). At
temperatures below ~150 °C these calculated waters plot to the left of the global meteoric
water line (i.e. have lower δ18O values than GMWL; Fig. 7.15) and thus likely discounts these
low formational temperatures.

For samples that contained >50 wt % kaolinite, illite or smectite, the isotope fractionation
factor between the dominant mineral and water was used to estimate the formational fluid
composition. The inferred temperature range of formation for kaolinite was between 150-
200 °C, 320-360 °C for illite and 150-200 °C for smectite. The results of these calculated
fluid compositions depict a linear trend between inferred Miocene meteoric waters and the
felsic magmatic water box Fig. 7.16).

248
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.16: Calculated formational fluids for kaolinite, illite and smectite at specific temperature for samples in which these clay minerals comprise of at least 50 wt %. The analytical error is
displayed in the top left.

249
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

7.4.3 Modelling of exchanged meteoric and magmatic waters and


mixing curves
The fluids responsible for kaolinitisation and illitisation plot in distinct fields between
purely magmatic and meteoric compositions (Fig. 7.16), suggestive of a model involving the
mixing of meteoric and magmatic fluids. Nevertheless, a counter hypothesis can be
constructed involving the calculation of a single fluid exchanging with the host rocks to
deliver a O and H isotopic trajectory that emerges from this data.

As a fluid flows through a rock it isotopically exchanges with the rock. Assuming a closed
system, whereby the fluid equilibrates with the rock, the exchanged isotopic composition
of the fluid can be calculated (Ohmoto and Rye 1974):

𝑤
𝛿 −𝛥 + 𝑟 ∗𝛿
𝛿 = 𝑤
1 + (𝑟 )

𝛿 =the final exchanged water composition,

𝛿 is the initial isotopic composition of the rock,

𝛥 is the fractionation factor between the rock and the water at the specified
temperature,

is the ratio of exchangeable atoms of the water relative to the rock, and;

𝛿 is the initial water composition (Ohmoto and Rye 1974).

The initial isotopic composition of the rock 𝛿 is the inferred composition of the Corona
Diorite (δ18O =+7.0 ‰ and δD =-70 ‰; section 6.12).

The model of water: rock exchange assumes that the oxygen composition of the fluid is
dominantly controlled by the interaction with plagioclase and hydrogen isotope
composition is controlled by the interaction with biotite. This is based on the observation
that plagioclase phenocrysts and biotite alter to clay (Chapter 3, section 3.6.4).

The modelled oxygen isotope fractionation factor between plagioclase and water is
extrapolated to lower temperatures from experimental studies of water-plagioclase
fractionation between 350-800 °C (O’Neil and Taylor 1967). This fractionation is
demonstrated to be controlled by the anorthite content of the plagioclase (𝛽), such that:

250
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

1000𝑙𝑛 = (2.91 − 0.76 ∗ 𝛽) ∗ (10 ∗ 𝑇 ) − 3.41 − 0.41 ∗ 𝛽

In this study, the average fractional anorthite content of the phenocrysts is used
(An=0.408) (Chapter 3, section, 3.4.1.3).

The modelled hydrogen isotope exchange between biotite and water is extrapolated below
the experimental conditions of 450-800 °C to lower temperatures (Suzuoki and Epstein
1976).

1000𝑙𝑛 = −21.3 ∗ (10 ∗ 𝑇 ) − 2.8

The ratio of exchangeable atoms of oxygen (or hydrogen) relative to the rock is

calculated:

𝑤𝑎𝑡𝑒𝑟
𝑤 𝑚𝑎𝑠𝑠 𝑟𝑎𝑡𝑖𝑜
= 88.98 ∗ 𝑟𝑜𝑐𝑘
𝑟 𝑝𝑒𝑟𝑐𝑒𝑛𝑡 𝑜𝑥𝑦𝑔𝑒𝑛 𝑟𝑜𝑐𝑘

𝑤𝑎𝑡𝑒𝑟
𝑤 𝑚𝑎𝑠𝑠 𝑟𝑎𝑡𝑖𝑜
= 11.11 ∗ 𝑟𝑜𝑐𝑘
𝑟 𝑝𝑒𝑟𝑐𝑒𝑛𝑡 ℎ𝑦𝑑𝑟𝑜𝑔𝑒𝑛 𝑟𝑜𝑐𝑘

The 𝑚𝑎𝑠𝑠 𝑟𝑎𝑡𝑖𝑜 effectively decreases from 1 as the water mass flows through and

equilibrates with an equal mass of rock.

The percentage of oxygen and hydrogen in the initial rock are estimated to be 49.0 ± 2.3 %
and 0.44 ± 0.04 %, respectively. These percentages are calculated from the average
composition of the four least altered biotite diorites, assuming all LOI is H2O (Appendix 7).

Meteoric waters (δ18O = -17.5 ‰ and δD = -130 ‰), hydrous dacite fluid (δ18O = +7.5 ‰
and δD = -65 ‰), hydrous andesite (δ18O = +7 ‰ and δD = -40 ‰) and fluids in equilibrium
with low aluminium amphibole (δ18O = +7.5 ‰ and δD = -80 ‰) are used as model end-
member fluid compositions (𝛿 ).

The δD compositions of hydrous andesites and dacites were determined assuming


equilibrium between amphiboles and fluids in the nearby Yanacocha district (Chambefort
et al. 2013). The compositions of the hydrous andesite water and hydrous dacite water plot
within the “felsic magmatic water” and “primary residual magmatic water” boxes,
respectively (Taylor 1992; Taylor 1974). The fluids in equilibrium with the low aluminium
amphibole (LAA) are interpreted to have been predominantly derived from 40 % degassing
of a dacitic melt (Chambefort et al. 2013).

251
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

7.4.4 Results of modelling exchanged fluids

Modelling kaolinite-forming fluids

Based on the internal fractionation of oxygen in kaolinite, the fluid temperatures that
formed kaolinite are probably >120 °C, but based on kaolinite stability < 200°C, thus
modelling a variety of fluid compositions around this temperature provides an insight into
potential sources of fluid that formed the kaolinite alteration.

Meteoric fluids

Modelling modern meteoric fluid compositions exchanging with host rock at temperatures
between 170 - 240 °C with water: rock ratios of 1:8 to 1:18 clearly plot within the kaolinite
fluid composition (Table 7.7, Fig. 7.17). The average fluid composition of kaolinite can
simplistically be modelled as local meteoric water at ~196 °C exchanging with ~ 12 times
as much rock.

Table 7.7: Calculated water: rock exchange ratios and temperatures required to evolve meteoric water to
compositions comparable to those of kaolinite-forming fluids.

Kaolinite forming fluids Exchanged meteoric water

Sample Temperature a δ18OH2O δDH2O Temperature b Water : rock

JL1.68 236 -1.60 -84 243 1:12

JL2.154 160 -3.97 -68 176 1:15

JL2.216 243 -3.09 -63 189 1:18

JL2.230 257 -3.68 -92 209 1:8

JL2.236 331 -4.56 -88 184 1:9

Average 245 -3.38 -79 196 1:12

The temperature a is the temperature calculated from internal fractionation of oxygen


(Zheng 1993), whereas temperature b is the temperature of the exchanging meteoric
water

Magmatic fluids

252
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Magmatically sourced fluids are unlikely to be exclusively involved in exchanging with host
Corona diorite to form kaolinite alteration. As magmatically derived fluids interact with
increasing amounts of host rock their isotopic compositions shift to higher δD and lower
δ18O values (Fig. 7.17). Thus, these fluids cannot realistically be modelled to be the
exclusive fluid involved in the formation of kaolinite.

253
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.17: Evolving compositions of fluid from local meteoric, hydrous dacite and a hydrous dacite that has degassed by ~40 % (low Al-amphibole) sources by interaction with host rock at
specified temperatures and decreasing water: rock ratios. See explanation in text.

254
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Modelling illite-forming fluids

The results of internal fractionation of oxygen in illite suggest a formational temperature


was ~300 °C. Modelling meteoric and magmatic fluids at this temperature and exchanging
them with various proportions of diorite host rock may provide an indication of the
ultimate source of the fluids that formed the illite alteration.

Meteoric fluids

To obtain isotope compositions equivalent to those of the illite-forming fluids, a hot


meteoric fluid is required to exchange with ~ 20 to 60 times as much rock at temperatures
of 300-440 °C (Table 7.8, Fig. 7.18). These results are similar to the temperature calculated
from internal fractionation of illite.

Table 7.8: Calculated water: rock exchange ratios and temperatures required to evolve meteoric water to
compositions comparable to those of illite-forming fluids.

Illite forming fluids Exchanged meteoric water

Sample Temperature a δ18OH2O δDH2O Temperature b Water : rock

JL1.196 306 1.67 -76 328 1:20

JL1.25 306 4.73 -57 438 1:57

JL2.98 43 1.81 -48 297 1:45

Average 2.74 -60 344 1:35

The temperature a is the temperature calculated from internal fractionation of oxygen


between the bulk illite and hydroxyl site (Zheng 1993), whereas temperature b is the
temperature of the exchanging meteoric water

Magmatic fluids

The fluid from a degassing of a dacitic melt (i.e. low-aluminium amphibole fluid,
Chambefort et al. 2013) could evolve to form a fluid of similar isotopic composition to the
average illite fluid composition by exchanging with about seven times as much rock (i.e.
W:R ratio of 1:7) at ~264 °C (Table 7.9, Fig. 7.18). This is well below the temperature

255
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

calculated by internal fractionation method and not all samples can be modelled using this
composition fluid.

Table 7.9: Calculated water: rock exchange ratios and temperatures required to evolve an exsolved dacitic fluid
to compositions comparable to those of illite-forming fluids

Illite forming fluids Exchanging fluid from degassing dacitic melt

Sample δ18OH2O δDH2O Temperature Water : rock

JL1.196 1.67 -76 N/A N/A

JL1.25 4.73 -57 374 1:14

JL2.98 1.81 -48 250 1:13

Average 2.74 -60 264 1:7

A fluid that exsolved from hydrous dacite cannot solely form a fluid of similar composition
to the average isotope illite-forming fluid by exchange with the rock at any temperature.

256
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.18: Evolving compositions of fluid from local meteoric, hydrous dacite and a hydrous dacite that has degassed by ~40 % (low Al-amphibole) source by interaction with host rock at
specified temperatures and decreasing water: rock ratios. See explanation in text.

257
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Model of mixing fluids

Mixing between meteoric and felsic magmatic water may explain the overall trend of
smectite, kaolinite and illite distribution (Fig. 7.19). Formational fluids of smectite at ~150-
200 °C is dominantly meteoric, requiring less than 30 wt. % of a magmatic fluid input. Fluids
responsible for the formation of kaolinite appear to plot as a mixture between meteoric and
magmatic fluids at temperature ranges between 150-200°C. Illite can be simplistically
modelled to be formed from fluids of predominantly magmatic origin at ~ 320-360°C (Fig.
7.19).

258
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Figure 7.19: Model of magma mixing between felsic magmatic and meteoric water. Fluid in equilibrium with smectite is dominated by >70 wt% meteoric waters, whereas kaolinite is dominated
by 40-65 wt. % and illite is predominated by magmatic waters. Analytical error is shown in top left corner.

259
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

7.5 Summary
The use of single mineral oxygen isotope geothemometry in illite and kaolinite has been
shown to have merit in the case for the Cerro Corona alteration system, providing more
realistic estimates of the temperature of mineral deposition than those inferred from fluid
inclusion data or calculated from apparently coexisting mineral phases. More broadly, the
temperatures calculated using this method agree with argillic and sericitic alteration
assemblages in porphyry and epithermal environments (Bodnar et al. 2013). Illitisation
formed at temperatures ~300 °C and kaolinitisation at temperatures >150 °C.

Although no direct temperature constraint was placed on the smectite, it seems plausible
that this mineral formed at temperatures >150°C. It seems highly unlikely that this
alteration formed at temperatures below ~150°C, as the modelled fluids at this
temperature in equilibrium with smectite plot to the left of the GMWL.

Models of meteoric fluids exchanging with the Corona quartz diorite host at variable
temperatures demonstrates that these fluids may form the kaolinite and illite minerals.
Meteoric waters are required to be heated to calculated temperatures, potentially a deeper
magmatic body which may cause a high geothermal gradient. Kaolinite forms at meteoric
water: rock of ~1:10, whereas illite requires meteoric water: rock of ~1:35.

Modelling also suggests that magmatic fluids may have been involved in the formation of
illite alteration. This requires a low magmatic water: rock of 1:10 to cause illite alteration.
Alternatively, a model involving mixing of a large proportion of magmatic fluid mixing with
meteoric waters is plausible.

It seems highly unlikely that purely magmatic fluids are responsible for the smectite and
kaolinite dominant argillic alteration. A more plausible explanation is a mixture of meteoric
and magmatic fluids at temperatures of ~ 150-200°C. At these temperatures, smectite is
formed from predominantly meteoric fluids (>70 %), whereas kaolinite alteration requires
approximately an equal mix of magmatic and meteoric fluid.

In conclusion, the smectite-dominant argillic alteration at Cerro Corona appears to be a


product of supergene alteration, where meteoric fluids are heated to > 150°C. The kaolinite-
dominant argillic alteration appears to be formed from equal mixture of magmatic and
meteoric waters and/or where meteoric waters are heated whilst percolating through host
rock with a water: rock ratio of ~ 1: 10. Fluid forming illite-dominant argillic alteration
appears to have a high magmatic component and/or meteoric fluids have been heater to
~300°C and interacted with host rock at a ratio of ~ 1: 35. If all clays are considered to form
260
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

during a single alteration event and meteoric source, then increasing temperature and
decreasing fluid: rock ratio would result in an illite-dominant argillic alteration at depth, a
kaolinite-dominant argillic alteration at intermediate depth and smectite-dominant
alteration would form at/near surface level by heated groundwater.

Further work

The application of internal isotope fractionation is not considered to be confined to this


type of hydrothermal alteration system, but rather where there is variable temperature
contrasts if clay mineral formation and variable isotope composition of the fluid. This lends
itself well to fluid flow within faults, around veins and in large geothermal convection cells.
It is recommended that this procedure is applied to a variety of alteration systems to
compare results in these environments.

To make this method of analyses more robust, detailed experimentation of the isotopic
fractionation of oxygen in these clay minerals is recommended. Subsequently, systematic
refinement of the oxygen isotope extraction is required. The author recommends the
addition of an infrared spectrometer, with or without a gravimetric analysis, to the sample
treatment chamber. This would allow for improvement in specific bond degradation and
internal isotope fractionation.

7.6 References
Akagi, T., Franchi, I. A. and Pillinger, C. T. (1993) Oxygen isotope analysis of quartz by
neodymium-yttrium aluminium garnet laser-fluorination. Analyst. 118 (12), 1507-1510.

Bechtel, A. and Hoernes, S. (1990) Oxygen isotope fractionation between oxygen of


different sites in illite minerals: a potential single-mineral thermometer. Contributions to
Mineralogy and Petrology. 104 (4), 463-470.

Bodnar, R. J., Lecumberri-Sanchez, P., Moncada, D., and Steele-MacInnis, M. (2014). 13.5—
Fluid inclusions in hydrothermal ore deposits. Treatise on Geochemistry, Second Edition
edn. Elsevier, Oxford, 119-142.

Borthwick, J. and Harmon, R. S. (1982) A note regarding CIF3 as an alternative to BrF5 for
oxygen isotope analysis. Geochimica Et Cosmochimica Acta. 46 (9), 1665-1668.

Capuano, R. M. (1992) The temperature dependence of hydrogen isotope fractionation


between clay minerals and water: Evidence from a geopressured system. Geochimica Et
Cosmochimica Acta. 56 (6), 2547-2554.

261
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Chambefort, I., Dilles, J. D., Longo, A. A. (2013) Amphibole Geochemistry of the Yanacocha
Volcanics, Peru: Evidence for Diverse Sources of Magmatic Volatiles Related to Gold Ores.
Journal of Petrology. 54 (5), 1017-1046.

Craig, H. (1961) Standard for reporting concentrations of deuterium and oxygen-18 in


natural waters. Science. (133), 1833–1834

Cuadros, J., Huertas, F., Delgado, A. and Linares, J. (1994) Determination of hydration (H2O-
) and structural (H2O+) water for chemical analysis of smectites. Application to Los
Trancos smectites, Spain. Clay Minerals. 29 (2), 297-300.

Donnelly, T., Waldron, S., Tait, A., Dougans, J. and Bearhop, S. (2001) Hydrogen isotope
analysis of natural abundance and deuterium-enriched waters by reduction over
chromium on-line to a dynamic dual inlet isotope-ratio mass spectrometer. Rapid
Communications in Mass Spectrometry, 15 (15), 1297-1303.

Gibson, J. J., Birks, S. J. and Edwards, T. W. D. (2008) Global prediction of δA and δ2H-δ18O
evaporation slopes for lakes and soil water accounting for seasonality. Global
Biogeochememical Cycles, 22 (2), 1-12.

Giggenbach, W. F. (1992) Isotopic shifts in waters from geothermal and volcanic systems
along convergent plate boundaries and their origin. Earth and Planetary Science Letters.
113 (4), 495-510.

Gilg, H. A. and Sheppard, S. M. F. (1996) Hydrogen isotope fractionation between kaolinite


and water revisited. Geochimica Et Cosmochimica Acta. 60 (3), 529-533.

Girard, J. P., and Savin, S. M. (1996). Intracrystalline fractionation of oxygen isotopes


between hydroxyl and non-hydroxyl sites in kaolinite measured by thermal
dehydroxylation and partial fluorination. Geochimica Et Cosmochimica Acta, 60(3), 469-
487.

Godfrey, J. D. (1962) The deuterium content of hydrous minerals from the East-Central
Sierra Nevada and Yosemite National Park. Geochimica Et Cosmochimica Acta. 26 (12),
1215-1245.

Guggenheim, S., and Van Groos, A. K. (2001). Baseline studies of the clay minerals society
source clays: thermal analysis. Clays and Clay Minerals, 49(5), 433-443.

262
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

Hamza, M. S. and Epstein, S. (1980) Oxygen isotopic fractionation between oxygen of


different sites in hydroxyl-bearing silicate minerals. Geochimica Et Cosmochimica Acta. 44
(2), 173-182.

Hedenquist, J. W., Arribas, A. and Reynolds, T. J. (1998) Evolution of an intrusion-centered


hydrothermal system; Far Southeast-Lepanto porphyry and epithermal Cu-Au deposits,
Philippines. Economic Geology. 93 (4), 373-404.

Hemley, J. J., and Jones, W. R. (1964). Chemical aspects of hydrothermal alteration with
emphasis on hydrogen metasomatism. Economic Geology, 59(4), 538-569.

Hut, G. (1987). Consultants' Group Meeting on Stable Isotope Reference Samples for
Geochemical and Hydrological Investigations, IAEA, Vienna, 16 - 18 September 1985;
report to the Director General. International Atomic Energy Agency, Vienna

International Atomic Energy Agency and World Meteorological Organization. (2016)


Global Network of Isotopes in Precipitation. The GNIP Database. Available from:
http://www.iaea.org/water.

Kahle, M., Kleber, M., and Jahn, R. (2002). Review of XRD-based quantitative analyses of clay
minerals in soils: the suitability of mineral intensity factors. Geoderma, 109 (3), 191-205.

Lawrence, J. R. and Taylor, H. P. (1972) Hydrogen and oxygen isotope systematics in


weathering profiles. Geochimica Et Cosmochimica Acta. 36 (12), 1377-1393.

Méheut, M., Lazzeri, M., Balan, E. and Mauri, F. (2007) Equilibrium isotopic fractionation in
the kaolinite, quartz, water system: Prediction from first-principles density-functional
theory. Geochimica Et Cosmochimica Acta. 71 (13), 3170-3181.

Moore, D. M., and Reynolds, R. C. (1989). X-ray Diffraction and the Identification and
Analysis of Clay Minerals 378, 227-260.

Nelson, S. T. (2000). A simple, practical methodology for routine VSMOW/SLAP


normalization of water samples analyzed by continuous flow methods. Rapid
Communications in Mass Spectrometry, 14(12), 1044-1046.

Ohmoto, H. and Rye, R.O. (1974) Hydrogen and oxyen isotopic compositions of fluid
inclusions in the Kuroko deposits, Japan. Economic Geology, 69, 468-481.

O’Neil, J. R. and Pickthorn W. J. (1988) Single-mineral oxygen isotope thermometry.


Chemical Geology. 71 (4), 369.

263
Chapter Seven: Oxygen and hydrogen isotope analyses of argillic alteration

O'Neil, J. R., and Taylor, H. P. (1967). The oxygen isotope and cation exchange chemistry of
feldspars. American Mineralogist, 52, 1414-1437.

Savin, S. M. and Epstein, S. (1970) The oxygen and hydrogen isotope geochemistry of clay
minerals. Geochimica Et Cosmochimica Acta. 34 (1), 25-42.

Savin, S. M. and Lee, M. (1988) Isotopic studies of phyllosilicates. In: Bailey, S.W. (Eds.)
Hydrous Phyllosilicates. Reviews in Mineralogy (19), Mineralogical Society of America,
Washington, D.C. 189-223

Sharp, Z. D. (1990) A laser-based microanalytical method for the in situ determination of


oxygen isotope ratios of silicates and oxides. Geochimica Et Cosmochimica Acta. 54 (5),
1353-1357.

Sharp, Z. D. (1992). In situ laser microprobe techniques for stable isotope analysis.
Chemical Geology: Isotope Geoscience section, 101(1), 3-19.

Sheppard, S. M., Nielsen, R. L., and Taylor, H. P. (1971). Hydrogen and oxygen isotope ratios
in minerals from porphyry copper deposits. Economic Geology, 66(4), 515-542.

Sheppard, S. M. F. and Gustafson, L. B. (1976) Oxygen and hydrogen isotopes in the


porphyry copper deposit at El Salvador, Chile. Economic Geology. 71 (8), 1549-1559.

Sheppard, S. and Gilg, H. A. (1996) Stable isotope geochemistry of clay minerals. Clay
Minerals. 31 (1), 1-24.

Suzuoki, T. and Epstein, S. (1976) Hydrogen isotope fractionation between OH-bearing


minerals and water. Geochimica Et Cosmochimica Acta. 40 (10), 1229-1240.

Taylor, H. P. J. (1974) The application of oxygen and hydrogen isotope studies to problems
of hydrothermal alteration and ore deposition. Economic Geology. 69: 843-883

Taylor B. E. (1992) Degassing of H2O from rhyolitic magma during eruption and shallow
intrusion, and the isotopic composition of magmatic water in hydrothermal systems.
Geological Survey of Japan Report 279: 190–194

Valley, J.W., Kitchen, N., Kohn, M.J., Niendorf, C.R., and Spicuzza, M.J. (1995) UWG-2, a garnet
standard for oxygen isotope ratios: Strategies for high precision and accuracy with laser
heating: Geochimica Et Cosmochimica Acta, (59), 5223- 5231.

Zheng, Y. F. (1993). Calculation of oxygen isotope fractionation in hydroxyl-bearing


silicates. Earth and Planetary Science Letters, 120(3-4), 247-263.
264
Chapter Eight: Discussion

8 Chapter Eight: Discussion

8.1 Geodynamic setting of Cerro Corona and source


magmas
The role of the discussion chapter is to bring together the findings of each chapter into a
coherent model for the development of the Cerro Corona deposit and to consider the
implications of this for hydrothermal alteration in porphyry systems in general.

Despite the Hualgayoc district lacking published whole rock geochemistry, the Cerro Corona
intrusive complex is placed within the context of the temporal equivalents of the Yanacocha
district, situated ~80 km to the south, and together encompassing the broader Cajamarca
district (Chapter 3, Fig. 3.5). The magmas associated with porphyry ore deposits in the
Cajamarca district are most likely linked to the thickening of the continental crust in the
Miocene, coincident with subduction of the buoyant “Inca Plateau” beneath the Cajamarca
district (Chapter 3, section 3.3, Rosenbaum et al. 2005). The adakite-like geochemical
signatures observed can be explained by the periodic recharge of deeply-derived hydrous
basaltic or andesitic melts, and mixing with silicic melts derived by crustal melting (Castillo et
al. 1999; Macpherson et al. 2006; Chiaradia et al. 2009; Davies 2002; Longo 2005; Longo et al.
2010). The ascent of these melts into shallow-level magma chambers was controlled by pre-
existing structures and the upper crustal stress regime, which was predominantly compressive
at the time owing to the high convergence rate of the subducting Nazca plate. The melting and
assimilation of upper crustal material is thought to have contributed chlorine and water to
sulphur- and metal-bearing, mantle-derived melts (Chambefort et al. 2013) - crucial for
formation of these types of deposits.

The igneous petrology of the Cerro Corona intrusions reveals an increasing crystallisation
temperature, pressure (or depth, Fig. 4.47) and hydration (Fig. 4.48) of the source magmas.
This is exemplified by the amphibole phenocrysts from Cerro Corona which have a similar
paragenetic sequence to those in the Yanacocha District (Chambefort et al. 2013). Their
chemistry reflects crystallisation from an oxidised, hydrous melt containing magnetite and
anhydrite with minor oligoclase, most likely recording the earliest major crystallisation event.

265
Chapter Eight: Discussion

Subsequently, biotite phenocrysts crystallised and contain magnetite and apatite inclusions
with minor rutile and ilmenite. Finally, late quartz phenocrysts display embayment’s,
potentially indicating multiple events of bubble nucleation and fluid exsolution (Vasyukova et
al. 2013). Interestingly, plagioclase microprobe analyses show that nearly all crystals do not
display an ‘excess’ of aluminium (Fig. 4.46) seen in other “fertile” porphyry deposits and linked
to high melt water content (Williamson et al. 2016). However, amphibole hygrometer
calculations (Ridolfi et al. 2010) suggest that magmas at Cerro Corona were hydrous,
consistent with the adakite-like geochemistry. This brings into question the validity of the use
of Al excess in feldspars as a tool for porphyry copper deposit exploration.

The exact timing of uplift and erosion in the Cajamarca district remains uncertain; suffice to
say there was a dramatic increase in crustal uplift rate in the late Miocene. The current studies
indicate that uplift increased from ~200-300 m/My in the late Miocene, to a remarkable ~600-
3000 m/My during the Pliocene. The conclusion of this thesis suggests that the significant uplift
rate may have played a critical role in the telescoping of typically lower temperature and more
distal alteration assemblages onto the porphyry deposit at Cerro Corona.

Porphyry and high sulphidation deposits in the Cajamarca district are controlled by pre-
existing structural weaknesses, such as those observed at Michiquillay (Hollister et al. 1974)
and El Galeno (Davies et al. 2005). Epithermal deposits occur trenchward of the porphyry belt,
suggesting an increased rate of post-deposition uplift with increasing distance from the trench.

This thesis re-evaluates the Cerro Corona complex and defines at least five intrusive phases:
two biotite quartz diorites (QD1 and QD2); hornblende granodiorite porphyry; andesitic dykes
and rhyolitic intrusions, building on the observations of mine site geologists (Uzategui and
Ayala pers. comm 2011) and previous authors (James 1998 and Sillitoe 1997).

The QD1 is the most intensely altered intrusive and contains abundant quartz veins that are
associated with early potassic alteration and the main ore-forming event. These are crosscut
by the QD2 intrusions, termed the ”Barren Cores” by mine geologists. The QD2 show intrusive
crosscutting relationships with QD1 in drill core and clearly displace the grade shells, thus the
QD2 are in fact, separate, later intrusive event(s) and thus the term “core” is a misnomer. The
very similar primary mineralogy and igneous geochemistry of QD1 and QD2 suggest they are
derived from a similar magmatic source, begging the question as to why the early phase is the
most intensely mineralised. The relative timing between the two post-ore intrusions (QD2) is
still uncertain.
266
Chapter Eight: Discussion

8.2 Mineralogical controls of chemical mass transfer


effects
The similar primary mineralogy and geochemistry of the intrusions at Cerro Corona provide
an ideal situation for a mass transfer analysis case study. The mass transfer study has
demonstrated an overall conservative element dilution due to bulk mass gain. These bulk mass
additions are accounted for by large gains in certain major element oxides (e.g. SiO2, Fe2O3 and
S) in each alteration stage, primarily related to change in primary alteration mineralogy. A
more in-depth evaluation of these net gains (and losses), together with the behaviour of minor
and trace elements, have provided greater insights into the nature of the fluids involved.

The hydrothermal alteration can be broadly understood in (Na+K)/Al versus K/Al geochemical
space (Fig. 8.1). The results from this study match the typical trends for hydrothermal
alteration and overprinting relationships seen in porphyry deposits worldwide (e.g. Bajo de la
Alumbrera porphyry Cu-Au, Argentina (Ulrich and Heinrich 2002); Batu Hijau porphyry Cu-
Au, Indonesia (Idrus et al. 2009); Chuquicamata porphyry Cu-Au, Chile (Arnott 2003); Duobuza
porphyry Cu-Au, Tibet (Li et al. 2011); Maher-Abad porphyry Cu-Au, Iran (Siahcheshm et al.
2014)).

267
Chapter Eight: Discussion

Figure 8.1: Molar K/Al versus (Na+K)/Al plot showing samples used in mass transfer analysis. Black squares are ideal
mineral compositions: “K-spar” is endmember orthoclase, “Plag” is end-member albite, “K-mica” is muscovite mica. “Ave
K-spar” is the average analysed composition of K-spar that has replaced [albite – or plagioclase]. “BD ave” is the bulk
rock composition of the least-altered biotite-quartz diorite. For explanation of alteration types, see Chapter 5.

The paragenesis of hydrothermal alteration observed at Cerro Corona is broadly consistent


with that reported previously (James 1998 and Sillitoe 1997). Early potassic to propylitic
alteration is coincidental with K-addition and significant copper and gold precipitation
(Chapter 5). This alteration and ore mineralisation is cross cut by the QD2 biotite-quartz
diorites. The planar habit of veins associated with sericitic alteration suggests that this
alteration is structurally controlled, thus a better understanding of the timing of faults is
important in modelling the distribution of sericitic alteration. Permeability is highest along

268
Chapter Eight: Discussion

intrusive contacts and areas of brecciation; these permeability features appear to control the
distribution of the argillic alteration, especially at depth.

8.2.1 Potassic alteration


Potassically-altered rocks have a bulk mass gain of ~50% relative to their presumed protolith
(Chapter 5), primarily due to quartz vein formation and effective expansion of the bulk rock
volume (Chapter 4). The gains in K2O, Ba and Rb at the expense of CaO and Sr, suggest a
replacement of andesine plagioclase, the dominant plagioclase phenocryst composition
(Chapter 4, section 4.2.4), by alkali-feldspar, following the reaction:

𝑁𝑎 . 𝐶𝑎 . 𝐴𝑙 . 𝑆𝑖 . 𝑂 + 0.91𝐾𝐶𝑙 + 1.2𝑆𝑖(𝑂𝐻)4
= 1.3𝐾 . 𝑁𝑎 . 𝐴𝑙𝑆𝑖 𝑂 + 0.3𝐶𝑎𝐶𝑙2 + 0.31𝑁𝑎𝐶𝑙 + 2.4𝐻2𝑂

𝐴𝑛𝑑𝑒𝑠𝑖𝑛𝑒 + 𝐾𝐶𝑙 + 𝑆𝑖𝑂2(𝑎𝑞) = 𝐾𝑓𝑒𝑙𝑑𝑠𝑝𝑎𝑟 + 𝐶𝑎𝐶𝑙(𝑎𝑞) + 𝑁𝑎𝐶𝑙(𝑎𝑞) + 𝐻2𝑂

There appears to be little removal of Na2O (Fig. 5.11), perhaps due to incorporation of an albitic
component into the alkali feldspar.

The net losses in CaO and minor loss of MgO during potassic alteration can be accounted for by
the observed breakdown of hornblende to biotite and rutile (Chapter 4 and Fig. 8.2). The
resulting fluid is enriched in iron and is acidified. Biotitization of hornblende is commonplace
in porphyry deposits (e.g. Bajo de la Alumbrera; Ulrich and Heinrich 2002). These reactions
have been shown to be plausible under conditions considered responsible for potassic
alteration (Brimhall et al. 1985).

2(𝑁𝑎, 𝐾) (𝐶𝑎 . , 𝐹𝑒 . 𝑁𝑎 . 𝑀𝑛 . )(𝑀𝑔 . , 𝐹𝑒 . 𝐴𝑙 . 𝑇𝑖 . )(𝑆𝑖 𝐴𝑙)𝑂 (𝑂𝐻, 𝐹, 𝐶𝑙) + 0.4𝑇𝑖𝑂


+ 6.1𝐻 𝑆𝑂
= (𝐾 . 𝑁𝑎 . )(𝑇𝑖 . 𝐹𝑒 . 𝑀𝑔 . )(𝑆𝑖 . 𝐴𝑙 . )𝑂 (𝑂𝐻, 𝐹, 𝐶𝑙) + 3.2𝐶𝑎𝑆𝑂
+ 2.9𝑀𝑔𝑆𝑂 + 0.5𝐹𝑒𝑂. 𝐹𝑒 𝑂 + 0.4 𝐴𝑙 + 8.4𝑆𝑖𝑂 + 4𝐻 𝑂 + 2.1𝐻

(𝑁𝑎 . 𝐾 . )(𝐶𝑎 . 𝐹𝑒 . 𝑁𝑎 . 𝑀𝑛 . )(𝑀𝑔 . 𝐹𝑒 . 𝐴𝑙 . 𝑇𝑖 . )(𝑆𝑖 . 𝐴𝑙 . )𝑂 (𝑂𝐻 . 𝐹 . 𝐶𝑙 . )


+ 𝐻 𝑆𝑂 ( ) + 𝐾 𝑆𝑂 ( ) → 𝐵𝑖𝑜𝑡𝑖𝑡𝑒 + 𝑎𝑛ℎ𝑦𝑑𝑟𝑖𝑡𝑒 + 𝑄𝑢𝑎𝑟𝑡𝑧

269
Chapter Eight: Discussion

Figure 8.2: Breakdown of amphibole phenocryst. Numbered red spots are locations of electron microprobe analyses.

The loss of trace elements V and Cr provide further evidence for hornblende destruction, the
primary igneous mineral inferred to host these elements (Stanton, 1994). Microprobe analyses
of unaltered amphiboles from post-mineralisation dykes contain up to 0.27±0.02 wt.% and
0.1±0.01 Cr2O3 and V2O3, respectively.

Aside from the expected gains of Cu and Au, related to hypogene precipitation of chalcopyrite
(and bornite) and potentially native gold in the potassic zone (Chapter 4), the most consistent
addition in potassic alteration is the addition of Ga. Ga3+ is inferred to substitute into the Al3+
site within alkali feldspars during this alteration (Goldsmith 1950). There is currently limited
research on the distribution of gallium in magmatic-hydrothermal systems (Prokof’ev et al.
2014). What is known is that the concentration of gallium varies by 1-3 orders of magnitude
within a single ore deposit, with the highest concentrations of gallium being linked to fluids of
magmatic origin, particularly those involved in copper-porphyry deposits (Prokof’ev et al.
2016). The results of this study suggest that hydrothermal fluids involved in potassic alteration
were significantly enriched in gallium with respect to aluminium and most likely derived from
a magma.
270
Chapter Eight: Discussion

The REE profile of the potassically-altered samples cannot be solely explained by dilution
resulting from bulk mass gain (Fig. 8.3), except for sample C11.JL.183. The altered samples are
consistently depleted in Light-REE (LREE) and Heavy-REE (HREE), with a positive Eu anomaly
relative to diluted protolith. This suggests that the REE were removed during potassic
alteration, except for europium which is preferentially retained. This can be explained by the
retention of Eu2+ in alkali feldspar, whilst trivalent REE were removed via amphibole
breakdown.

Figure 8.3: Chondrite-normalized (Sun and McDonough 1989) REE diagram for the average least-altered biotite-
quartz diorite porphyry and the modelled dilution of this REE pattern by bulk mass gain of 20% increments compared
with potassically-altered biotite-quartz diorite samples.

The samples with the largest bulk mass gains have the greatest additions of SiO2, Fe2O3 and
variable CaO addition. This supports the observation that this stage alteration assemblage is
spatially associated with quartz-magnetite veinlets and variable carbonate addition.
Interestingly, Sr does not appear to correlate with CaO, in that Sr is readily removed during
this alteration. This is consistent with the common observation of Sr depletion in whole rock
in proximal propylitic alteration, which is reflected in chlorite mineral chemistry (Wilkinson et
al. 2015).

271
Chapter Eight: Discussion

8.2.2 Propylitic alteration


Propylitic alteration, much like the argillic alteration, results in less substantial bulk mass gains
of ~10 to ~30%, due to the less drastic metasomatism that is seen in the potassic and sericitic
alteration assemblages. These mass changes are reflected in the chloritization of mafic
minerals and the addition of carbonate, that broadly define this assemblage.

Gains in LOI, assumed to be predominantly due to the presence of water (±CO2?), indicates
significant hydration during propylitic alteration. K2O appears to be moderately removed from
some samples, presumed to be controlled by the alteration of biotite to chlorite (Chapter 4, Fig.
4.26). These changes are reflected in the reaction:

2K(Fe, Mg) AlSi O (OH) + 2H O + 𝑂 + 2CO + 2CaCl ( )

→ (Fe, Mg) (Si, Al) O (OH) + 4SiO + 2KCl + 2CaCO + 𝐶𝑙

2𝑏𝑖𝑜𝑡𝑖𝑡𝑒 + 2𝑤𝑎𝑡𝑒𝑟 + 𝑜𝑥𝑦𝑔𝑒𝑛 + 2𝑐𝑎𝑟𝑏𝑜𝑛 𝑑𝑖𝑜𝑥𝑖𝑑𝑒 + 2𝑐𝑎𝑙𝑐𝑖𝑢𝑚 𝑐ℎ𝑙𝑜𝑟𝑖𝑑𝑒


→ 𝑐ℎ𝑙𝑜𝑟𝑖𝑡𝑒 + 4𝑞𝑢𝑎𝑟𝑡𝑧 + 2𝐾𝐶𝑙( ) + 2𝑐𝑎𝑙𝑐𝑖𝑡𝑒 + 𝑐ℎ𝑙𝑜𝑟𝑖𝑛𝑒

P2O5 remains within error of no net change indicating that apatite may be stable. The observed
conservation of MgO is consistent with the biotite-chlorite reaction given above. The apparent
gains of K2O in some samples may be in part explained by an unrecognised prior weak potassic
alteration or a transitional assemblage from potassic to propylitic.

The bulk removal of Na2O suggests that alkali feldspar is altered to illite/muscovite in the
propylitic zone which may be macroscopically cryptic. This breakdown of feldspar may
account for some of the losses in Ba and Rb, which probably occupy K-sites. Feldspar
destruction is supported by REE patterns in that the positive Eu anomaly is inverted in some
samples (Fig. 8.4). These net changes can be partially explained by the reaction:

2.8(𝑁𝑎 . 𝐾 . )𝐴𝑙𝑆𝑖 𝑂 + 1.6𝐻 𝑂 + 𝑀𝑔


→ 0.8𝐾𝐴𝑙 𝑆𝑖 𝑂 (𝑂𝐻) + 2𝑁𝑎 + 0.2(𝑀𝑔 𝐴𝑙)(𝐴𝑙𝑆𝑖 )𝑂 (𝑂𝐻) + 5.4𝑆𝑖𝑂

2.8𝐴𝑙𝑘𝑎𝑙𝑖 𝑓𝑒𝑙𝑑𝑠𝑝𝑎𝑟 + 1.6 𝑤𝑎𝑡𝑒𝑟 + 𝑀𝑔 → 𝐼𝑙𝑙𝑖𝑡𝑒 + 2𝑁𝑎 + 0.2𝑐ℎ𝑙𝑜𝑟𝑖𝑡𝑒 + 5.4𝑞𝑢𝑎𝑟𝑡𝑧

In samples with sulphur addition, there is an increase in Fe2O3, Cu and Au suggestive of pyrite
(and/or chalcopyrite) precipitation. This is supported by petrographic observations in which
chlorite is spatially associated with chalcopyrite (Chapter 4, Fig. 4.24) and shows that
significant ore metal addition still occurred within the proximal propylitic zone, irrespective
272
Chapter Eight: Discussion

of whether it was an overprint of an existing potassic alteration stage or previously unaltered


rocks.

The REE patterns of the typical propylitically-altered samples cannot be explained by simple
mass addition to the average, potassically altered protolith (Fig. 8.4). It can be argued that some
of the LREE removed during potassic alteration were added back during propylitic alteration.
The REE pattern is controlled predominantly by secondary minerals and their relative
abundance, presumed to be principally carbonates such as calcite.

8.2.3 Sericitic alteration


Sericitisation of previously potassically-altered rocks results in highly variable bulk mass gains
from ~60 to >250%, depending on the intensity of the alteration. The strong spatial association
of this facies with planar quartz-pyrite veins (Chapter 4) indicates that variable expansion of
the rock mass through quartz veining is the principal control. This veining and is also largely
responsible for the observed gains in SiO2, Fe2O3 and S. Gains in LOI reflect the addition of water
and CO2 (bound in the muscovite and calcite structures), formed during alteration of
plagioclase to muscovite mica and calcite (Chapter 4, Fig. 4.30). This process also accounts for
the apparent gain in K2O at the expense of CaO and Na2O and the generation of abundant quartz
as seen by the reaction below, in which the albite and muscovite compositions were derived
from EDX analysis (sample C12.JL.170, Chapter 4):

3𝑁𝑎 . 𝐶𝑎 . 𝐴𝑙 . 𝑆𝑖 . 𝑂 + 2.55𝐻𝐶𝑙( ) + 1.04𝐾𝐶𝑙( ) + 0.08𝐹𝑒𝐶𝑙 ( ) + 0.23 𝑀𝑔𝐶𝑙 ( )

+ 0.29𝑂 + 0.04𝐶𝑂
→ 1.27𝐾 . 𝐶𝑎 . (𝐴𝑙 . 𝐹𝑒 . 𝑀𝑔 . )(𝑆𝑖 . 𝐴𝑙 . )𝑂 (𝑂𝐻) + 4.65𝑆𝑖𝑂
+ 0.04𝐶𝑎𝐶𝑂 + 2.91𝑁𝑎𝐶𝑙( ) + 0.65𝐶𝑙

𝐴𝑙𝑏𝑖𝑡𝑒(𝑎𝑛𝑎𝑙𝑦𝑠𝑒𝑑) + ℎ𝑦𝑑𝑟𝑜𝑐ℎ𝑙𝑜𝑟𝑖𝑐 𝑎𝑐𝑖𝑑 + 𝐾𝐶𝑙(𝑎𝑞) + 𝑓𝑒𝑟𝑟𝑜𝑢𝑠 𝑐ℎ𝑙𝑜𝑟𝑖𝑑𝑒


+ 𝑚𝑎𝑔𝑛𝑒𝑠𝑖𝑢𝑚 𝑐ℎ𝑙𝑜𝑟𝑖𝑑𝑒 + 𝑜𝑥𝑦𝑔𝑒𝑛 + 𝑐𝑎𝑟𝑏𝑜𝑛 𝑑𝑖𝑜𝑥𝑖𝑑𝑒
→ 𝑠𝑒𝑟𝑖𝑐𝑖𝑡𝑒(𝑎𝑛𝑎𝑙𝑦𝑠𝑒𝑑) + 𝑞𝑢𝑎𝑟𝑡𝑧 + 𝑐𝑎𝑙𝑐𝑖𝑡𝑒 + 𝑠𝑜𝑑𝑖𝑢𝑚 𝑐ℎ𝑙𝑜𝑟𝑖𝑑𝑒 + 𝑐ℎ𝑙𝑜𝑟𝑖𝑛𝑒

Biotite is observed to alter to muscovitic mica (Chapter 4, Fig. 4.28). This reaction primarily
controls the removal of Fe2O3, MgO and minor removal of K2O with little effect on the water
content as both are hydrous minerals with similar OH content.

The removal of niobium during sericitisation of unaltered biotite-quartz diorite is attributed


to the breakdown of biotite, as biotite commonly accounts for between 60-80% of the Nb
273
Chapter Eight: Discussion

content in similar composition biotite granodiorites (Parker et al. 1968). The niobium
redistribution warrants further investigation, as the fluids leaving this assemblage will be
relatively Nb-rich.

Sericitisation of potassically-altered biotite-quartz diorite protolith results in K2O being


reworked into the muscovite structure. However, sericitisation of fresh biotite-quartz diorite
requires K2O addition. The behaviour of Rb mimics that of K2O in that it is added during
sericitisation of plagioclase (sericitisation of unaltered biotite-quartz diorite), but is within
error of no net change in rocks that have undergone prior potassic alteration (sericitisation of
K-feldspar). Barium, unlike K2O and Rb, is removed during all stages of sericitisation. This
suggests that Ba is not accommodated in the muscovite structure and thus is partitioned into
solution under sericitic alteration conditions. Strontium behaves similarly to CaO: during
sericitisation of plagioclase in unaltered biotite-quartz diorite it is removed, whereas with
increasing sericitisation of potassically-altered biotite-quartz diorite it may be reworked into
calcite. Biotite is largely altered to greyish sericite (Chapter 4, Fig. 4.28), which is broadly
reflected by the removal of Nb of previously unaltered protolith (Chapter 5, section 5.3).

The REE patterns of sericitised samples clearly demonstrate that bulk mass gain has diluted
the concentration of the REE (Fig. 8.4). The sericitised least-altered biotite-quartz diorite
composition is comparable with mass dilution of 20-40% (Fig. 8.4), similar to the calculated
bulk mass gain. Likewise, the REE pattern of a sericitised potassically-altered biotite-quartz
diorite protolith is in broad agreement with modelled bulk mass gains. The development of a
negative Eu anomaly is indicative of preferential Eu removal in both previously unaltered and
potassically-altered protoliths (Fig. 8.4)

274
Chapter Eight: Discussion

Figure 8.4: REE patterns of sericitised biotite-quartz diorites normalised to volatile-free C1 chondrite (Sun and
McDonough, 1989). “Sericite_U” is the average of sericitised, previously unaltered biotite-quartz diorite (“BD least
alt”), “Sericite_M” and “Sericite_VS” are the averages of moderately and very strongly sericitised, previously
potassically-altered biotite-quartz diorites (“Potassic average”) respectively. Modelled bulk mass gains (grey lines) of
20% and 40% of the biotite-quartz diorite and 60%, 150% and 275% are modelled for bulk mass gain of the average
potassically-altered biotite-quartz diorites.

The abundance of pyrite in these samples is reflected in the net gains of Fe2O3 and S, as well as
the majority of the chalcophile elements. Microprobe analyses of pyrite detected trace Co (up
to 0.57±0.06 wt. %) and Zn (up to 0.26±0.04 wt. %), and chalcopyrite with trace As (up to
0.31±0.02 wt. %) is probable (Fig. 8.5), accounting for the additions of these metals.

Ga, V and Cr are inferred to reside in muscovitic mica, however the trace amounts of these
elements were below the detection limits of EDX analysis.

275
Chapter Eight: Discussion

Figure 8.5: Electron microprobe analyses of pyrite from Cerro Corona (sample C11.JL.122) confirming: (A)
cobaltiferous pyrite containing up to 0.57±0.06 wt.%; (B) pyrite with Zn up to 0.26±0.04 wt.% and sphalerite infill
along fractures; (C) arsenian chalcopyrite infilling fractures with As up to 0.31±0.02 wt.%. Red spots are location of
microprobe analyses.

8.2.4 Argillic alteration


In the argillic domains, illite-smectite and kaolinite-smectite alteration facies have mass gains
of ~25%, whereas dickite-altered samples exhibit dramatic gains of ~150% indicating a
marked difference in processes.

These mass changes are attributed principally to the variable quartz veining/silicification
within the samples and may not be entirely related to the argillic alteration stage, but could
include a pre-existing alteration that was not confidently determined due to the alteration
intensity (e.g. the bulk mass gain due to veining/silicification could have been entirely
introduced in a prior sericitic alteration that could not be recognised). The consistent removal
of Nb in all stages of argillic alteration suggests breakdown of biotite, the dominant Nb-bearing
mineral in equivalent composition porphyries (Parker et al. 1968).

276
Chapter Eight: Discussion

Dickite-dominant alteration displays significant SiO2, Fe2O3, S and LOI additions which may be
attributed to the observed pyrite formation, generation of “smoky” quartz and general
hydration involved in the formation of the clay minerals. The alteration of feldspars to clay can
account for the losses in Na2O and CaO (plagioclase), and K2O and Rb (alkali feldspar,
muscovite):

7.7𝑁𝑎 . 𝐶𝑎 . 𝐴𝑙 . 𝑆𝑖 . 𝑂 + 5𝐻 𝑂 + 10𝐻𝐶𝑙
→ 5𝐴𝑙 𝑆𝑖 𝑂 (𝑂𝐻) + 10.79𝑆𝑖𝑂 + 5.39𝑁𝑎𝐶𝑙( ) + 2.31𝐶𝑎𝐶𝑙 ( )

𝑃𝑙𝑎𝑔𝑖𝑜𝑐𝑙𝑎𝑠𝑒 + 𝑤𝑎𝑡𝑒𝑟 + ℎ𝑦𝑑𝑟𝑜𝑐ℎ𝑙𝑜𝑟𝑖𝑐 𝑎𝑐𝑖𝑑


→ 𝑘𝑎𝑜𝑙𝑖𝑛 + 𝑞𝑢𝑎𝑟𝑡𝑧 + 𝑠𝑜𝑑𝑖𝑢𝑚 𝑐ℎ𝑙𝑜𝑟𝑖𝑑𝑒 + 𝑐𝑎𝑙𝑐𝑖𝑢𝑚 𝑐ℎ𝑙𝑜𝑟𝑖𝑑𝑒

2𝐾𝐴𝑙𝑆𝑖 𝑂 + 2𝐻𝐶𝑙( ) → 𝐴𝑙 𝑆𝑖 𝑂 (𝑂𝐻) + 4𝑆𝑖𝑂 + 2𝐾𝐶𝑙( )

𝐾𝑠𝑝𝑎𝑟 + ℎ𝑦𝑑𝑟𝑜𝑐ℎ𝑙𝑜𝑟𝑖𝑐 𝑎𝑐𝑖𝑑 → 𝑘𝑎𝑜𝑙𝑖𝑛 + 𝑞𝑢𝑎𝑟𝑡𝑧 + 𝑝𝑜𝑡𝑎𝑠𝑠𝑖𝑢𝑚 𝑐ℎ𝑙𝑜𝑟𝑖𝑑𝑒

Calcium released from the destruction of plagioclase may not be completely removed, because
the released CaO may combine with the phosphorus that is added in this alteration step,
resulting in Ca-phosphate precipitation. The addition of Ca-phosphates to the rock can explain
the drastically different REE pattern produced during dickite-dominated alteration in which
REE are strongly added, with greater enrichment in the LREE than the HREE (Fig. 8.6).
Strontium addition is also consistent with apatite precipitation. The large gains in metals
during the dickite-stable alteration implies that the fluids involved at this stage were
metalliferous and it is suggested that they are related to an epithermal overprint (see below),
which is usually more distal to the porphyry intrusion (such as Lepanto epithermal and Far
Southeast porphyry, Hedenquist et al. 1998).

The kaolinite-smectite assemblage is controlled by similar reactions to the dickite-dominant


alteration, although probably occurring at lower temperatures (Chapter 7, Fig. 7.12).
Plagioclase was altered to kaolinite with the removal of CaO and Na2O. The gains in SiO2, Fe2O3,
and S in these facies indicate quartz and pyrite addition, although this could have taken place
in an unrecognised, preceding sericitic alteration stage. CaO removal is mimicked by Sr during
this stage indicating that Ca-phosphate precipitation does not play a role, as clearly indicated
by the REE dilution patterns (Fig. 8.6). Rubidium and Ba follow K2O in that they remain within
error of no change, as most of feldspar is potentially destroyed by pre-existing alteration and
the residence of potassium is within the illite clay structure.

277
Chapter Eight: Discussion

The illite-smectite alteration does not appear to show significant gains and losses, indicating
that this alteration stage did not form from very aggressive or trace metal-rich fluids and that
significant rock fracturing and veining did not take place. The observed breakdown of
plagioclase and relic mafic phases to illite-smectite shows little net mobility of major
components (CaO, Fe2O3, K2O, MgO and P2O5) indicating local re-incorporation into secondary
minerals (K into illite; Ca, Fe, Mg into smectite and carbonates). The net removal of Na2O and
addition of LOI is explained by alteration of plagioclase to calcite and illite, similar to the
reaction suggested for sericitisation of plagioclase, although at lower temperatures. The minor
addition of sulphur may be related to limited pyrite formation, with the Fe2O3 being derived
from the alteration of biotite to illite. The REE show no definitive net change and are consistent
with whole rock dilution of the least-altered biotite-quartz diorite by limited bulk mass gain
(Fig. 8.6). Strontium released from plagioclase destruction appears to be removed and is not
incorporated into calcite or illite. Barium and Rb appear to have different behaviour: Ba is
removed whereas Rb is gained. This suggests that, following the breakdown of plagioclase and
biotite, Ba remained in solution, whereas excess Rb was incorporated into illite. This conflicting
behaviour is not fully understood and should be evaluated in further studies of the mineral
chemistry of these alteration products.

Figure 8.6: REE chondrite-normalised (Sun and McDonough 1989) patterns of average argillically-altered biotite-
quartz diorite compared with least-altered, potassically-altered and propylitically-altered biotite-quartz diorite (“BD
least alt”, Potassic average” and “VS Prop”, respectively). Grey lines indicate 30% bulk mass gain of least-altered
biotite diorite and 20% and 40% bulk mass gain of potassically-altered biotite-quartz diorite.

278
Chapter Eight: Discussion

8.2.5 3D Mass transfer modelling


The underlying weakness in the model of mass transfer in 3D is that only the average of
potassically-altered biotite-quartz diorite was utilised as a representative protolith. There are
multiple overprinting alteration stages and variable alteration intensities, which may blur the
detail of the mass transfer. Furthermore, in the ideal case, field logging should constrain the
precursor alteration types as well as the intensity of each alteration stage. This method of
characterisation would provide a more realistic representation of mass transfer. Nonetheless,
the simplified approach presented in Chapter 5 provides some general insights into bulk mass
and net chemical redistribution during alteration at Cerro Corona.

The predominant area of bulk mass gain is above the 3000 m level and to the north of the
deposit (Fig. 8.7), likely to be due to the intense quartz veining in this region. This corresponds
to the area of greatest net addition of the major components K2O (Chapter 5, Fig. 5.49), a plume
of CaO loss (Chapter 5; Fig. 5.55) and an upward deflection of the central zone of Na2O loss
(Chapter 5; Fig. 5.52). This is inferred to reflect a zone of “secondary” potassic alteration,
potentially relating to an unknown porphyry intrusion at depth (Fig. 8.8). Notably, this zone is
also coincidental with logged areas containing >5% pyrite and >4% magnetite and an “inverted
cup” shaped Cu grade shell. The areas of no bulk mass change (post initial potassic) within the
central and northern areas have copper grades >0.5%, indicating that copper was introduced
during the initial potassic alteration event (Fig. 8.8) and in the secondary potassic event.

279
Chapter Eight: Discussion

Figure 8.7: Cross section across 763300 showing inferred intrusion causing drastic bulk gains in northern area of the
deposit and their coincidence with logged pyrite, magnetite and chalcopyrite, and an inverted cup-shaped Cu grade
shell.
280
Chapter Eight: Discussion

The central region, corresponding to ~60% mass gains is coincident with K2O loss and a
downward deflection of the zone of Na2O loss. This region follows the distribution of chlorite
(Fig. 8.8), a distinctive mineral of the propylitic alteration assemblage (Chapter 4), indicating
strong propylitic alteration in the central region.

281
Chapter Eight: Discussion

Figure 8.8: N-S cross section showing net loss of Na2O and logged areas where chlorite >10%.

282
Chapter Eight: Discussion

8.3 The origin of clay alteration at Cerro Corona


The majority of clay minerals at Cerro Corona occur at shallow levels, predominantly above the
3500m level. This close spatial link to the surface has led to the inference that the incursion of
later meteoric water has caused this alteration, a common feature in porphyry deposits as they
are uplifted and eroded (Sillitoe 2010). Infrared spectroscopy has aided in the differentiation
of clay mineral groups which, when mapped out, have variable distribution. These specific clay
distribution patterns are interpreted along with the findings of the stable isotope study to
provide an understanding of the controls of argillic alteration at Cerro Corona.

Smectite is by far the most abundant clay mineral at Cerro Corona. Smectite-rich assemblages
are most abundant in the eastern and central parts of the host biotite quartz diorite intrusion
where they occur in areas of increased permeability, along structural conduits. This clay
mineral group commonly forms under low temperature, near neutral conditions. Stable
isotope calculations based on the smectites at Cerro Corona reveal that the fluids responsible
for the assemblage were primarily hot (> ~150°C) meteoric waters (Chapter 7, Fig. 7.19). This
is supported by the mass transfer analysis which reveals that smectite-dominant alteration
does not result in drastic addition and removal of major and minor components (Chapter 5,
Fig. 5.38). It is most plausible that the formation of the (illite-)smectite facies occurred in the
waning stages of the hydrothermal system as the Western Cordillera was significantly uplifted
during the Quechua II deformation phase (Chapter 3, section 3.2.2). This period of rapid
regional uplift along the Western Cordillera (Michalak et al. 2016) exhumed the cooling
intrusions and heated local meteoric waters.

Kaolinite is most abundant at surface, focused in the centre of the deposit (Chapter 6, Fig. 6.18)
and is patchy at depth. The kaolinite crystallinity increases with depth consistent with an
increase in formation temperature, which may be interpreted either in terms of a normal
geothermal gradient or heating of the kaolinite-forming fluids by a still warm porphyry stock.
The results of the isotope study suggest that the kaolinite-forming fluids were at ~150°C. These
fluids may be either an equal mixture of magmatic and meteoric waters (Fig. 7.19); or meteoric
waters at ~200°C that have interacted with host rock with a water: rock ratio of ~ 1: 10 whilst
being heated (Fig. 7.17). The mass transfer analysis revealed that kaolinite-smectite alteration
resulted in minor bulk mass gain, with limited changes in major elements (Fig. 5.38), with
minor changes in trace element (Sr removal, due to plagioclase destruction) and LREE removal.

283
Chapter Eight: Discussion

This is indicative of a hydration-dominant alteration process that is perhaps more consistent


with a purely meteoric fluid origin.

Dickite, although only a minor constituent of the argillic alteration assemblages overall, occurs
concentrated in pockets, along the southern margin of the intrusive complex and along
fractures. The mass transfer analysis revealed that dickite-dominated alteration resulted in
significant mass gain and, with significant gains in Fe2O3, P2O5 and S (Chapter 5, Fig. 5.38). The
trace metals are largely added in this alteration assemblage, with significant addition of ore
metals (Fig. 5.42). The REE pattern is significantly modified, potentially due to the formation
of currently unidentified APS minerals. These suggest that there were pulses of acidic, higher
temperature fluids that interacted with the host porphyry along the southern margin of the
intrusion, most consistent with magmatic fluid input in these areas, potentially during the
period of uplift and erosion.

Illite commonly occurs with smectite and/or kaolinite and is the predominant clay at depth.
The stable isotope study indicates that illite-forming fluids were at ~300°C. These fluids can
be modelled as purely magmatic fluids with low magmatic water: rock ratios of ~1:10 (Fig.
7.18). Alternatively, a model involving mixing of a large proportion of (~70-90%) magmatic
fluid with meteoric waters is also plausible (Fig. 7.19). A majority meteoric fluid cannot be
completely disregarded, as heated meteoric waters interacting with ~20-60 times as much
host rock at ~340°C may also fall within the illite-forming fluid composition O-H isotope field
(Fig. 7.18). The IR spectroscopy study indicated there are two groups of illite, based on the
2200nm absorption feature (Chapter 6, Fig. 6.27). The more crystalline illite (Fig. 6.31) is
coincidental with the zone of “secondary” potassic-propylitic transitional alteration and
deeper porphyry emplacement (Fig. 8.8). Whereas the lower-crystallinity illite is more
intrinsically associated with the smectite and kaolinite, at shallower levels.

Although no mass transfer analysis was conducted on purely illite-dominated argillic alteration
assemblages, the illite-smectite alteration appears to have little bulk mass gain and net change
in major or minor elements (Fig. 5.38, Fig. 5.42). This suggests that illite associated with other
clays may be related to meteoric fluids.

Succinctly, it appears that the both illite groups- (the lower-crystallinity illite, associated with
kaolinite and smectite and a higher crystallinity illite group) formed at ~300°C. The higher
crystalline illite is inferred to be transitional to sericitic alteration event in the northern margin
of the deposit, with a higher magmatic fluid component. Whereas, the coexistence of illite with
284
Chapter Eight: Discussion

other clay minerals in the shallower portions of the deposit is interpreted to reflect the
overprinting of the secondary, cooler, meteoric-driven argillic alteration onto this hotter
alteration facies, once the flow of magmatic fluids had ceased.

8.4 Fluid flow model for the genesis of argillic


alteration at Cerro Corona
The most plausible explanation for the overall distribution of clay assemblages is that a well-
developed convection system was established within the Cerro Corona deposit after the main
hypogene mineralisation stage. The fluids involved in the formation of dickite potentially had
a higher magmatic proportion and were primarily controlled by fault conduits. The illite-
forming fluids reflect the waning magmatic-hydrothermal activity, or the onset of an extensive
meteoric-driven convective circulation. As the area was uplifted and eroded and the
geothermal gradient retreated, lower temperature kaolinite-stable fluids, which are usually
more distal, were superimposed onto the deposit, controlled primarily by zones of pre-existing
high permeability. The lowest temperature fluids (~150°C) were predominantly meteoric and
formed the smectite-dominant and kaolinite-dominant assemblages, overprinting downwards
on the system as the geotherms retreated.

This telescoping of argillic alteration assemblages onto one another as well as onto pre-existing
potassic, sericitic and propylitic alteration suggests that there was substantial uplift and/or
erosion occurring at the time of formation, consistent with the Quechua II deformation event
when significant exhumation of the Cajamarca terrain is inferred.

8.5 References
Arnott, A. M. (2003) Evolution of the hydrothermal alteration at the Chuquicamata porphyry
copper system, northern Chile. PhD thesis. Dalhousie University, Nova Scotia, Canada.

Brimhall, G. H., Agee, C., and Stoffregen, R. (1985). The hydrothermal conversion of hornblende
to biotite. Canadian Mineralogist, 23, 369-379.

Castillo, P. R., Janney, P. E. and Solidum, R. U. (1999) Petrology and geochemistry of Camiguin
Island, southern Philippines: insights to the source of adakites and other lavas in a complex arc
setting. Contributions to Mineralogy and Petrology. 134 (1), 33-51.

285
Chapter Eight: Discussion

Chambefort, I., Dilles, J. D., Longo, A. A. (2013) Amphibole Geochemistry of the Yanacocha
Volcanics, Peru: Evidence for Diverse Sources of Magmatic Volatiles Related to Gold Ores.
Journal of Petrology. 54 (5), 1017-1046.

Chiaradia, M., Merino, D. and Spikings, R. (2009) Rapid transition to long-lived deep crustal
magmatic maturation and the formation of giant porphyry-related mineralization (Yanacocha,
Peru). Earth and Planetary Science Letters. 288 (3), 505-515.

Davies, R. C. I. (2002) Tectonic, magmatic and metallogenic evolution of the Cajamarca mining
district, Northern Peru. James Cook University.

Davies, R. C. and Williams, P. J. (2005) The El Galeno and Michiquillay porphyry Cu-Au-Mo
deposits: geological descriptions and comparison of Miocene porphyry systems in the
Cajamarca district, northern Peru. Mineralium Deposita. 40 (5), 598-616.

Goldsmith, J. R. (1950) Gallium and germanium substitutions in synthetic feldspars. The


Journal of Geology. 518-536.

Hollister, V. F. and Sirvas, E. B. (1974) The Michiquillay porphyry copper deposit. Mineralium
Deposita. 9 (3), 261-269.

Idrus, A., Kolb, J. and Meyer, F. M. (2009) Mineralogy, Lithogeochemistry and Elemental Mass
Balance of the Hydrothermal Alteration Associated with the Gold-rich Batu Hijau Porphyry
Copper Deposit, Sumbawa Island, Indonesia.Resource Geology. 59 (3), 215-230.

James, J. (1998) Geology, alteration, and mineralization of the Cerro Corona porphyry copper-
gold deposit, Cajamarca Province, Peru. Master’s thesis. University of British Columbia.

Li, J., Li, G., Qin, K., Xiao, B., Chen, L. and Zhao, J. (2012), Mineralogy and Mineral Chemistry of
the Cretaceous Duolong Gold-Rich Porphyry Copper Deposit in the Bangongco Arc, Northern
Tibet. Resource Geology, 62, 19–41.

Longo, A. A. (2005) Evolution of volcanism and hydrothermal activity in the Yanacocha mining
district, northern Peru. PhD thesis. Oregon State University.

Longo, A. A., Dilles, J. H., Grunder, A. L. and Duncan, R. (2010) Evolution of Calc-Alkaline
Volcanism and Associated Hydrothermal Gold Deposits at Yanacocha, Peru. Economic
Geology. 105 (7), 1191-1241.
286
Chapter Eight: Discussion

Macpherson, C. G., Dreher, S. T. and Thirlwall, M. F. (2006) Adakites without slab melting: high
pressure differentiation of island arc magma, Mindanao, the Philippines. Earth and Planetary
Science Letters. 243 (3), 581-593.

Michalak, M. J., Hall, S. R., Farber, D. L., Audin, L., and Hourigan, J. K. (2016). (U-Th)/He
thermochronology records late Miocene accelerated cooling in the north-central Peruvian
Andes. Lithosphere, 8(2), 103-115.

Prokof’ev, V. Y., Naumov, V. B. and Dorofeeva, V. A. (2016) Gallium concentration in natural


melts and fluids. Geochemistry International. 54 (8), 691-705.

Prokof’ev, V. Y., Naumov, V. B. and Selektor, S. L. (2014) Content of gallium in melt and fluid
inclusions. Asian Current Research on Fluids Inclusions.

Ridolfi, F., Renzulli, A., and Puerini, M. (2010). Stability and chemical equilibrium of amphibole
in calc-alkaline magmas: an overview, new thermobarometric formulations and application to
subduction-related volcanoes. Contributions to Mineralogy and Petrology, 160(1), 45-66.

Rosenbaum, G., Giles, D., Saxon, M., Betts, P. G., Weinberg, R. F. and Duboz, C. (2005) Subduction
of the Nazca Ridge and the Inca Plateau: Insights into the formation of ore deposits in
Peru. Earth and Planetary Science Letters. 239 (1), 18-32.

Siahcheshm, K., Calagari, A. A., Abedini, A. and Sindern, S. (2014) Elemental mobility and mass
changes during alteration in the Maher-Abad porphyry Cu–Au deposit, SW Birjand, Eastern
Iran. Periodico Di Mineralogia. 83 (1), 55-76

Sillitoe, R. (1997) Comments on the geological model for the Cerro Corona porphyry copper-
gold prospect and exploration potential elsewhere in the Hualgayoc district, Northern Peru.
RGC Exploration Pty Limited.

Sillitoe, R. H. (2010) Porphyry Copper Systems. Economic Geology. 105 (1), 3-41.

Stanton, R,L. (1994) Ore elements in arc lavas. Oxford University Press on Demand.

Sun, S. S. and McDonough, W. F. (1989) Chemical and isotopic systematics of oceanic basalts:
implications for mantle composition and processes. Geological Society, London, Special
Publications. 42 (1), 313-345.

287
Chapter Eight: Discussion

Ulrich, T. and Heinrich, C. A. (2002) Geology and Alteration Geochemistry of the Porphyry Cu-
Au Deposit at Bajo de la Alumbrera, Argentina. Economic Geology. 97 (8), 1865-1888.

Vasyukova, O. V., Kamenetsky, V. S., Goemann, K., and Davidson, P. (2013). Diversity of primary
CL textures in quartz from porphyry environments: implication for origin of quartz eyes.
Contributions to Mineralogy and Petrology, 166(4), 1253-1268.

Wilkinson, J. J., Chang, Z., Cooke, D. R., Baker, M. J., Wilkinson, C. C., Inglis, S., Chen, H. and
Gemmell, J. B. (2015) The chlorite proximitor: A new tool for detecting porphyry ore deposits.
Journal of Geochemical Exploration. (152) 10-26.

Williamson, B., Herrington, R. and Morris, A. (2016) Porphyry copper enrichment linked to
excess aluminium in plagioclase. Nature Geoscience. 9, 237–241.

288
Chapter Nine: Conclusions

9 Chapter Nine: Conclusions

9.1 Development of mineral-bearing arc volcanism in


northern Peru
Based on literature reviews, the Hualgayoc igneous district was placed within the broader
context of the North Peruvian arc. It lies on the eastern margin of the Western Peruvian Trough
(WPT), closer to a major Andean-trending fault-the Puntre Fault than the Yanacocha
epithermal district to the Southwest Fig. 3.1 and Fig. 3.5). This is the major structural weakness
that formed during the Cretaceous period and controlled sedimentation within the WPT
(Scherrenberg et al. 2014). This normal fault is reactivated during subsequent phases of
compression and may explain the variable rates of uplift and erosion within the wider
Cajamarca district, with higher rates of uplift occurring along proximal to the reactivated fault,
which had not been previously considered.

Gradational crustal thickening during the period leading up to the emplacement of Calipuy
Volcanic Group is evidenced by high pressure fractional crystallisation along with melting
(Chiaradia et al. 2009) and assimilation of the lower crust and the development of an adakite-
like signature. During this stage the magmatic arc shifts eastward.

The development of shallow level magma chambers is evidenced by the development of


porphyritic textures comprising predominantly plagioclase, biotite and amphibole
phenocrysts. The emplacement of porphyry intrusions appears to have a structural control,
especially proximal to the Puntre Fault. This is seen in the Michiquillay (Hollister et al. 1974)
and El Galeno (Davies et al. 2005) intrusions. Trenchward of these intrusions is a belt of
epithermal mineralisation (the Yanacocha District), considered to represent shallow degassing
of deeper magmas where there were presumed lower rates of uplift and erosion.

The Hualgayoc district records a sequence of early sill emplacement within the hinge of a NW-
SE trending anticline. This is followed by porphyry emplacement (Cerro Jesus, Cerro
Coymolache and Cerro Corona) and synchronous Cu-Pb-Zn-Ag stratiform replacement
mineralisation (Macfarlane et al. 1994) preferentially along permeable horizons. The
hydrothermal alteration within the district falls within a ~2 Ma window. This is followed by a
felsic intrusion ~9 Ma (Cerro Hualgayoc) and presumably the undated Las Gordas rhyolite.

289
Chapter Nine: Conclusions

The intrusive history of the Cerro Corona deposit is constrained by crosscutting relationships,
which suggest multi-phase intrusions of biotite quartz diorite porphyries, a late hornblende
quartz diorite porphyry and andesite dykes. The deposit is volumetrically dominated by biotite
quartz diorites in which the major and trace-element chemistry of least altered pre- and post-
ore mineralisation biotite quartz diorite porphyries are indistinguishable (Fig. 4.21 and Fig.
4.22).

The magnesio-hornblende from the hornblende quartz diorite porphyry crystallized from
magmas that were more oxidised, lower temperature and less hydrous, than the pargasite of
the late andesite dykes as reflected in the amphibole geothermobarometry (Fig. 4.45). This
sequence is reflected within the Yanacocha district in which hydrous basaltic melts are
emplaced late in the intrusive history (Chambefort et al. 2013), coincident with the subducted
buoyant “Inca Plateau” beneath the Cajamarca district at ~12-8 Ma (Rosenbaum et al. 2005).

The subduction of the Inca Plateau is related to flat slab subduction which is considered to
prevent recharging of magmatic reservoirs within this section of the magmatic arc. The final
magmatic cycle within this segment of the arc is the extrusion of hydrous oxidised felsic
pyroclastics of the Negritos Formation (Longo 2005).

9.2 Implications from mass transfer models of


hydrothermal alteration in a porphyry deposit
A primary aim of this thesis was to resolve what the overall bulk mass changes and net gains
and losses of major, minor and trace components are during various stages of hydrothermal
alteration; where and how these patterns inform our understanding of the evolution of
porphyry hydrothermal systems.

The similar geochemistries of the pre- and post-ore mineralisation biotite quartz diorites
provided an ideal opportunity in which the mass transfer of components could be evaluated,
as the protolith compositions are indistinguishable from one another. The ISOCON method was
adopted to evaluate the mass changes. This method is significantly improved upon by the
incorporation of the propagation of errors, which allows for increased confidence in the
results.

Mass gain is ubiquitous within all alteration stages, with significant gains in SiO2 as reflected
by quartz veining. Potassic and sericitic alteration have the largest mass gains and net gains in

290
Chapter Nine: Conclusions

SiO2. This is reflected in hand sample and drillcore logs in which these alteration stages show
significant stockwork and sheeted veining.

Significant bulk mass gain in the northern section of the Cerro Corona deposit corresponds to
an upward deflection of net Na2O loss coupled with K2O gain and a general “mushroom” of CaO
loss. These net changes correspond to potassic alteration, seen in detailed mass transfer
analyses of hand samples. This zone is thus inferred to be a second potassic alteration stage
from a deeper degassing magmatic body, perhaps a smaller “pencil porphyry” at depth? The
apex of this potassic alteration corresponds to an inverted cup shape of Cu and Au grade
distribution, similar to the distribution of grade seen in Bingham (Gruen et al. 2010), albeit
smaller.

These mass transfer models indicate that potassic alteration is dominated by large bulk mass
gains, predominantly by net gains of SiO2, K2O, Na2O and FeO (total) addition. This
geochemistry is reflected in typical potassic alteration assemblages dominated by k-silicates
and is associated with quartz and magnetite stockwork veining. These mass changes are
similar to those seen in potassic alteration within porphyry deposits (Fig. 9.1). More subtle
trace element mobility is reflected by mineralogical changes, such as Cr and V removal during
amphibole alteration to biotite. The LREE are more readily remobilised than HREE during
potassic alteration, causing a general flattening of the REE pattern. The gain of REE’s within
propylitic alteration suggests that the REE are redeposited within this stage, presumably
within apatite, calcite or plagioclase. This alteration is a common process in porphyry systems,
similar trends in major oxide gains and losses are seen in which K2O is gained at the expense
of CaO and/or Na2O (Fig. 9.1). Broadly speaking, the bulk mass change appears largely
controlled by SiO2 related to the abundance of stockwork quartz veins.

291
Chapter Nine: Conclusions

Figure 9.1: Calculated net change during potassic alteration in other deposits. Bulk mass change and SiO2 are plotted
on the left axis, as their changes are larger than other components read off the right hand axis. Abbreviations are in
the order of deposit, alteration and lithology where: BdlA., Arg. = Bajo de la Alumbrera porphyry Cu-Au, Argentina
(Ulrich and Heinrich 2002); B.H., Ind. = Batu Hijau porphyry Cu-Au, Indonesia (Idrus et al. 2009); Chu, Chi. =
Chuquicamata porphyry Cu-Au, Chile (Arnott 2003); Du., T., = Duobuza porphyry Cu-Au, Tibet (Li et al. 2011); M-A,
Ir.= Maher-Abad porphyry Cu-Au, Iran (Siahcheshm et al. 2014); Pot. Alt= Potassic alteration; Early P3= Early P3
porphyry; Late P3=Late P3 porphyry; NW Por=NW porphyry; E Por=East porphyry; Unk.=lithology uncertain; Late
Grd.= Late Granodiorite; QM= Quartz monzonite; And. Clast= Andesitic volcaniclastics; I.T.= Intermediate tonalite;
EQD= Equigranular quartz diorite; Y.T=Young tonalite.

The propylitic alteration stage is characterised by the presence of chlorite, calcite and quartz
alteration of mafic minerals along with the alteration of feldspar to illite. The net gains of CaO,
FeO (total) and LOI is reflected in the overall hydration of these rocks, by increasing abundance
of hydrous minerals along with calcite and magnetite. Cr and V removal during potassic
alteration may be added to the rock during this stage, presumably hosted within the magnetite
structure. Strontium removal during this confirms the observation that proximal propylitic
alteration has lower total Sr in chlorite (Wilkinson et al. 2015).

The sericitic alteration assemblage comprises of dominantly sericite, quartz and pyrite. This
alteration stage does not cause significant change in K2O content, but rather the removal of
Na2O by feldspar destruction and overall hydration of the rock by LOI (water?) addition. The
alteration of feldspars to sericite buffers the KCl/HCl activity of the fluid, which, based on
292
Chapter Nine: Conclusions

mineral stability, is about 1-3 (Fig. 1.7). The metal gains of this alteration provide strong
evidence for a magmatic fluid, which buffered by sericite-k-feldspar alteration with decreasing
temperature. This leads to radical drop in divalent metal sulphide solubility (Crerar et al. 1985)
during pyrite precipitation. This pyrite precipitation and base metal content (Pb and Zn) is
typical for roots of epithermal zones with this associated with this sericitisation.

Argillic alteration does not contribute as large a mass transfer as other alteration assemblages.
The dickite-dominant alteration samples show significant gains in phosphate, which are
inferred to cause dramatic gains in the REE’s.

Simplistically, the mass transfer of major components can be seen in figure 8.1. The potassic
alteration typically shows increase in both K/Al and (Na+K)/Al, subsequent sericitic and
propylitic alteration removes Na2O, resulting in a lower (Na+K)/Al over a similar interval K/Al.
Argillic alteration removes both K2O and Na2O resulting in a low K/Al and (Na+K)/Al.

The intriguing distribution of gallium during hydrothermal alteration stage, may be an


important exploration tool. It appears to be significantly added during potassic and sericitic
alteration stages, but relatively conserved in propylitic and argillic alteration, thus Ga
anomalies may reflect hydrothermal alteration related to a deeper magmatic body.

9.3 Inferences from stable isotope and infrared


spectroscopy of clay minerals
Another primary objective of this thesis was to evaluate the clay mineral assemblages present
in the argillic alteration zones and could different alteration facies be defined that reflect
different processes or different fluid sources.

In this regard, the overall distribution of argillic alteration shows that the most intense argillic
alteration occurs up to a depth of 400 m from the current surface level, with deeper clay
mineral distribution being predominantly controlled by structures.

Based on IR spectroscopy, illite-dominant argillic alteration commonly occurs as a mixture


with smectite and/or kaolinite. This is contradictory to the isotope study which demonstrates
that illite is formed from high temperature (~350°C), predominantly magmatic fluids, kaolinite
from a mixture of meteoric and magmatic fluids at intermediate temperatures (~200°C) and
smectite is dominated by lower temperature (~150-200°C), predominantly meteoric fluids. A

293
Chapter Nine: Conclusions

plausible explanation for this may be that the argillic alteration mixtures occur due to
overprinting of events of cooler meteoric fluids during the retreating of the geothermal
gradient, resulting in an inverted clay mineral distribution (i.e. telescoping). This telescoped
alteration is common in porphyry deposits where uplift and erosion results in argillic
alteration overprinting earlier higher temperature alteration assemblages.

The distribution of illite-dominant argillic alteration appears to be somewhat patchy at ~3600


m level (Fig. 6.26). This alteration appears to have the largest magmatic component,
comprising of up to 70% magmatic fluid and is considered the earliest argillic alteration with
alteration fluids at ~350 °C.

With retreating geothermal gradient and larger influx of meteoric fluids, kaolinite alteration
telescopes and overprints earlier, higher temperature illite-dominant alteration. The smectite-
dominant clay contributes the most significant proportion of clay mineral within the Cerro
Corona deposit (Fig. 6.13), predominantly of the montmorillonite-type (Fig. 6.32). The
distribution of smectite-dominant argillic alteration is thought to be controlled primarily by
permeability- such as intrusive contacts. Stable isotope values of smectite-dominant clays
suggest these clays formed from lower temperature meteoric waters (~150-200°C), perhaps
at during the end of the waning hydrothermal system as the area was uplifted and eroded.

The exposure of sulphides to air and water results in the generation of sulphuric acid. This
generation of sulphuric acid, lowers the pH of the water and thus result in lower (KCl+K+) /
(HCl+H+). This lower activity ratio causes kaolinite-dominant argillic alteration rather than
smectite-dominant alteration at equivalent temperatures (Fig. 9.2 pathway A to B versus E to
F). The kaolinite-dominant alteration in the centre of the pit may be related to this ground-
water acidification, as the distribution of Fe-oxides show abundant hematite and limonite
within a core of the kaolinite-dominant alteration (Fig. 9.2 pathway A to B). The northern
margin shows less kaolinitisation suggesting that meteoric fluids may have been more buffered
(Fig. 9.2, pathway F to G).

294
Chapter Nine: Conclusions

Figure 9.2: Cross section with clay mineral distribution pattern with interpretations of their distribution based on
temperature and acidity.

295
Chapter Nine: Conclusions

Hydrothermal alteration at Cerro Corona is comparable with typical telescoped porphyry


copper deposits (Gustafson and Hunt 1975, Sillitoe 2010). Early potassic alteration, associated
with quartz stock work ± magnetite ± chalcopyrite is crosscut by propylitic and sericitic
alteration. The transition from potassic to propylitic alteration is considered to cause ore
mineralisation. This is crosscut by intrusions of barren biotite quartz diorite intrusions. The
inferred intrusion of a deeper porphyry is suggested to result in a secondary potassic and
alteration zone in the northern margin. Subsequent sericitic and argillic alteration appears to
be predominantly controlled by relative permeability within the deposit. Further work is
required to understand the overall permeability changes in porphyry environment to further
constrain the fluid pathways controlling alteration.

9.4 Future work


Numerous outstanding questions still exist, outside the scope of this study, some of which are
outlined below:

9.4.1 Regional and local geology


1) What is the timing of the monoclinal folds seen in the Llama and Huambos Volcanics?
2) Timing of regional post-intrusion mineralisation?
a. What is the relationship between the three main fracture networks associated
with polymetallic mineralisation in the Hualgayoc District?
b. Are these related to porphyry emplacement or are they related to a later
orogenic event (such as the Quechua III?)
3) What is the geochemical composition of the Hualgayoc intrusions? Do they follow a
similar geochemical evolution as the magmatic sequence in Yanacocha and elsewhere
in the Cajamarca district? Are these intrusions linked to the same parental melt and if
so at what depth does this batholith exist? If so, is the degassing of this parental
magmatic chamber responsible for local mantos formation?
4) Why does the Tantahuatay have such an intense advanced argillic (quartz-pyrite-
pyrophyllite) alteration and a thick (up to 200-300meters) gussano texture, whereas
Cerro Corona has no alunite or pyrophyllite?
5) What is the chronology of Andean and transandean faults? What is the displacement
along these faults? What is the relationship of these faults with the Hualgayoc
Anticline?

296
Chapter Nine: Conclusions

6) Is the contact between the Corona quartz diorite and the Pariatambo and the QD1 and
Chulec formations the same as the sharp contact between the Yumagual and quartz
diorite? How different is the contact in the north versus the south? How extensive are
permeable horizons in the Chulec and Pariatambo formations? What is the extent of
the mineralisation in the North?

9.4.2 Alteration geochemistry


1) Compare the mass transfer in whole rocks to fluids considered responsible for this
alteration, via the chemical analysis of fluid inclusions.
2) Comparison of mass transfer in other alteration assemblages in other hydrothermally
altered rocks.

9.4.3 Infrared spectroscopy


1) Incorporate fault maps and proximity to intrusive contacts to understand the link of
clay mineral distribution to permeability.
2) Compare the results of clay mineral distribution to satellite images of the area to
identify hydrothermal alteration within the larger context of the district.
3) Compare the vertical clay mineral distribution to deposits nearby to understand
regional the level of uplift and overprinting alteration on a regional scale.

9.4.4 Oxygen and hydrogen isotopes


1) The use of internal fractionation of oxygen between different sites within the kaolinite
and illite clay structure appears to provide reasonable estimates of the temperatures
of formation. This method should be further developed in combination with
precipitation experiments, whereby clays are formed at specific temperatures with
hydrothermal fluids, such kaolinite precipitation experiments (Huertas et al. 1999).
2) Can the internal fractionation of oxygen be applied to kaolinite forming in other
deposits? Analyses of internal fractionation of oxygen should be applied to kaolinite
forming in other environments, from normal weathering processes.

297
Chapter Nine: Conclusions

9.4.5 Generic model


1) The relationship of post-mineralisation structures should be evaluated and
incorporated into the geological model.
2) What is the relationship of the post-mineralisation biotite quartz diorite intrusions to
one another?
3) Are the nearby mantos related to magmatic fluids flowing through permeable
horizons? It seems plausible given the proximity of these deposits to intrusions and fit
with current porphyry models (Sillitoe 2010). Incorporation of the local stratigraphy
may aid in exploration of skarn-type mineralisation along more permeable horizons.

9.5 References
Arnott, A. M. (2003) Evolution of the hydrothermal alteration at the Chuquicamata porphyry
copper system, northern Chile. PhD thesis. Dalhousie University, Nova Scotia, Canada.

Chambefort, I., Dilles, J. D., Longo, A. A. (2013) Amphibole Geochemistry of the Yanacocha
Volcanics, Peru: Evidence for Diverse Sources of Magmatic Volatiles Related to Gold Ores.
Journal of Petrology. 54 (5), 1017-1046.

Chiaradia, M., Merino, D. and Spikings, R. (2009) Rapid transition to long-lived deep crustal
magmatic maturation and the formation of giant porphyry-related mineralization (Yanacocha,
Peru). Earth and Planetary Science Letters. 288 (3), 505-515.

Crerar, D., Wood, S., Brantley, S., and Bocarsly, A. (1985). Chemical controls on solubility of ore-
forming minerals in hydrothermal solutions. The Canadian Mineralogist, 23(3), 333-352.

Davies, R. C. and Williams, P. J. (2005) The El Galeno and Michiquillay porphyry Cu-Au-Mo
deposits: geological descriptions and comparison of Miocene porphyry systems in the
Cajamarca district, northern Peru. Mineralium Deposita. 40 (5), 598-616.

Gruen, G., Heinrich, C. A. and Schroeder, K. (2010) The Bingham Canyon Porphyry Cu-Mo-Au
Deposit. II. Vein Geometry and Ore Shell Formation by Pressure-Driven Rock Extension.
Economic Geology. 105 (1), 69-90.

Gustafson, L. B. and Hunt, J. P. (1975) The porphyry copper deposit at El Salvador,


Chile. Economic Geology. 70 (5), 857-912.

298
Chapter Nine: Conclusions

Hollister, V. F. and Sirvas, E. B. (1974) The Michiquillay porphyry copper deposit. Mineralium
Deposita. 9 (3), 261-269.

Huertas, F. J., Fiore, S., Huertas, F., & Linares, J. (1999). Experimental study of the hydrothermal
formation of kaolinite. Chemical Geology, 156(1), 171-190.

Idrus, A., Kolb, J. and Meyer, F. M. (2009) Mineralogy, Lithogeochemistry and Elemental Mass
Balance of the Hydrothermal Alteration Associated with the Gold-rich Batu Hijau Porphyry
Copper Deposit, Sumbawa Island, Indonesia.Resource Geology. 59 (3), 215-230.

Li, J., Li, G., Qin, K., Xiao, B., Chen, L. and Zhao, J. (2012), Mineralogy and Mineral Chemistry of
the Cretaceous Duolong Gold-Rich Porphyry Copper Deposit in the Bangongco Arc, Northern
Tibet. Resource Geology, 62, 19–41.

Longo, A. A. (2005) Evolution of volcanism and hydrothermal activity in the Yanacocha mining
district, northern Peru. PhD thesis. Oregon State University.

Macfarlane, A. W., Prol-Ledesma, R. and Conrad, M. E. (1994) Isotope and Fluid Inclusion
Studies of Geological and Hydrothermal Processes, Northern Peru. International Geology
Review. 36 (7), 645-677.

Rosenbaum, G., Giles, D., Saxon, M., Betts, P. G., Weinberg, R. F. and Duboz, C. (2005) Subduction
of the Nazca Ridge and the Inca Plateau: Insights into the formation of ore deposits in
Peru. Earth and Planetary Science Letters. 239 (1), 18-32.

Scherrenberg, A.F., Holcombe, R.J., and Rosenbaum, G. (2014) The persistence and role of basin
structures on the 3D architecture of the Marañón Fold-Thrust Belt, Peru. Journal of South
American Earth Sciences 51, 45-58.

Siahcheshm, K., Calagari, A. A., Abedini, A. and Sindern, S. (2014) Elemental mobility and mass
changes during alteration in the Maher-Abad porphyry Cu–Au deposit, SW Birjand, Eastern
Iran. Periodico Di Mineralogia. 83 (1), 55-76

Sillitoe, R. H. (2010) Porphyry Copper Systems. Economic Geology. 105 (1), 3-41.

Ulrich, T. and Heinrich, C. A. (2002) Geology and Alteration Geochemistry of the Porphyry Cu-
Au Deposit at Bajo de la Alumbrera, Argentina. Economic Geology. 97 (8), 1865-1888.

299
Chapter Nine: Conclusions

Wilkinson, J. J., Chang, Z., Cooke, D. R., Baker, M. J., Wilkinson, C. C., Inglis, S., Chen, H. and
Gemmell, J. B. (2015) The chlorite proximitor: A new tool for detecting porphyry ore deposits.
Journal of Geochemical Exploration. (152) 10-26.

300

You might also like