Lithogeochemical Analysis of the Heath Steele E Zone Volcanogenic Massive Sulphide Deposit, Bathurst Mining Camp, by

Josue Jimenez-Gonzalez

B.Sc. in Geological Engineering, National Polytechnic Institute, 2018

A Thesis, Submitted in Partial Fulfillment of the Requirements for the Degree of

Masters of Sciences

in the Graduate Academic Unit of Earth Sciences

Supervisor: David R. Lentz, Ph.D., Department of Earth Sciences

Examining Board: Bruce Broster, Ph.D., Department of Earth Sciences

Michelle Gray, Ph.D., Faculty of Forestry and Environmental Management

This thesis is accepted by the Dean of Graduate Studies

THE UNIVERSITY OF NEW BRUNSWICK

August, 2020

©Josue Jimenez-Gonzalez, 2020

Abstract

The Heath Steele E zone Zn-Pb-Cu-Ag volcanogenic massive sulphide deposit lies in the Heath Steele belt in the Bathurst Mining Camp, northeast New Brunswick.

The Heath Steele E zone deposit is hosted mainly by felsic volcanic and related volcano- sedimentary rocks of the Nepisiguit Falls Formation (Tetagouche Group), which were deposited in the Tetagouche-Exploits back-arc basin in the Middle . The host sequence is affected by locally intense, deposit-related hydrothermal alteration, and polyphase deformation and mid- to upper-greenschist grade regional metamorphism related to inclusion of the host sequence in the Brunswick Complex.

Numerous geological events have affected the E zone deposit and the host rocks, which have complicated a proper stratigraphic interpretation in the area. Furthermore, the similarity among the various volcano-sedimentary units precludes confident unit correlation among adjacent drill cores on the basis of macro-scale observations alone.

For this reason, chemostratigraphy is employed to discriminate among the various volcano-sedimentary units, and for assessment of deposit related hydrothermal alteration using a portable X-ray fluorescence spectrometry (pXRF) as the main tool. The pXRF is a useful analytical tool for acquiring high-quality results in real-time with a level of resolution that surpasses most other techniques, thereby providing at least 30 potential variables for use in chemostratigraphic characterization and correlation.

The pXRF analysis of eight drill cores of the study area facilitated: 1) the construction of discrimination diagrams that show that the host felsic volcanic rocks in the E zone deposit are rhyodacite/dacite and rhyolite with a tholeiitic magmatic affinity;

ii consistent with an intracontinental back-arc environment, 2) the identification of geochemically distinct rock units in this case, a clear difference between structural hanging wall and footwall was recognized, and 3) qualitative characterization of hydrothermal alteration in the footwall (mainly chlorite-carbonate and chlorite-pyrite- sericite assemblage) and the hanging wall (K-feldspar-sericite and sericite-chlorite- pyrite assemblage).

iii

Acknowledgements

First of all, a special thanks to Dr. David Lentz for continued guidance and direction throughout the project. I wish to acknowledge the economic, logistical, and technical support given by Trevali Mining Corporation. I want to express my gratitude to Dr. James Walker for his valuable time invested in this project, and for his always punctual observations to improve this work. I would like to thank all the staff of the New

Brunswick Geological Survey, Bathurst branch for the support during my fieldwork seasons, access to the Madran drill core storage facility and for the use of their pXRF.

Thanks to Tim Kingsley and Peter Hall for the facilities provided during my fieldwork at the Caribou mine and for making me feel part of the team. I would like to thank my internal and external examiner for the time to review this work and for their specific observations to improve it. I thank the Department of Earth Sciences for giving me the opportunity to be part of the UNB community. Finally, I want to thank my family and everyone who has been part of this project and in one way or another supported me to achieve this goal.

iv

Table of Contents Abstract ...... ii

Acknowledgements ...... iv

Table of Contents ...... v

List of Figures ...... vii

List of Tables ...... x

Chapter 1 ...... 1

Introduction ...... 1

1.1 General Statement ...... 1

1.2 Objectives ...... 2

1.3 Justification ...... 3

1.4 Analytical methods ...... 4

1.4.1 Whole-rock lithogeochemistry and pXRF analytical methods ...... 5

1.4.2 Statistical Analysis ...... 8

1.5 Exploration History ...... 11

1.6 Previous (scientific) work ...... 13

Chapter 2 ...... 16

Geology of the Bathurst Mining Camp ...... 16

2.1 Tectonic setting ...... 16

2.2 Stratigraphy of the Bathurst Mining Camp ...... 19

2.3 Volcanogenic Massive Sulphide Deposits in the BMC ...... 25

2.3.1 Review of VMS systems ...... 25

2.3.2 Massive Sulphide Deposits at the BMC ...... 31

Chapter 3 ...... 34

Geology of the Heath Steele area ...... 34

v

3.1 Geologic setting ...... 34

3.2 Heath Steele deformation ...... 37

3.3 Heath Steele E zone deposit ...... 39

Chapter 4 ...... 46

Lithogeochemical interpretation ...... 46

4.1 Rock Type classification ...... 47

4.2 Magmatic affinity ...... 49

4.3 Tectonic settings ...... 51

Chapter 5 ...... 54

Chemostratigraphy ...... 54

5.1 Chemostratigraphic analysis of the felsic volcanic rocks ...... 54

5.2 Exhalites (iron formation) analysis ...... 63

Chapter 6 ...... 72

Hydrothermal Alteration trends ...... 72

Chapter 7 ...... 80

Conclusions and Recommendations ...... 80

Conclusions ...... 80

Recommendations ...... 83

References ...... 85

Appendix A ...... 109

Appendix B ...... 114

Appendix C ...... 116

Appendix D ...... 127

Appendix E ...... 131

Appendix F ...... 135

Curriculum Vitae

vi

List of Figures

Figure 1.1. Lithotectonic divisions of the northeastern Appalachian and simplified geology map of New Brunswick…………………………………………………….. 1

Figure 1.2. Portable XRF correction curves for selected elements used in this project………………………………………………………………………………... 10

Figure 2.1. Simplified tectonic map of Brunswick Complex.………………………. 16

Figure 2.2. Schematic block diagrams illustrating stages of the tectonic and petrogenetic evolution of the Tetagouche-Exploits back-arc basin in four time intervals………………………………………………………………………………. 18

Figure 2.3. Simplified geological map of the Bathurst Mining Camp………………. 21

Figure 2.4. Schematic geological model for the formation of VMS deposits…….…. 27

Figure 2.5. Distribution and characteristic mineralogical assemblages of alteration facies in a typical VMS system …………………………………….……………….. 29

Figure 2.6. Idealized Kuroko-type VMS deposit …….……………………………… 30

Figure 3.1. Geological map of the central part of the Heath Steele area showing location of B, E, F, and ACD zones……………………………………………….… 35

Figure 3.2. Generalized stratigraphic column of the Tetagouche and the Miramichi groups in the Heath Steele area…………………………………………...…………. 36

Figure 3.3. Map of the Heath Steele E zone showing drill core locations in black circles and section line A-A’………………………………………………………… 39

Figure 3.4. Schematic cross sections A-A’ of the Heath Steele E zone deposit……... 40

Figure 3.5. Zn-Pb-Cu base metal classification scheme of E zone VMS deposit...... 41

Figure 3.6. Photograph of volcanic rock at the footwall of the E zone deposit…….... 42

vii

Figure 3.7. Photograph of quartz-feldspar crystal tuff with sub-hedral to euhedral pink K-feldspar and anhedral phenocrysts of quartz………………………………… 42

Figure 3.8. Photograph of laminated siltstone (ash?) with sericite-chlorite alteration. 43

Figure 3.9. Photograph of volcanic rock at the middle-section……………………… 44

Figure 3.10. Photograph of iron formation with breeched texture……………..……. 44

Figure 3.11. Photograph of quartz-feldspar crystal tuff of the hanging………..….… 45

Figure 4.1. Nb/Y vs Zr/TiO2 discrimination diagram applied to data collected from the Heath Steele E zone deposit…………………………………………………..….. 48

Figure 4.2. Y/TiO2 versus Zr/TiO2 discrimination diagram for magmatic affinity….. 50

Figure 4.3. Nb versus Y tectonic discrimination diagram from volcanic and related sedimentary rocks of the Heath Steele E zone……………………………………..… 52

Figure 4.4. Rb versus Y+Nb tectonic discrimination diagram from volcanic and related sedimentary rocks of the Heath Steele E zone……………………………..… 53

Figure 5.1. Zr/TiO2 average signatures per unit with pXRF analyses performed on samples taken from drill cores of the Heath Steele E zone………………….………. 57

Figure 5.2. TiO2 versus Zr plot demonstrates that each of the alteration lines forms a distinct cluster of samples………………………………………………………….. 60

Figure 5.3. Zr/Al2O3versus TiO2/Al2O3immobile elements ratio-ratio plots showing distinct clusters of population of samples……………………………………………. 61

Figure 5.4. Zr/TiO2 signatures with pXRF analyses performed on samples taken from drill cores of the Heath Steele E zone………………………………………….. 62

Figure 5.5. Simplified cross sections of volcanogenic massive sulphide deposits showing different types and morphologies of exhalites…………………………..…. 64

Figure 5.6. Ca-Fe-Mn discrimination diagram…...………………………………….. 65

viii

Figure 5.7. Fe/Mn versus Zn+Pb discrimination diagram..…….……………………. 66

Figure 5.8. Al-Fe-Mn discrimination diagram of Boström…………………………. 67

Figure 5.9. Plot of Fe/Ti versus Al/(Al+Fe+Mn) for whole-rock compositions of exhalite samples from the Heath Steele E zone…………………………………….... 68

Figure 5.10. MnO versus Fe/Mn discrimination diagram for clastic and hydrothermal sedimentary rocks ……………………………………..……………… 69

Figure 5.11. Schematic evolution of the Heath Steele E zone……………………….. 71

Figure 6.1. Representative example of alteration zoning in volcanogenic massive sulphide deposits in the Bathurst district………………………………………...….. 73

Figure 6.2. Fe/Mn values of the Nepisiguit Falls Formation. From select drill cores at the Heath Steele E zone …………………………………………………………... 75

Figure 6.3. Alteration box plot modified for use with the suite of elements acquired with the pXRF…………………………………………………………………...…… 77

Figure 6.4. Alteration box plot modified for use with the suite pXRF acquired data collected during this study…………………………………………………..……….. 79

Figure 7.1. Geological map of the Heath Steele E zone deposit…………..………… 83

ix

List of Tables

Table 1.1. Drill cores analyzed in this study from the Heath Steele E zone deposit… 6

Table 1.2. Element scoring criteria………..………………………………………... 11

Table 1.3. Element scoring based on least squares regression (Mining Plus and Soil are analysis modes on the Inovex X 5000 equipment)……………………………… 11

Table 5.1. Average immobile elements and immobile-element ratios of the hanging wall and footwall rocks intersected by drill cores in the Heath Steele E zone …….. 56

Table 5.2. Shows the average immobile element ratios of the four lithogeochemical (chemostratigraphic) volcanic units in the E zone deposit…………………………... 59

x

Chapter 1

Introduction

1.1 General Statement

The Heath Steele E zone Zn-Pb-Cu-Ag volcanogenic massive sulphide (VMS) deposit is located approximately 37 km southwest of Bathurst and lies in northern part of the Miramichi zone in the Bathurst Mining Camp (BMC; Figure 1.1).

Figure 1.1. A) Lithotectonic divisions of the northeastern Appalachian orogen (modified after Hibbard et al. 2006). B) Simplified geology map of New Brunswick. The BMC area outlined in Figure 1.1B is presented in Figure 2.1 (from the New Brunswick Department of Natural Resources and Energy Development website).

The E zone deposit is an economically important part of the complex Middle

Ordovician volcano-sedimentary sequence dominated by felsic volcanic rocks (quartz- feldspar crystal tuff) and intercalated sedimentary rocks of the Nepisiguit Falls

1

Formation (Tetagouche Group), which were deposited in the Tetagouche-Exploits back- arc basin in the Middle Ordovician.

1.2 Objectives

A clear understanding of the host stratigraphy is difficult by the similarity among units, as well as intense penetrative poly-phase deformation, intense alteration, and faulting. These features make classification of rock types and the magmatic affinity challenging and make a detailed stratigraphic division based only on primary textures all but impossible. To better understand the E zone and discern the complications listed above, the current study was initiated by examining recently acquired drill core from the

Heath Steele E zone deposit. The objective of which was the application of the lithogeochemical and chemostratigraphic techniques to define rock types, magmatic affinity, and tectonic environment. This was used to reconcile regional and local geological evolution including the effects of hydrothermal alteration. This was achieved by the following:

1) Detailed core logging to identify units.

2) Detailed whole-rock geochemical analysis using portable X-Ray Fluorescence

(pXRF) technology to characterize the chemistry of structural hanging wall and footwall sequences.

3) Use of lithogeochemical and chemostratigraphy techniques to establish a local chemostratigraphy of the host sequence to the E-zone deposit.

4) Characterize (qualitatively) the deposit-related hydrothermal alteration in footwall and hanging wall rocks to develop exploration vectors.

2

The results will be used to improve the correlation of units and resolve the complex lithostratigraphic relationships in the E zone deposit that are a key component of mineral exploration. In the case of hydrothermal alteration analysis, this study aims to characterize qualitatively the alteration geochemistry by developing exploration vectors.

1.3 Justification

Lithogeochemical and chemostratigraphic techniques are essential tools for mineral exploration in poly-deformed and/or hydrothermally altered sequences.

Application of these techniques allow us to better understand the host sequences, geological processes, alteration patterns, and in some cases the original geometries of these deposits. In addition, better correlations among lithological units post-alteration, deformation, and metamorphism, can enhance stratigraphic and structural interpretations, thereby improving future exploration success rates. Unfortunately, there are few specific lithogeochemical studies that try to understand the petrogenesis, chemostratigraphy, and depositional environment of volcanic and sedimentary rocks hosting massive sulphide deposits in the study area.

These techniques employed herein are based on: 1) immobile-element variations which reflect tectono-magmatic affinity of volcanic rocks and provenance of the terrigenous material and 2) mobile hydrothermal-hydrogenous-element differences indicative of depositional paleo-environment of the rocks. In this context, immobile elements, i.e., the high field strength elements (HFSE) refer to elements that are either added to or removed from the rock during alteration or metamorphism. Immobile elements may be involved in phase changes and perhaps be mobile at the millimetre scale, but their mass in the altered rock remains unchanged. Although the proportions

3

(concentrations) of individual immobile elements may change, due to net mass changes in the size of the system, their inter-element ratios remain constant (Floyd and

Winchester 1978, MacLean and Barrett 1993, Jenner 1996, Lentz 1996a). In contrast, the mobile elements (Si, Fe, Mg, Ca, Na and K) refer to the major elements that are relatively mobile during alteration and metamorphism (MacLean and Barrett 1993,

Rollinson 1993, Jenner 1996, Ross and Bedard 2009). Many studies of VMS deposits have shown that Al and the HFSE (Zr, Hf, Nb, Ta, Y, Sc, Ti, V), and heavy rare earth elements (REE; Yb, Lu) remain essentially immobile during alteration (MacLean and

Barrett 1993, Lentz 1996a, Piercey 2010, Piercey and Devine 2014). Their immobility has been documented even in the most intense hydrothermally altered zones directly beneath deposits, e.g., in the Bathurst Camp (Brunswick no. 6 and Heath Steele B zone deposits; Lentz 1996a).

1.4 Analytical methods

Several portable X-ray fluorescence (pXRF) studies have been conducted in a variety of geological environments and deposit types. These studies have shown that portable XRF analyser technology provides credible and relevant geochemical information (Lemiere 2018, Lemiere and Uvarova 2019). Comparing conventional laboratory and pXRF data, it has been shown that there is a good correlation between both analysis (i.e. r2 >0.80) (Piercey and Devine 2014). Therefore, pXRF is an effective tool for exploration lithogeochemistry and chemostratigraphy. In addition, the pXRF technology can be used to collect large amounts of multi-element data which includes:

HFSE and REE useful for magmatic affinity or tectonic discrimination diagrams; precious metals at low concentrations; and even major elements essential for the

4

calculation of many hydrothermal alteration indices, rapidly at a relatively low cost

(Fisher et al. 2014, Ross et al. 2014). It can assist in differentiating among volcanic units, especially in the pervasively hydrothermally altered lithologies related to volcanogenic massive sulphide deposits, where the primary protolith is unrecognisable on the basis of macroscopic textures (Sack and Lewis 2013, Ross et al. 2014).

The fundamental principle utilized in pXRF-EDS analysis is the photoelectric effect, where incident of x-ray energy flux strikes the inner shell electrons of the atoms of elements within the sample with sufficient energy to cause an electron in the K- and

L-shells to be displaced. An electron in a higher orbital will fill the vacancy in the inner orbital, thereby producing a photon with energy equal to the difference in energy between the two orbitals. Each element has diagnostic photon emission energies from its various orbitals, and the total emission spectrum for specific incident X-ray energy can be deconstructed to calculate the absolute concentration of most elements (Glanzman and Closs 2007, Potts and West 2008).

1.4.1 Whole-rock lithogeochemistry and pXRF analytical methods

This project involved logging and sampling of 8 selected diamond drill cores

(Appendix A) from the Heath Steele E zone deposit stored at the New Brunswick

Department of Energy and Resource Development’s Madran drill core facility and at the

Caribou Mine, west of Bathurst, New Brunswick (Tables 1.1). The work was conducted in one field season from June to September of 2019. The portable X-ray Fluorescence

Spectrometry (pXRF) analysis was conducted on 856 samples from a variety of rock types, e.g. quartz-feldspar crystal tuff, fine-grained sedimentary rocks, and exhalative rocks or iron formation, intersected by these drill cores.

5

Table 1.1. Drill cores from the Heath Steele E zone deposit analyzed in this study.

Hole_ID Easting Northing Dip Azimuth Length (m) Samples Stored S-362 721883 5242456 -55˚ 342 ˚ 195.68 172 Madran S-806 721868 5242540 -47 ˚ 343 ˚ 136.24 78 Madran S-832 721838 5242533 -48 ˚ 339 ˚ 139.29 102 Madran S-908 721922 5242425 -52 ˚ 339 ˚ 288.33 82 Madran S-912 721922 5242473 -50 ˚ 339 ˚ 236.49 100 Madran S-913 721911 5242456 -48 ˚ 339 ˚ 245.39 86 Madran S-916 721871 5242481 -48 ˚ 340 ˚ 209.09 144 Madran HS17-005 721882 5242394 -52 ˚ 343 ˚ 321.00 92 Caribou

Next, analysis points were selected on the basis on textural representativeness within a given core interval of the various hanging wall and footwall units.

Representative samples were collected every 1 meter where possible. It is important to note that in some cases the sampling interval was greater than 1 meter to avoid localized alteration features (e.g., quartz, carbonate, and sulphide veins) whereas, in other instances, sampling intervals were as little as 50 centimeters. Finally, each core sample was cut to produce a flat and smooth surface of between 6 and 10 cm in length to create a suitable uniform sampling site.

Data were obtained using a field-portable benchtop Olympus/ Innov-X X5000 energy dispersive pXRF equipped with a Ta X-ray tube with a maximum tube voltage of

50 kV, maximum tube power of 10 W, detector area of 25 mm2, and <165 eV spectral resolution. This field pXRF can provide very precise data for selected major, minor, and trace elements, generally with relative standard deviation (RSD) values of <7.5 % and many <5 %, except at very low concentrations (i.e., approaching the limit of detection).

Accuracy is variable and ranges from excellent to reasonable for many major and minor

6

elements (±15–20 % relative difference (RD), for Al2O3, SiO2, K2O, CaO, Fe2O3, TiO2, and MnO±S), base metals (±20 % for Cu, Zn), the low field strength (LFSE) and high field strength (HFSE) elements (±15 % RD for Rb, Ba, Zr; ±20 % RD for Nb) and poor accuracy was obtained for MgO and P2O5 (Piercey and Devine 2014).

In this study the pXRF was operated in two different analytical modes (Soil and

Mining Plus modes). Soil mode is used to collect data for minor and trace elements with concentrations less than 1 wt.% where concentrations are measured in ppm. Mining Plus for analyzing at higher elemental concentrations, with concentrations >0.5–1 wt.%, in this case, the measured concentrations are output as weight percent (wt.%). Soil mode employs three beams: beam 1 at 50 kV, beam 2 at 35 kV, and beam 3 at 15 kV, and

Mining Plus employs two beams: beam 1 at 35 kV and beam 2 at 10 kV. For every analysis, each beam was integrated over 80 seconds, for a total analysis time of 240 and

160 seconds, respectively.

After every twenty analyses, the calibration check was performed by analyzing silica blank and selected certified reference materials (CRM), i.e., SY-4, SY-2, SBC1, and RHY1. These standards were chosen because they have similar concentrations and matrixes to the samples being analysed in the study. Given that the sampling window of the pXRF analyzer used in this study is 8 mm in diameter, each sample was analyzed three to five times depending on the texture of each sample (fewer analyses are required in finer-grained more homogenous rocks relative to coarse-grained rocks) in order to get an average representative value. To obtain a calibration-correction, the raw data for a given element from the calibration standards ([w]푟푒푓푒푟푒푛푐푒 푚푎푡푒푟푖푎푙,푟푎푤) were divided by literature values ([w]푙푖푡푒푟푎푡푢푟푒) to provide a correction factor Corr푖 for a given element i:

7

[w]푙푖푡푒푟푎푡푢푟푒 Corr푖 = ) [w]푟푒푓푒푟푒푛푐푒 푚푎푡푒푟푖푎푙,푟푎푤

Finally, this correction factor was then applied to the raw concentration of element i in the unknown sample or blank j (i.e.,[i]푟푎푤,푗) to provide a calibration corrected concentration for element [i] in sample j (Piercey and Devine 2014; Appendix

C contains the corrected data of the analyzed samples):

i푐표푟푟푒푐푡푒푑,푗 = corr푖 ∗ [i]푟푎푤,푗

1.4.2 Statistical Analysis

In this type of study quality assurance and quality control (QA/QC) are critical in assessing integrity of the geochemical data from field acquisition through to the generation of the final results. In the present study, accuracy is quantified by percent relative difference (%RD) and precision by percent relative standard deviation (%RSD)

(Jenner 1996, Piercey 2014). The %RD (Equation 1) is determined by analysing a CRM and is the difference between the accepted elemental concentration and the elemental concentration measured by the instrument. The %RSD (Equation 2) is determined by analysing duplicate samples and determining the deviation between measurements of the same material (Piercey and Devine 2014). In general, accuracy and precision can be classified using %RD and %RSD as excellent (0 to ≤3%), very good (>3 to ≤7%), good

(>7 to ≤10%), or inaccurate/imprecise (>10%) precision (Jenner 1996).

휇푖−푆푇퐷푖 %푅퐷푖 = 100( )……….Eq. 1 푆푇퐷푖

푆푖 %푅푆퐷푖 = 100 ( )…………..Eq. 2 휇푖

8

Where %푅퐷푖 = percent relative difference of element 푖, 휇푖 = mean value of element 푖 during analysis, 푆푇퐷푖 = accepted concentration of element 푖 in the CRM,

%푅푆퐷푖 = percent relative standard deviation for element 푖, and 푆푖 = standard deviation from mean of element 푖 over duplicate runs (Piercey and Devine 2014).

To improve the accuracy of raw pXRF output compositions, corrections were applied to the data based on the multiple analyses of CRMs, as well as a silica blank.

The data were normalized to the background by subtracting the average composition of the silica blank from every analysis (Garcelon et al. 2016). The data were then corrected to the CRMs by establishing correction curves (Figure 1.2) over a range of concentrations for each element. X-Y plots of the accepted values (documented concentrations plotted on the x-axis) versus the values measured by the pXRF (on the y- axis) were plotted for each element for all CRM analyses. Using least squares regression analysis, a line of best fit with a y-intercept of zero was established to determine the slope (m) of the regression line. The inverse of the slope was used as the correction factor for each element (Garcelon et al. 2016). This method allows for simple quantification of the quality of the correction factor, based on the coefficient of determination (푅2) value of the linear regression (Hall et al. 2011, 2013). The coefficient of determination, R2, provides an assessment of the robustness of the regression. The closer the corrected composition data is to the true composition data the closer the R2 value is to 1 (Fisher et al. 2014).

9

Figure 1.2. Portable XRF correction curves for selected elements used in this project. CRM=certificate reference material values, pXRF=analytical results from portable XRF.

10

Following the results of least squares regression analysis (Figure 1.2; see

Appendix B for more correction curves), elements were scored (see Tables 1.2 and 1.3) based on the slope and goodness-of-fit (accuracy and precision, respectively) using the criteria employed by Hall et al. (2011).

Table 1.2. Element scoring criteria (modified after Hall et al. 2011).

Accuracy Precision Good <0.05 >0.9 Fair 0.05-0.10 0.8-0.9 Poor 0.10-0.20 0.5-0.8 Bad >0.20 <0.5

Table 1.3. Element scoring based on least squares regression for Mining Plus and Soil analytical modes on the Innovex X5000 pXRF used in this study.

Accuracy Precision Element Mining Plus Soil Mining Plus Soil Al 0.13 - 0.78 - K 0.16 0.19 0.93 0.76 Ca 0.08 0.45 0.95 0.90 Ti - 0.12 - 0.83 Mn 0.33 0.10 0.89 0.95 Fe 0.18 0.13 0.90 0.94 Y - 0.08 - 0.85 Zr 0.17 0.03 - 0.88 Nb - 0.04 - 0.92 Th - 0.03 - 0.84

1.5 Exploration History

In 1954, Little River (renamed Heath Steele) was discovered as a result of an airborne electromagnetic (AEM) survey that was conducted by the American Metal

Company (AMCO later AMAX) and International Nickel Company (INCO), the very

11

first AEM discovery in the world (Luff 1995, McCutcheon et al. 2003, McCutcheon and

Walker 2019). Numerous airborne electromagnetic anomalies were identified at that time and the second hole of a follow–up drill program intersected the A zone in that year. Further drilling of other airborne electromagnetic anomalies led to the discovery of the B, C, D, and E zones (McCutcheon et al. 2003 and references therein).

In 1964 the Heath Steele N-5 zone was discovered in the northern part of the mining lease (Hamilton et al. 1993). In 1965, the B-5 zone was found at a depth of over

300 m below surface and approximately 750 m southeast of the B zone. In 1966, the

West Grid zone was discovered approximately 1.5 km to the west of the Heath Steele

ACD zone. Two of the six discoveries resulted from drill testing coincident soil geochemical & weak electromagnetic anomalies, while one was found by prospecting.

In 1981, discovered the Heath Steele C-North zone, approximately 550 m north of the ACD zone, by drill testing coincident ground magnetic and electromagnetic anomalies. In 1987, Noranda discovered the Heath Steele H-2 zone by drill testing a combined soil geochemical-electromagnetic anomaly in the southwestern part of the mine lease. Continued work by Noranda led to discovery of the Heath Steele

HC-4 deposit in 1991 (McCutcheon et al. 2003).

Heath Steele E zone was discovered in 1954 thanks to a geological mapping program and an AEM of the Central Highlands (McCutcheon et al. 1993). The total geological ore reserves in 1979 for the B and ACD zones was estimated at 24 Mt grading 5.2% Zn, 1.8% Pb, 0.93% Cu, and 65.6 g/t Ag, and the reserve for E zone was

12

estimated to be 1.2 Mt grading 4.33% Zn, 1.98% Pb, 1.56% Cu, and 73.1 g/t Ag (Gates

1972).

Mining operations at Heath Steele were carried out by a variety of companies intermittently between 1957 and 1999. Production began in 1957, but low metal prices and poor recoveries forced a shut down from 1958 to 1962; the original production rate of 800 tons per day was increased to 4000 tons per day by 1975 (McBride 1976). At the end of 1970, reserves for the B and ACD zones were estimated at 34.6 Mt grading

4.55% Zn, 1.64% Pb, 1.15% Cu and 49.6 g/t Ag.

The E zone, (the subject of this research project) has not been mined; however, a historic resource estimated at 1.3 Mt grading 4.33% Zn, 1.95% Pb, 1.56% Cu, and 66.3 g/t Ag was reported (Gates 1972). In March 2017, Trevali Mining Corporation commenced an initial 7-hole, 1,200 m drill program to validate the historic resource estimate of the E zone. The first drill hole, HS17-001 intersected 56.2 metres of sulphide mineralization grading at 3.84% Zn, 1.51% Pb, 2.15% Cu, 64.3 g/t Ag, and 1.01 g/t Au, within which are several higher grade intervals. At the time of writing this thesis, exploration is ongoing. As part of this exploration, funding was provided to UNB for chemostratigraphic analysis of the host sequence, the results of which are reported herein.

1.6 Previous (scientific) work

Since 1950 the Geological Survey of has conducted important geological, geochemical, and geophysical studies in the BMC. These include McAllister

(1960), Whitehead and Goodfellow (1978), van Staal and Fyffe (1991), Lentz (1994,

13

1995, 1996a, 1996b, 1996c, 1997, 1999), Lentz and Goodfellow (1993), Whalen et al.

(1998), Currie et al. (2003), Goodfellow and McCutcheon (2003), Rogers et al. (2003a,

2003b), and van Staal et al. (2003) who have described the geology, tectonic setting, deformation, and metamorphic history of the BMC in general, as well as detailed studies on the stratigraphy, hydrothermal ore-forming processes, and hydrothermal alteration of specific deposits.

From 1994 to 1999 the Exploration Science and Technology Initiative (EXTECH

II) multi-disciplinary research project, conducted by the New Brunswick Department of

Natural Resources and Geological Survey of Canada, focused on the BMC with the objective of developing new exploration ideas and methods. This program was a co- operative program involving the federal and provincial governments, universities, and the mineral industry. Research was focused in six main areas: 1) geologic and tectonic setting; 2) genesis of massive sulphide formation; 3) surficial geology and geochemistry;

4) hydrology and hydrochemistry; 5) geophysical surveys including ground, borehole, and airborne methods; and 6) compilation of a digital database using a Geographical

Information System (GIS) (Langton and McCutcheon 1996, Goodfellow et al. 2003).

In the Heath Steele area specifically numerous studies have focused on a variety of research topics, including: geology, tectonic setting, deformation, and metamorphic history (McBride 1976, Harley 1979, Owsiacki and McAllister 1979, de Roo et al. 1990,

1991, van Staal and Fyffe 1991, Hamilton 1992, Lentz 1998, de Roo and van Staal 2003,

Currie et al. 2003), detailed studies on the petrology, geochemistry, and stratigraphy

(Whitehead and Goodfellow 1978, Langton 1996, Rogers 1995, Wilson et al. 1998,

Lentz 1999, Downey 2005, Downey and Lentz 2006), chemostratigraphy (Whitehead

14

1973, Whitehead and Govett 1974, Whitehead and Goodfellow 1978, Lentz 1996a,

1997, Lentz et al. 1996, 1997, Lentz and Wilson 1997, Downey et al 2006), hydrothermal ore-forming processes, and hydrothermal alteration (Lusk 1969, Owsiacki and McAllister 1979, McDonald 1983, McCutcheon 1992, Lentz and Goodfellow 1993,

Lentz and Goodfellow 1994a, Peter and Goodfellow 1996, Peter 2003, Peter and

Goodfellow 2003, Peter et al. 2003a, 2003b) in specific zones of the Heath Steele area.

15

Chapter 2

Geology of the Bathurst Mining Camp

2.1 Tectonic setting

The Bathurst Mining Camp (BMC) is made up of several different tectonic blocks and slivers (Fournier, California Lake, Tetagouche, and Sheephouse Brook blocks and the blueschist and Bamford Brook slivers) (Figure 2.1).

Figure 2.1. Simplified tectonic map of Bathurst Mining Camp, New Brunswick. C=Carboniferous; FB=Fournier block; HSN=Heath Steele nappe; LLN=Lucky Lake nappe; LRLA=Little River Lake antiform; MLMB=Moose Lake–Mountain Brook ; NA=Nepisiguit antiform; NMS=Nine Mile synform; NN= Nepisiguit nappe; PRN=Portage River nappe; RA=Restigouche antiform; RBMF=Rocky Brook Millstream fault; S=Silurian rocks; SB=Sheephouse Brook block; SDG=Silurian-

16

Devonian granite; SLN=Strachens Lake nappe; SPN=Spruce Lake and Mount Brittain nappes; TA=Tetagouche antiform; TN=Tomogonops nappe. The area of this figure is outlined in Figure 1.1B (modified from van Staal et al. 2008).

These blocks and slivers (Figure 2.1) belong to the Dunnage (Bathurst and

Fournier supergroups) and Gander (Miramichi Group) zones of the Canadian

Appalachians. They represent widely separated, ensialic to ensimatic portions of the

Tetagouche-Exploits back-arc basin formed during the Middle Ordovician by rifting of the submarine Popelogan Arc on the continental margin of Ganderia (Figure 2.2A) on the southern margin of the Iapetus Ocean (van Staal et al. 2003, McCutcheon and

Walker 2019 and references therein).

Sea floor volcanism in the back-arc basin was spawned by crustal melting induced by mantle upwelling related to slab roll back (van Staal et al. 2003). Early ensialic volcanism was marked by submarine eruptions of silicic flows and submarine epiclastic, volcaniclastic, and pyroclastic deposits of dacitic to rhyolitic composition.

Texturally these rocks range from crystal-rich tuffs, hyaloclastite, to later effusive aphyric lavas and subvolcanic rhyolitic cryptodomes (Figure 2.2B and 2.2C; Sullivan and van Staal 1990, Sullivan and van Staal 1996, Goodfellow et al. 2003). Early felsic magmatism and related VMS mineralization was followed by a second pulse of felsic magmatism and associated minor VMS mineralization. The second pulse of felsic magmatism gives way up-section to alkaline and tholeiitic basaltic lava flows and hyaloclastites and related sedimentary rocks. This change from felsic to mafic magmatism is coincident with continued back-arc rifting and basin widening (van Staal et al. 2003, Goodfellow et al. 2003).

17

18

Figure 2.2. Schematic block diagrams illustrating stages of the tectonic and petrogenetic evolution of the Tetagouche-Exploits back-arc basin in four time intervals a) 478–475 Ma, b) 474–469 Ma, c) 466 Ma, and d) 457 Ma (modified from Wills 2014).

Likewise, the progression from alkaline to tholeiitic mafic composition may reflect the transformation from a back-arc basin into oceanic marginal sea (Figure 2.2D; van Staal et al. 2003), which is also marked by the deposition of maroon-coloured pelagic mudstone, siltstone, and chert intercalated with flows of alkaline

(Goodfellow et al. 2003). The textures and composition of sedimentary rocks indicate that the back-arc basins formed in varying water depths; some basins had shallow water submarine to subaerial environments, whereas others were in much deeper marine environments (Goodfellow and Peter 1996). The back-arc basin ultimately closed by northwest-directed subduction during the Late Ordovician to Early Silurian (van Staal

1994, van Staal and de Roo 1995, van Staal et al. 2003).

The various parts of the back-arc basin were tectonically juxtaposed by thrusting

and internally imbricated into thrust panels during their incorporation into an accretionary prism, i.e., the Brunswick Subduction Complex (van Staal 1994). Within the Brunswick Subduction Complex, the rocks were subjected to poly-phase deformation and middle- to upper-greenschist grade metamorphic conditions. Detailed geothermobarometry using a variety of methods consistently gave metamorphic temperatures ranging between 340 and 370 °C and pressures between 5.5 and 6.0 kbar in rocks of the Tetagouche block and the Spruce Lake nappe of the California Lake block

(Figure 2.1; Currie and van Staal 1999, Currie et al. 2003, van Staal et al. 2003)

2.2 Stratigraphy of the Bathurst Mining Camp

The following description of regional geology is meant to be a summary; for complete detailed descriptions of the stratigraphy and structure the reader is referred to van Staal and Fyffe 1991, Rogers 1994, McCutcheon and Walker 2009, Walker and

19

McCutcheon 2011, Wilson 2003, Wilson et al. 2015, or in the Bedrock Lexicon of the

Geological Surveys Branch of New Brunswick Department of Energy and Resource

Development available online: http://dnrmrn.gnb.ca/Lexicon/

Lexicon/Lexicon_Search.aspx?lang=e.

First, it is important to clarify that the rock names used throughout this text are protolith names; however, as mentioned in the previous section, the BMC was intensely deformed and metamorphosed during multiple collisional events related to subduction of the basin, so all the rocks are metamorphosed to greenschist facies. It was not possible to subdivide these rocks into mappable units on the basis of their primary volcanic features alone. The volcanic rocks were stratigraphically subdivided on the basis of 2,321 high- quality whole-rock geochemical analyses in combination with detailed

geochronological, petrologic, and structural studies (e.g., van Staal et al. 1991, 2003;

Rogers and van Staal 2003, Rogers et al. 2003a, 2003b).

The rocks of the Bathurst Mining Camp are divided into three major packages, i.e., 1) Miramichi Group, 2) Bathurst Supergroup, and 3) Fournier Supergroup. The

Miramichi Group (Figure 2.3) is a Cambro-Ordovician clastic sedimentary succession, formed as a flysch on the stable Gondwana margin and is overlain conformably to locally unconformably by the Bathurst Supergroup. The Miramichi Group is divided into three formations that, in ascending order, are: Chain of Rocks (fine- to medium- grained quartzose sandstone intercalated with phyllitic siltstone and shale), Knights

Brook (sandstone (quartzite), quartz wacke, minor feldspathic wacke, and interbedded shale, fine- to medium-grained, micaceous sandstone and siltstone), and Patrick Brook

(generally thin-bedded shale siltstone, feldspathic wacke, and local fine-grained

20

sandstone characterized by abundant volcanic quartz phenoclasts) formations (van Staal and Fyffe 1991, van Staal et al. 2003, Wilson et al. 2015).

Figure 2.3. Simplified geological map of the Bathurst Mining Camp, the red box shows the study area, the light blue represents undivided Silurian-Carboniferous cover sequences, whereas the grey dark and pink units represent Ordovician and Siluro- Devonian intrusive rocks, respectively (from the New Brunswick Department of Energy and Resource Development website).

The Bathurst Supergroup consists of Middle Ordovician bimodal volcanic and related sedimentary rocks deposited during the early rifting and subsequent widening phase of the Tetagouche-Exploits back arc basin. The Bathurst Supergroup is divided into three more-or-less coeval groups, i.e., Sheephouse Brook, Tetagouche, and

California Lake groups (Figure 2.3). All three groups are conformable on the Miramichi

21

Group and in the case of the Tetagouche Group locally disconformable on the Miramichi

Group and represent distinct eruptive centers in the back-arc, and all three host volcanogenic massive sulphide (VMS) mineralization (see section 2.3.2).

The Sheephouse Brook Group is divided into three formations that, in ascending stratigraphic order are: Clearwater Stream, Sevogle River, and Slacks Lake formations. The Clearwater Stream Formation has returned U-Pb zircon age of 469.0 ±

0.9 Ma (Wilson and Kamo, 2008) and consists of dacitic to rhyolitic tuffs. This formation is host to the Chester VMS deposit. The Clearwater Stream Formation is conformably overlain by alkali feldspar-phyric rhyolite and minor shale of the Sevogle

River Formation, alkalic to tholeiitic basalt, graphitic shale, and minor ferromanganiferous shale and chert of the Slacks Lake Formation (Wilson et al. 1999).

The Tetagouche Group is divided into five formations that, in ascending stratigraphic order, are: Nepisiguit Falls, Flat Landing Brook, Little River, Tomogonops, and Melanson Brook formations.

The Nepisiguit Falls Formation consists of rhyodacitic to rhyolitic crystal tuff, quartz-feldspar porphyritic felsic volcanic rock, quartz-feldspathic volcaniclastic rocks, fine- to medium-grained quartz-feldspathic wacke, and dark greenish grey shale and siltstone, as well as massive sulphides and related chemical exhalative sedimentary rocks (mainly oxide and silicate facies iron formation). The Nepisiguit Falls Formation has yielded several U-Pb zircon dates ranging between 474 and 469 Ma; van Staal and

Fyffe 1991, Rogers et al. 2003a).

22

The Flat Landing Brook Formation is conformable on the Nepisiguit Falls

Formation and consists of aphyric to sparsely fine feldspar- (and rare fine quartz-) phyric rhyolite flows, flow-breccias, and hyaloclastites. The Flat Landing Brook Formation also includes locally abundant tholeiitic to transitional mafic volcanic rocks including massive to pillowed basalt flows, felsic crystal ± lithic ± vitric tuff and minor porphyritic felsic flows, and related clastic sedimentary rocks, ferromanganiferous shale and chert, and ironstone are minor constituents. This unit has yielded a U-Pb zircon age of 466 ± 2

Ma (Rogers et al. 1997).

The Little River Formation is dominated by dark grey shale and siltstone, black shale and chert, and red and green ferromanganiferous shale and chert; significant volumes of transitional to alkalic pillow (van Staal et al. 1991, Wilson et al.

1998, van Staal et al. 2002).

The Tomogonops Formation is a syn-tectonic sedimentary unit derived by erosion of underlying units as they were incorporated into the Brunswick Subduction

Complex. The Tomogonops Formation comprises a coarsening-upward sequence of light- to medium- grey, thin- to medium-bedded, calcareous siltstone, shale, lithic wacke and quartz wacke, and thicker bedded, non-calcareous coarse-grained sandstone and conglomerate containing clasts of mafic and felsic volcanic and sedimentary rock

(Langton 1996, Sullivan and van Staal 1996, Rogers et al. 2003a, van Staal et al. 2003,

Wilson et al. 2015).

The Melanson Brook Formation consists mainly of medium to dark grey or greenish grey, massive to very prominently laminated, typically moderately to strongly

23

calcareous, but locally non-calcareous slaty siltstone that locally grades to calcisiltite or calcilutite (Wilson 2003, Wilson et al. 2015).

The California Lake Group comprises six formations Spruce Lake, Canoe

Landing Lake, Mount Brittain, Boucher Brook, and Middle River formations. The first three formations are more-or-less coeval and are all conformably overlain by Boucher

Brook Formation. The Spruce Lake and Mount Brittain formations consist in rhyodacitic to rhyolitic feldspar-crystal flows and tuff and crystal lithic tuff, with subordinate alkalic to tholeiitic basalt, and fine-grained sedimentary rocks. The Canoe Landing Lake

Formation is dominated by basalt flows tuffs and related fine-grained clastic sedimentary rocks including massive to prominently laminated, non-calcareous slate siltstone, and thin- to medium-bedded quartzose sandstone. Most of the VMS deposits in

this group occur in the Spruce Lake Formation, whereas two deposits occur in the Mount

Brittain Formation and one in the Canoe Landing Lake Formation (Wilson et al. 2015).

The Middle River Formation comprises syntectonic sedimentary rocks that overlie or are in tectonic contact with the Boucher Brook Formation in the Canoe Landing Lake thrust panel (Wilson et al. 2015).

The Fournier Supergroup represents a dismembered ophiolite sequence and comprises three major thrust nappes that, from structural base to top, are: the Sormany

Group, Pointe Verte Group, and Devereaux Complex (van Staal and Wilson 2014). The

Fournier Supergroup structurally overlies the California Lake Group (part of the

Bathurst Supergroup) along a thrust fault that is spatially associated with a belt of high- pressure metamorphic rocks (van Staal et al. 1992). In the BMC the Fournier

24

Supergroup is restricted to the northern limb of the Tetagouche Antiform and the core of the Nine-Mile Synform (Figures 2.1 and 2.3).

The various groups were juxtaposed by thrusting and internally imbricated into thrust nappes during their successive incorporation into the Brunswick subduction complex (van Staal 1994, van Staal et al. 2003). The Brunswick subduction complex formed as a result of Late Ordovician-Silurian northwest-directed subduction of the wide

(~1000 km) Tetagouche-Exploits back-arc basin, which extended along most of the length of the northern Appalachians (van Staal et al. 1998, 2003, Valverde-Vaquero et al. 2006, van Staal et al. 2008), following the immediately preceding, Taconic accretion of the Popelogan-Victoria arc to Laurentia (van Staal et al. 2008).

2.3 Volcanogenic Massive Sulphide Deposits in the BMC

2.3.1 Review of VMS systems

Volcanogenic massive sulphide (VMS) deposits are epigenetic to syngenetic, stratiform and stratabound/lenticular bodies of sulphide minerals, precipitated from metallifeous hydrothermal fluids at or immediately below the sea-floor. The deposits typically occur as polymetallic, massive sulphide lenses dominated by Fe-sulphides, but commonly containing significant amounts of Cu, Zn, Pb, Ag, and Au. These deposits form through the precipitation of metal sulphides from metal-bearing hydrothermal fluids as they mix with cold seawater (Franklin 1997, Franklin et al. 1981, 2005, Galley

1993, Galley et al. 2007, Gibson et al. 1999, Gibson and Galley 2007, Lydon 1988a,

1998b).

25

The VMS deposits can be classified in several ways, but most recently classifications on the basis of lithostratigraphy are popular (Piercey 2011). This classification scheme divides VMS deposits into five types on the basis of their respective host rocks, specifically: 1) mafic, 2) mafic-siliciclastic, 3) bimodal-mafic, 4) bimodal-felsic, and 5) felsic-siliciclastic (Barrie and Hannington 1999, Galley et al.

2007, Piercey 2011).

The first three types are hosted by sequences dominated by mafic footwall rocks with varying amounts of siliciclastic and chemical sedimentary rocks, minor felsic rocks in the case of bimodal mafic environments, and mafic to ultramafic intrusive rocks; these sequences are commonly juvenile and have very little continental crustal influence.

In contrast, felsic volcanic and sedimentary rocks dominate the last two types which are

commonly associated with evolved continental crust. According to this classification scheme, the VMS deposits of the BMC are included in felsic-siliciclastic type (van Staal et al. 2003, Galley et al. 2007, Piercey 2011).

The accepted genetic model is based mostly on empirical observations from ancient terrestrial and modern seafloor systems, suggesting that VMS deposits are the product of district-scale hydrothermal convection resulting from anomalous thermal input into the shallow crust (Galley 1993, Franklin et al. 2005). The development and maturation of a subseafloor hydrothermal system can be divided into three stages

(Franklin et al. 2005):

1) The relatively deep emplacement of beneath a rift or caldera and the establishment of a shallow circulating, low-temperature seawater convection system.

26

This circulating hydrothermal fluid results in shallow subseafloor alteration and mass changes and formation of hydrothermal exhalative sediments (Figure 2.4A).

Figure 2.4. Schematic geological model for the formation of VMS deposits with submarine geological rift setting, physical geologic, geochemical, and hydrothermal characteristics (modified from Franklin et al. 2005, Galley et al. 2007, Gibson et al. 2007).

2) Shallowly emplaced subvolcanic and resultant generation of a deep- seated, sub-seafloor seawater convection system in which net gains and losses of elements are dictated by subhorizontal isotherms (Figure 2.4B).

3) Development of a mature, large-scale hydrothermal system in which subhorizontal isotherms control the formation of semi-conformable hydrothermal

27

alteration assemblages. The high-temperature reaction zone adjacent the cooling intrusion is periodically breached due to seismic activity or dyke emplacement, allowing focused up-flow of metal-rich fluids to the seafloor and formation of VMS deposits

(Figure 2.4C; Galley et al. 2007).

Six principal factors are common to VMS-forming hydrothermal systems (see

Figure 2.4C): 1) a heat source (magma) at depth, 2) a high-temperature reaction zone formed through the interaction of footwall strata with changed seawater, 3) deep penetrating, syn-volcanic faults and seismic activity allowing egress of seawater and focused discharge of venting fluids, 4) hanging wall and footwall alteration zones that are controlled by variations in wall rock (e.g., composition, permeability, and porosity), fluid composition, temperature, and water/rock ratios with respect to the core of the

hydrothermal vent, 5) stratiform to locally stratabound massive sulphide lenses formed at or near the seafloor and whose metal zoning is a product of zone refining process

(Ohmoto 1996), and 6) hydrothermal plume fall out which contributes to sedimentary rocks distal to the deposit (Galley et al. 2007).

As a result of fluid convection and variation in the water/rock ratio in the subsurface, VMS deposits have extensive distal semi-conformable (recharge) and proximal discordant (pipe-like/discharge) zones of hydrothermal alteration. The size distribution and mineralogy and of these alteration zones varies with host rock characteristics (chemistry, permeability and porosity etc.) and proximity to the hydrothermal vent site (footwall moderate to locally intense and hanging wall typically weak to absent) (Figure 2.5). The alteration zones are typically zoned laterally and vertically with more intense chloritic alteration developed proximal to the upwelling

28

zones in the footwall, whereas sericitic alteration is developed in more distal parts. In some systems, ferruginous, silicic, and carbonate alteration is pervasive in the footwall

(Franklin et al. 2005, Galley et al. 2007, Piercey 2007).

Figure 2.5. Distribution and characteristic mineralogical assemblages of alteration facies in a typical VMS system (from Wills 2014).

A typical VMS deposit consists of two major types of sulphide bodies: a stratiform or stratabound massive sulphide lens where the sulphide contents generally exceed 50% by volume and a stockwork (stringer, or siliceous) zone consisting of veinlets and disseminated mineralization in the footwall rock of the massive sulphide body (Figure 2.6; Ohmoto 1996). The stratiform massive sulphide body is commonly mound shaped with typical dimensions of 20 m in height and ~300 m in radius (i.e., a very small height/length ratio). Stockwork zones commonly have a downward- narrowing, funnel-shape with an average radius of ~100 m at the top and a downward length of ~100 m (Ohmoto 1996).

29

Figure 2.6. Idealized Kuroko-type VMS deposit (modified from Ohmoto 1996).

In general VMS deposits (more specifically the Kuroko deposits, see Ohmoto

1996) have a strong metal zonation, from top to bottom, barite ore (barite > sulphides, in abundance), black ore or Zn-Pb-rich ore (sphalerite ≈ barite > pyrite ≈ galena), yellow ore or Cu-rich ore (chalcopyrite ≈ pyrite > sphalerite), and pyrite ore (nearly monomineralic). The stockwork zones are characterized by abundant quartz and rare barite. Metal zoning is similar to that in the massive ores, but expressed laterally instead of vertically (i.e., pyrite-rich core, to yellow ore, and to peripheral black ore), may be present. Many of the massive sulphide bodies are overlain by thin (< 1 m) and discontinuous layers of ferruginous chert, which is called "chert-hematite ore",

"hematite-quartz ore", "tetsusekiei ore", or "exhalite". Massive bodies of gypsum/anhydrite are commonly developed beneath the massive sulphide body in areas peripheral (distal to) the stockwork (Figure 2.6; Ohmoto 1996).

30

2.3.2 Massive Sulphide Deposits at the BMC

The BMC has been the subject of intensive study since the discovery of major

VMS deposits in the early 1950s. The BMC hosts forty-five massive sulphide deposits that collectively account for a pre-mining massive sulphide resource in excess of 0.5 billion tonnes (McCutcheon et al. 2003, Goodfellow 2007, McCutcheon and Walker

2019). The Tetagouche Group hosts 31 of the deposits and the majority of the sulphide tonnages, whereas 13 deposits are hosted by the California Lake Group and one deposit is hosted by the Sheephouse Brook Group (McCutcheon et al. 2001, 2003). The most famous of the Tetagouche Group deposits is the world-class Brunswick No.12, which produced 136,643,367 tonnes of ore grading 3.44% Pb, 8.74% Zn, 0.37% Cu, and 91 g/t

Ag over its 49-year life; in addition the Brunswick No. 6, Austin Brook, and the Heath

Steele deposits are hosted by the Nepisiguit Falls Formation. Deposits hosted by the

Nepisiguit Falls Formation share several characteristics (McCutcheon et al. 2001):

 The massive sulphides are underlain by chloritic mudstone and/or very

fine -grained volcaniclastic rocks (McCutcheon 1992), which generally

have an aerial extent equal to or larger than the limits of the massive

sulphide bodies.

 Most (particularly in the eastern part of the BMC) are capped by and (or)

have a laterally equivalent oxide facies iron formation that is interbedded

with and passes into chloritic (silicate) iron formation along strike

(Goodfellow 2007).

 Various alteration facies can be recognized in the footwall volcanic rocks

(Lentz and Goodfellow 1994a), including proximal silicic-Fe-chloritic,

31

Fe-chloritic (±sericitic), Fe-Mg-chloritic sericitic, distal sericitic-Mg-

chloritic, and least altered (regional metamorphic).

 Large-scale mineralogical and/or chemical zoning may be present, both

vertically and laterally, in the deposits (Goodfellow and McCutcheon

2003).

Some deposits occur in the second cycle volcanic rocks e.g. the Flat Landing

Brook Formation. The characteristics common to these deposits are:

 Most are hosted by siliceous tuff and (or) fragmental rocks rather than

mudstone.

 Oxide iron formation is absent, except at the Louvicourt deposit where

red and green ferro-manganiferous mudstone overlie the barite-sulphide

exhalite (McClenaghan et al. 2006).

 The dominant footwall alteration is sericitic with subordinate Mg- and

Fe-chloritic alteration is restricted to the proximal footwall and extends

no more than a few hundreds of metres below the massive sulphide

lenses.

 Metal zoning is generally absent.

The deposits hosted by the California Lake Group differ from those in the

Tetagouche Group in that:

 The host sedimentary rocks may be dominantly pelagic rather than

volcanically derived.

32

 Oxide facies iron formation is absent; however, magnetite occurs within

the sulphides at some deposits, such as Caribou where it is attributed to

late-stage, vent-proximal replacement of sulphides (Goodfellow and

McCutcheon 2003).

 Footwall alteration is either less obvious or not as extensive as it is in

Nepisiguit Falls-hosted deposits. Beneath most deposits, Fe-rich chlorite

and disseminated sulphides occur for a short distance into the footwall,

but in the sediment-hosted deposits, at least some of this alteration may

be related to downward, rather than upward fluid flow. A silicified zone

is absent from most deposits but hanging wall sericitic alteration occurs

in many systems.

 Metal zoning is recognized.

33

Chapter 3

Geology of the Heath Steele area

The Heath Steele area is located in the Heath Steele belt, 37 km southwest of

Bathurst (Figure 2.1). The stratigraphic sequence of the Heath Steele belt comprises the

Miramichi and Tetagouche groups. Siltstones and shales of the Miramichi Group are overlain by carbonaceous shales and felsic volcanic and volcaniclastic rocks of the

Nepisiguit Falls Formation; that is in turn overlain by rhyolite tuffs, hyaloclastite, and tuffaceous sedimentary rocks of the Flat Landing Brook Formation (van Staal and Fyffe

1991, Moreton 1994, Lentz 1996a, 1997, Goodfellow et al. 2003, Peter et al. 2003a, van

Staal et al. 2003, Wilson 2015). The Heath Steele belt is bounded to the south by the east-west striking Heath Steele Fault zone (Figure 3.1), which juxtaposes the Nepisiguit

Falls Formation (north) against massive rhyolite and coarse-grained fragmental rocks of

Flat Landing Brook Formation (south) that are locally sheared (Wilson 1993, Lentz

1997).

3.1 Geologic setting

In the Heath Steel area (Figure 3.1), the volcanic part of Nepisiguit Falls

Formation consists of medium- to coarse-grained quartz-feldspar crystal-rich (10 to

45%), volcaniclastic rocks interlayered with quartz-feldspar crystal tuff. The volcanic rocks are interbedded with clastic and exhalative sedimentary rocks including quartzose siltstone and sandstone, shale, feldspathic wacke, and local massive sulphide and oxide facies iron formation. The clastic rocks are typically medium to dark green and commonly moderately to highly chloritic and locally sericitic. The most intense

34

chloritization occurs in proximity to massive sulphide lenses, where they are referred to as “chloritic tuffs” in mine terminology (McCutcheon et al. 1993, Wilson et al. 2015).

Figure 3.1. Geological map of the central part of the Heath Steele area showing location of B, E, F, and ACD zones (dark green box indicates the study area shown in the Figure 3.3, FLBF=Flat Landing Brook Formation, NFF=Nepisiguit Falls Formation; modified from Wilson et al. 2015).

The Flat Landing Brook Formation consists mainly of feldspar-phyric rhyolite and minor dacite flows that commonly exhibit perlitic or spherulitic textures, which are indicative of an original glassy groundmass. In addition to massive flows the Flat

Landing Brook Formation also includes volumetrically subordinate felsic hyaloclastites and pyroclastic rocks and thin intervals of interflow sedimentary rocks. Aphyric to sparsely feldspar- (and rarely quartz-) phyric rhyolite flows, flow-breccias, and hyaloclastites (Reids Brook Member) make up the bulk of the formation in the Heath

Steel area, whereas felsic crystal ± lithic ± vitric tuff and minor porphyritic felsic flows

35

are included in the Roger Brook Member (van Staal et al. 1991, McCutcheon et al. 1993,

Wilson et al. 2015).

Six economically significant massive sulphide deposits (ACD, B, E and F zones) are located along the Heath Steele belt hosted by sedimentary and volcanic rocks of the

Nepisiguit Falls Formation, which overlies the carbonaceous shales and siltstones of the

Patrick Brook Formation (the upper part of the Miramichi Group) and underlies felsic volcanic rocks of the Flat Landing Brook Formation (Figure 3.2; van Staal and Fyffe

1991, van Staal et al. 2003).

Figure 3.2. Generalized stratigraphic column of the Heath Steele area. MG=Miramichi Group, NFF=Nepisiguit Falls Formation, FLBF=Flat Landing Brook Formation (modified from de Roo et al. 1991, Moreton 1994, Lentz et al. 1997, Peter et al. 2003a).

The massive sulphide mineralization is spatially and temporally associated with iron formation (exhalative), and chlorite tuff (a weak or distal iron formation) that are

36

products of hydrothermal sedimentation (Peter et al. 2003a). This horizon is near the top of a sequence of interbedded laminated mudstone, siltstone, and crystal tuff (McBride

1976, Owsiacki 1979), which is overlain by massive quartz-feldspar crystal tuff. The true thickness of each stratigraphic unit is difficult to determine due to multiphase deformation that controls the morphology of the massive sulphide orebodies and the host rocks (Currie et al. 2003). All rocks along the Heath Steele belt have been metamorphosed to greenschist facies assemblages, with a peak temperature around

375°C and 5.8 kbars pressures (Currie et al. 2003, Peter 2003, Peter and Goodfellow

2003, Peter et al. 2003a, 2003b).

3.2 Heath Steele deformation

In general, the geology in the Heath Steele area strikes east-west and dips steeply

towards the north. Detailed structural work in the area (de Roo et al. 1990, 1991,

Moreton 1994) indicates that there are at least five deformation events, which are consistent with results elsewhere in the BMC. These deformational events can be summarized as follows:

D1): is the earliest deformation, it is represented by a penetrative composite foliation (S1), a stretching lineation (L1), and disharmonic, commonly noncylindrical asymmetrical folds (F1) (de Roo and van Staal 2003). D1 deformation is characterized by tight to isoclinal recumbent sheath folds indicative of high strain environment and developed in conjunction with thrusting (Moreton 1994). These folds have been documented at both outcrop- and mine-scale and have been invoked to account for younging reversals and local infolding of hanging wall units in the footwall. Zones of high D1 strain in the sedimentary and volcanic rocks are characterized by boudinage,

37

stratal disruption, sheath folds, mylonitization, spatially associated vein complexes, structural truncation of rock units, and repetitions of stratigraphy such that older units structurally overlie younger rocks (van Staal 1994, van Staal and Rogers 2000, de Roo and van Staal 2003).

D2): is mainly represented by steeply inclined or upright F2 folds, commonly accompanied by a crenulation cleavage (S2). F2 folds are generally tight to open structures in profile and are rarely associated with the marked disruption or boudinage of layering and other markers characteristic of D1 shear zones, indicating generally lower strain than that produced during D1 (de Roo et al. 2003). In general, these F2 folds have a moderate westerly plunge and a moderately south-southwest-dipping (approximately

45°) axial planar foliation (Moreton 1994). The enveloping surface to the F2 folds is

responsible for the tabular nature of the Heath Steele massive sulphide deposits. The intensity of the S2 foliation accompanying these folds is such that it generally obliterates evidence of S1 cleavage by transposition into a composite S1-2 schistosity (de Roo et al.

1991). The strike of S2 is variable in the Heath Steele area and ranges from 050° to 120°, mainly because of the effects of later deformations due to reorientation by later deformation (D3-5).

(D3-5): D3 deformation produced mostly open folds (F3) with a northwesterly striking axial plane cleavage S3, whereas D4 deformation produced mostly gentle, recumbent folds (F4). The D5 deformation appears to be localized, it is distinct in style, the orientation of the S5 cleavage is independent of its position of any fold. The fourth and fifth generations of folds (F4 and F5) have subvertical axial surfaces, and northwest- and northeast-plunging fold axes, respectively (de Roo et al. 1990, 1991).

38

3.3 Heath Steele E zone deposit

The Heath Steele E zone deposit (Figure 3.3) is a volcanogenic-sedimentary Pb-

Zn-Cu sulphide deposit located midway between the B and ACD zones. The Heath

Steele E zone deposit consists of two tabular subparallel sulphide lenses (North and

South lenses) that strike east-west dip 50 to 60 degrees to the south, the two lenses come close to the surface (sub-crop) and they merge at shallow depths ( see DDH S-806 on

Figure 3.4).

Figure 3.3. Map of the Heath Steele E zone showing drill core locations in black circles. The area of map is located on Figure 3.2. Geological section line A-A’ is presented in Figure 3.4 (modified from Wilson et al. 2015).

39

Figure 3.4. Schematic cross sections A-A’ of the Heath Steele E zone deposit. Line of Section is located on Figure 3.3 (modified from Trevali Mining Corporation).

The two sulphide lenses of the E zone are hosted by chloritic sedimentary rocks and crystal-rich volcanic/volcaniclastic rocks of the Nepisiguit Falls Formation. The

North lens is dominated by massive pyrite grading down-dip into magnetite iron formation, whereas the South lens is dominated by chalcopyrite-pyrrhotite (de Roo et al.

1990, 1991). Therefore, both massive sulphide lenses in the E zone deposit must have formed as a syn-genetic mineralization accumulated from exhalative solutions in subaqueous basin. The North lens (Pb-Zn rich and Cu poor) represents the distal zone, while the South lens (Cu rich and Pb-Zn present in economic quantities) is interpreted to be proximal to the feeder pipe (Whitehead 1973, Jambor 1979; Figure 3.5).

40

Figure 3.5. Zn-Pb-Cu base metal classification scheme applied to laboratory assay data from E zone VMS deposit (field boundaries are from Franklin et al. 1981).

The following description corresponds to the host rocks of the massive sulphide lenses from drill core investigated in the study area (Appendix A, Table 1.1 of section

1.6.1). For the purposes of this study, it was decided to refer to the rocks that overlie the

South lens as hanging wall rocks, rocks that underlie the North lens as footwall rocks, and to the rocks that lie between both lenses as middle-section rocks (Figure 3.4).

The footwall consists of light grey, coarse-grained quartz-feldspar crystal tuff

(Figure 3.6 and 3.7). White subhedral to euhedral feldspar crystals, ranging in size from

2 to 5 mm make up 10 to 15 % of the rock. Colourless sub-angular to sub-rounded quartz crystals in sizes ranging from 1 to 2 mm make up 10 to 15% of the rock. This unit is interbedded with thin (20-30 cm thickness) layers of dark grey fine grained siltstone.

41

Figure 3.6. Photograph of quartz-feldspar crystal tuff with euhedral feldspar (Fsp) and quartz (qtz) phenocrysts (0.1-1cm) from the footwall NFF of the E zone deposit. Photo is from drill core S-362 @185 m (core diameter 3.64 cm).

Figure 3.7. Photograph of quartz-feldspar crystal tuff with sub-hedral to euhedral pink

K-feldspar (Ksf) (1-3 mm) and anhedral phenocrysts of quartz (qzt) (1-3mm) from the footwall NFF of the E zone deposit. Photo is from drill core S-916 @209 m (core diameter 3.64 cm).

The sedimentary rocks are dark green, fine-grained siltstone with heavy oxidation along the foliation planes (Figure 3.8). In the footwall, alteration goes from intense K-feldspar-sericite (distal) to intense chlorite-sericite-pyrite (proximal) to the mineralization. The whole package is weakly to moderately deformed, and locally intense deformation is marked by a well-developed fabric.

42

Figure 3.8. Photograph of laminated siltstone (ash?) with sericite-chlorite alteration and strong Mn-oxide on the foliation planes. Photo is from drill core S-916 @ 198 m (core diameter 3.64 cm).

The middle-section part is made up of chloritic felsic volcanic rocks and sedimentary rock that are separated by a thin (60 to 90 cm thickness) layer of iron formation. The two volcanic layers are very similar and consist of dark green, medium- to coarse-grained, quartz-feldspar crystal tuff. K-feldspar crystals are white to pink, anhedral to subhedral crystals, ranging in size between 2 and 5 mm and account for 15 to

25% of the rock. Quartz occurs as colourless sub-angular to sub-rounded crystals ranging in sizes from 1 to 2 mm and account for up to 10 to 15% of the rock. Alteration in this zone goes from moderate chlorite to chlorite-carbonate to strong chlorite-pyrite- sericite towards mineralization, and disseminated pyrite close to the mineralization and

(or) the iron formation (Figure 3.9).

43

Figure 3.9. Photograph of quartz-feldspar crystal tuff from the middle-section from drill core S-362 @173.93 m. Showing K-feldspar (Kfs) crystals in sizes ranging between 2 and 5 mm width strong chlorite alteration and disseminated pyrite (py) (core diameter 3.64 cm).

The sedimentary rocks in the Middle-section consist of dark green, fine-grained, mica-rich mudstone and siltstone, with well-developed schistose textures. These rocks are also affected by a strong pervasive chlorite alteration, abundant quartz veining

(10%), and trace disseminated pyrite. Finally, the iron formation layer generally does not exceed one meter in thickness and it is predominantly fine-grained and commonly

laminated or brecciated. Iron-formation consists mainly of magnetic, hematite, (iron oxide), carbonate (siderite), and sulphide (disseminated pyrite) (Figure 3.10).

Figure 3.10. Photograph of a strongly chloritized tuff with disseminated pyrite (py), and thin layers of chloritic sedimentary rock (shale) in the upper part of the photo. Oxide

44

facies iron formation with brecciated texture in the lower part. Photo is from drill core S- 362 @172.50 m (core diameter 3.64 cm).

The hanging wall comprises light grey, medium- to coarse-grained massive quartz-feldspar crystal tuff and several layers of sedimentary rocks. Quartz crystals are equant, clear-white, mm-scale (up to 3 mm), evenly distributed throughout; however, locally the concentrations of crystals may constitute up to 10% of the unit. Feldspars are white, sub-euhedral to euhedral, and range from 1 to 5 mm, are sporadically distributed throughout, and account for up to 10 % of the unit (Figure 3.11). Weak foliation defined by thin bands of chloritic alteration, and moderate to strong sericite and chloritic alteration is developed in these rocks. Sedimentary rocks consist mainly of dark grey, fine-grained, well foliated/bedded siltstone with strong pervasive chlorite alteration and abundant quartz veining (10%).

Figure 3.11. Photograph of quartz-feldspar crystal tuff from the hanging wall of the E zone. Showing feldspar (Fsp) crystals in sizes ranging between 2 and 5 mm. Photograph is from drill core S-832 @ 25 m (core diameter 3.64 cm).

45

Chapter 4

Lithogeochemical interpretation

Lithogeochemical studies of igneous rocks in modern and ancient geological environments have been vital to understand magmatic and tectonic processes related to the geological evolution of the Earth (Piercey 2011). An important advance in this field was the creation of Alkali-total iron-magnesium (AFM) and the total alkali-silica (TAS) diagrams to define the magmatic affinity of volcanic rocks using major elements

(Na2O+K2O, FeO+Fe2O3, MgO, and SiO2); however, these large ion lithophile elements

(LILE) are commonly mobile under the low to high temperature seawater alteration processes associated with VMS systems. Consequently, reliable rock type and tectonic environment discrimination diagrams should be based on high field strength elements

(HFSE), such as Al, Zr, Y, V, Nb, Th, Ti, Ga, Sc, Hf, and Ta and most of the REE are immobile under most hydrothermal process (Winchester and Floyd 1977, Floyd and

Winchester 1978, MacLean and Barrett 1993, Rollinson 1993, Jenner 1996, Lentz

1996a, Ross and Bedard 2009).

Furthermore, various elemental ratios and indexes can be used to identify primary characteristics associated with the formation of these rocks or secondary alteration characteristics acquired during mineralization events (Piercey 2010). This lithogeochemical evidence can be crucial for the correct interpretation of regional to local geology as primary textures and mineralogy are commonly obliterated by subsequent deformation and (or) alteration processes (Lentz 1998).

46

4.1 Rock Type classification

A method of rock-type classification of fresh volcanic rocks using immobile elements to classify the differentiation products of subalkaline and alkaline magma series with minor and trace elements (Ti, Zr, Y, Nb, Ce, Ga, and Sc) was proposed by

Winchester and Floyd (1977). This method proposed discrimination diagrams to classify volcanic rock types in the three main magma series (tholeiitic, calc-alkalic, and alkalic) and can be applied to altered or metamorphosed rocks, since it is based on ratios of low- mobility HFSEs. The Zr/TiO2 versus Nb/Y discrimination diagram (Figure 4.1) illustrates the rhyodacitic/dacite to andesite composition of most of the footwall and hanging wall felsic volcanic rocks sampled and is consistent with Nepisiguit Falls

Formation compositions in the Heath Steele belt and elsewhere in the BMC (Whitehead and Goodfellow 1978, Lentz 1996a, 1997, 1999).

47

Figure 4.1. Nb/Y vs Zr/TiO2 discrimination diagrams applied to data collected from the 8 drill cores investigated at the Heath Steele E zone deposit. Field boundaries are from Winchester and Floyd (1977). (Alk-Bas=alkaline-basalt, Bsn/Nph=basanite/nephelinite, TrachyAnde=trachyandesite, Com/Pant=comendite/pantellerite).

48

4.2 Magmatic affinity

In order to distinguish between tholeiitic and calc-alkaline affinities in fresh and altered felsic volcanic rock suites, the Y/TiO2 versus Zr/TiO2 discrimination diagram

(Figure 4.2) has been shown to be useful. The diagram recognizes calc-alkaline (Zr/Y=7 to 25), transitional (Zr/Y=4.5 to 7.0), and tholeiitic (Zr/Y=2.0 to 4.5) affinities. The dacitic to rhyodacitic volcanic rocks in the E zone have Zr/Y values ranging between 2 and 7 indicative of a tholeiitic to transitional affinity, with increasing Zr/TiO2 fractionation, possibly indicative of Zr saturation or evolution to tholeiitic affinities

(Lentz 1998). This affinity is commonly associated with continental bimodal magmatism

(Lentz and Wilson 1997).

49

Figure 4.2. Y/TiO2 versus Zr/TiO2 discrimination diagrams for magmatic affinity for samples from the 8 drill cores investigated. Field boundaries are from Barrett and MacLean (1994). This diagram indicates a tholeiitic magmatic affinity for volcanic host rocks from Heath Steele E zone.

50

At the Heath Steele E zone deposit; there are no obvious compositional differences between footwall and hanging wall crystal tuff packages. The compositional similarity between the footwall and hanging wall rocks together with the presence of two subparallel sulphide lenses suggest the possibility of a fold-repeated sequence.

4.3 Tectonic settings

On the Nb versus Y (Figure 4.3) and Rb versus Nb+Y (Figure 4.4) tectonic discrimination diagrams, the felsic volcanic rocks from Heath Steele E zone fall across the volcanic arc and within-plate setting fields. This is consistent with the intracontinental ensialic back arc rift setting proposed for the BMC (van Staal et al.

1991, Lentz 1996a, van Staal et al. 2003, Wilson 2003, Goodfellow 2007, Wilson et al.

2015).

Previous models of the magmatic evolution during the Middle Ordovician indicate that volcanic activity began in the late Arenigian-early Llanvirnian with the eruption of felsic lavas and pyroclastic rock (Nepisiguit Falls Formation; van Staal et al.

1991). The abundance and within-plate setting of the felsic volcanic rocks (Figures 4.3 and 4.4), coupled with the lack of intermediate volcanic rocks, suggest generation from melting of thinned continental crust rather than extrusion in a mature island-arc or active continental margin (van Staal et al. 1991).

51

Figure 4.3. Nb (ppm) versus Y (ppm) tectonic discrimination diagram for volcanic and volcano-sedimentary rocks from the Heath Steele E zone. (VAG=volcanic arc granites, syn-COLG=syn-collisional granites, WPG=within-plate granites, ORG=ocean ridge granites. Field boundaries are from Pearce et al. (1984).

52

Figure 4.4. Rb (ppm) versus Y+Nb (ppm) tectonic discrimination diagram showing data from volcanic and volcano-sedimentary rocks from the Heath Steele E zone (VAG=volcanic arc granites, syn-COLG=syn-collisional granites, WPG=within-plate granites, ORG=ocean ridge granites. Field boundaries are from Pearce et al. (1984).

53

Chapter 5

Chemostratigraphy

The rocks in the Heath Steele are characterized by a strong, penetrative poly- phase deformation, intense alteration, and faulting. Collectively, these factors have all but obscured the primary textures in the volcanic and sedimentary host rocks at the E zone deposit. For this reason, chemostratigraphy is employed to elucidate the volcano- sedimentary stratigraphy of the rocks hosting the deposit.

5.1 Chemostratigraphic analysis of the felsic volcanic rocks

In this study, chemostratigraphic units are defined in terms of a point or range of values of selected elements, or a combination of elements. Previous work by Lesher et al. (1986), Maclean and Barrett (1993), Jenner (1996), Lentz (1996a,1996b, 1996c,

1997), Lentz and Wilson (1997), Lentz et al (1996, 1997), and Barrett and MacLean

(1998) show that the use of immobile element ratios can improve lithostratigraphic correlations in altered felsic volcanic sequences that host massive sulphide deposits. For example the Zr/TiO2 ratio (Winchester and Floyd 1977, MacLean and Barrett 1993) serves as a reliable indicator of fractionation even in intensely altered rocks, and is a robust chemostratigraphic discriminator (Lentz and Wilson 1997, Lentz 1997).

Key chemostratigraphic differences between structural hanging-wall and footwall sedimentary rocks at the Heath Steele E zone are: Al contents in the hanging wall sedimentary rocks exceed 14 wt.% Al2O3, whereas footwall sedimentary rocks have

<14 wt.% Al2O3. Likewise, Ti contents in the hanging wall sedimentary rocks (>1 wt.%

TiO2) are greater than the footwall (<0.8 wt.% TiO2). While average values of Zr/Al2O3

54

= 0.0042 for the hanging-wall sedimentary package and an average Zr/Al2O3 = 0.0025 for the footwall sedimentary rocks where obtained.

The hanging wall volcanic rocks have average Y and Zr contents of 34 ± 2 and

125 ± 5 ppm, respectively, whereas footwall volcanic rocks have average Y and Zr contents of 34 ± 5 ppm and 114 ± 8 ppm, respectively. The Zr/TiO2 and Zr/Y ratio are used to differentiate volcanic units in the hanging wall from those in the footwall; in the hanging wall rocks average values for Zr/TiO2 = 0.054 ± 0.010 and a Zr/Y = 3.9 ± 2.0, whereas footwall rocks have average Zr/TiO2 = 0.052 ± 0.015 and average Zr/Y = 3.7 ±

2.0; finally, the average Zr/Al2O3 in the hanging-wall and footwall volcanic rocks is

0.0021 ± 0.001 (Table 5.1; Figure 5.1; Appendix D).

55

Table 5.1. Average immobile elements and immobile-element ratios of hanging wall and footwall rocks intersected by drill cores in the Heath Steele E zone deposit.

Geochemical Structural Rock type Average STD discriminant position Hanging wall 14.3 ±2.4 Al O (wt.%) 2 3 Footwall 12.6 ±2.9 Hanging wall 1.06 ±0.32 TiO (wt.%) 2 Footwall 0.72 ±0.45 Hanging wall 90 ±10 Cr (ppm) Footwall 86 ±12 Hanging wall 168 ±40 Sedimentary V (ppm) Footwall 164 ±40 Hanging wall 0.0042 ±0.002 Y/TiO 2 Footwall 0.0072 ±0.004 Hanging wall 0.0042 ±0.001 Zr/Al O 2 3 Footwall 0.0025 ±0.001 Hanging wall 0.032 ±0.010 Zr/TiO 2 Footwall 0.030 ±0.020

Hanging wall 11 ±4 Nb (ppm) Footwall 11 ±3 Hanging wall 34 ±2 Y (ppm) Footwall 34 ±5 Hanging wall 125 ±5 Zr (ppm) Footwall 114 ±8 Hanging wall 3.9 ±2.0 Volcanic Zr/Y Footwall 3.7 ±2.0 Hanging wall 0.0021 ±0.0001 Zr/Al O 2 3 Footwall 0.0021 ±0.0001 Hanging wall 0.054 ±0.010 Zr/TiO 2 Footwall 0.052 ±0.015 Hanging wall 1.13 ±0.20 Th/Nb Footwall 0.95 ±0.20

56

57

Figure 5.1. Down hole correlation of Zr/TiO2 with rock type for 5 selected drill cores from the Heath Steele E zone. See Appendix A for detailed logs and Appendix C for pXRF data.

Comparing values of this study with previous regional data (Wilson 1993, Lentz

1996a, Lentz et al. 1997) for sedimentary rocks (quartzose siltstone and shale) of the

Miramichi Group, which have high Al2O3 (>15 wt.%) and TiO2 (0.9 to 1.1 wt.%) (Lentz et al. 1997), and those reported for the Nepisiguit Falls Formation (<15 wt.% Al2O3 and

0.4 to 0.9 wt.% TiO2) (Lentz et al. 1997) the geochemical data of sedimentary rocks of the Heath Steele reveals that they are more consistent with those reported for the

Nepisiguit Falls Formation. The high content of Al2O3 and TiO2 (generally associated with clay/mica components; Lentz et al 1996, Lentz and Wilson 1997) in some levels of the stratigraphic column of the E zone deposit are indicative a mixture of and/or reworked terrigenous from the Miramichi Group with rocks directly from hydrogenous/hydrothermal sedimentation.

This idea is supported by the high concentration of Cr and V (e.g., Cr>80 ppm and V>160 ppm) reflecting a mature provenance (Lentz and Wilson 1997, Lentz et al

1996), the presence of thin layers of exhalite (iron formation; see section 5.2) directly from a hydrothermal source, low Fe/Mn (oxidized signature), elevated base metal

(Zn+Pb) values, and strong enrichment in other exhalative components (Fe, Mn, B, and

P), e.g., result of local disruption of the stratified ocean by concurrent volcanism and/or convective overturn resulting from the thermal perturbations in the basin (Lentz 1996a).

Geochemical data from the Nepisiguit falls Formation at the Brunswick No 6 deposit (Wills et al. 2006), e.g. Nb = 16 ± 3 ppm, Y = 43 ± 3 ppm, Th = 16.3 ± 4.6 ppm, whereas Al2O3 = 14.49 ± 1.15 wt.%, and TiO2 = 0.57 ± 0.12 wt.% (Zr/TiO2 = 0.053 ±

0.004, Y/TiO2 = 0.009 ± 0.003). These values are very similar to those obtained in this study. However, the Nepisiguit Falls Formation at the Heath Steele E zone can be

58

divided into five volcanic chemo-stratigraphic units on the basis of distinct Zr/TiO2,

Al2O3/TiO2, Al2O3/Zr, Zr/Nb, and Zr/Y ratio (Table 5.2).

Table 5.2. Shows the average immobile element ratios of five chemostratigraphic volcanic units at the E zone deposit.

Structural Unit Zr/TiO Zr/Al O TiO /Al O Zr/Nb Zr/Y position 2 2 3 2 2 3 Footwall Average 0.052 0.0020 0.036 11.8 3.9 1 STD 0.010 0.0001 0.010 2.4 0.7 Average 0.040 0.0022 0.040 10.7 3.3 2 Middle STD 0.010 0.0001 0.011 1.8 0.3 Section 3 Average 0.041 0.0020 0.042 10.7 3.4 STD 0.010 0.0001 0.010 1.8 0.3 Average 0.054 0.0020 0.036 11.5 3.7 4 Hanging STD 0.010 0.0001 0.012 1.7 0.5 wall Average 0.058 0.0020 0.036 11.5 4.3 5 STD 0.010 0.0001 0.010 3.0 1.3

Table 5.2 shows that there are similarities between units in different structural levels (i.e., repetition of units in different structural levels). On a TiO2 versus Zr discrimination diagram (Figure 5.2), the five volcanic chemostratigraphic units form linear mass change alteration arrays, where Unit 1 and 4 have very similar Zr/TiO2 values of 0.058 to 0.060. Likewise units 2 and 3 have similar Zr/TiO2 values of 0.040 to

0.043, whereas unit 5 and the iron formation (IF) have Zr/TiO2 ratios of 0.052 and

0.015, respectively (structurally IF is situated between units 2 and 3).

59

Figure 5.2. TiO2 versus Zr plot demonstrates that units 1 and 4 have similar Zr/TiO2 ratios, likewise units 2 and 3 have similar average in Zr/TiO2; each of the alteration mass

change lines forms a distinct cluster of samples, which is definitive for each of the geochemical units.

The introduction of another immobile element as a denominator to the immobile elements in bivariate plots minimizes the effect of alteration-controlled mass balance variation (Barrett and MacLean, 1998). As demonstrated in Figure 5.3, the introduction of Al2O3 as a denominator to both Zr and TiO2 causes the samples from each chemo- stratigraphic unit to plot in a cluster rather than linear arrays. Three clusters are clearly identified, the first consist of iron formation, the second includes units 2 and 3, and the third includes units 1, 4 and 5. The third set demonstrates similar behaviour between footwall samples (unit 1) and hanging wall samples (units 4 and 5).

60

Figure 5.3. Zr/Al2O3 versus TiO2/Al2O3 discrimination diagram for data (see appendix B) from the Heath Steele E zone. Unit 1 corresponds to footwall samples, units 2 and 3 correspond to middle section samples, and units 4 and 5 correspond to hanging wall samples.

Finally, lithostratigraphic correlation of the volcanic units along the section A–

A’ is further refined by chemostratigraphic correlation of units 1 to 5 in drill cores S-

362, S-806, S-832, S-916, and HS17-005 (Figure 5.4 and Appendix E).

61

62

Figure 5.4. Average Zr/TiO2 values for stratigraphic units at the Heath Steele E zone. See Appendix A for detailed logs and Appendix C for pXRF data.

5.2 Exhalites (iron formation) analysis

According to previous work (Saif 1980, Saif 1983, Peter and Goodfellow 1996,

Peter and Goodfellow 2003, Peter et al. 2003a, 2003b), there are two major types of iron formation (exhalites) in the BMC. Type 1 is the carbonate-oxide-silicate iron formation that occurs on the Brunswick horizon (e.g., Brunswick No. 12 and 6, Austin Brook,

Heath Steele, Key Anacon deposits) and spatially related to massive sulphide mineralization at the top of the Nepisiguit Falls Formation; this key unit is more or less a time-stratigraphic horizon (Peter and Goodfellow 1996, Peter et al. 2003a). Type 2 consists of widely distributed Fe-Mn oxides that are not spatially or temporally associated with massive sulphide deposits (Goodfellow et al. 2003).

Type 1 iron formations are divided into five major facies based on the

mineralogy, i.e. oxide, carbonate, silicate, sulphide, and sulphate facies (Peter and

Goodfellow 1996, Peter et al. 2003a, Slack 2012). The facies of iron formation produced is dependent on bulk compositions and range of physicochemical parameters of ambient seawater and venting hydrothermal fluids involved in the mineralizing system, e.g. pressure, temperature, pH, ironic strength, fugacity of oxygen, sulphur, and carbon dioxide (Peter 2003). Oxide facies iron formation consists of jasper, hematite iron formation, and magnetite iron formation; carbonate facies, includes one or more Fe-Mg-

Ca-Mn carbonates, such as siderite, ankerite, dolomite, calcite, rhodochrosite, and kutnahorite; Silicate facies, comprises iron-rich minerals (such as greenalite and stilpnomelane), magnesian minerals (such as talc and chlorite), manganese-rich minerals

(such as spessartine garnet), and boron-rich minerals (such as tourmaline); Sulphide facies iron formation is composed of pyrite and (or) pyrrhotite with only minor base-

63

metal sulphides (chalcopyrite, sphalerite, galena); and sulphate facies, which comprises barite and, in a limited number of deposits, anhydrite and gypsum (Slack 2012; Figure

5.5).

Figure 5.5. Cross sections of volcanogenic massive sulphide deposits showing types and morphologies of exhalites. A) Proximal jasper and hematitic chert overlying mound-like deposit. B) Proximal and distal (regionally extensive) iron formation occurring immediately above and along strike from deposits, typical of the BMC. C) Zoned iron formation, grading outward from inner sulphide facies to carbonate facies to oxide facies to silicate-facies. D) Iron formation occurring immediately above a sheet-like deposit (from Slack 2012 and references therein).

64

The iron formation from Heath Steele E zone (unit IF) correspond to Type 1 iron formation (carbonate-oxide-silicate-sulphide iron formation) and are shown to correspond to carbonate facies, which is consistent with observations reported in Peter and Goodfellow (1996, 2003). The carbonate facies consists of interlaminated siderite

(see Ca-Fe-Mn discrimination diagram Figure 5.6) and magnetite.

Figure 5.6. Ca-Fe-Mn discrimination diagram for the classification of carbonate minerals using bulk geochemical data from iron formation (IF) and massive sulphide (MS) samples (this study) from Heath Steele E zone (from Peter et al. 2003b).

In general, Type 1 iron formations are zoned laterally such that the thickness and relative proportion of hydrothermal sediment increase with proximity to sulphide deposits, (i.e., proximal exhalites are thickest directly above sulphide bodies and become progressively thinner with increasing distance from a deposit) (Peter and Goodfellow

1996, 2003, Goodfellow 2007). Iron formation samples from the E zone are classified as

65

distal exhalites (i.e., distal to the site of hydrothermal venting) on the basis on the Fe/Mn versus Zn+Pb contents (Figure 5.7).

Figure 5.7. Fe/Mn versus Zn+Pb discrimination diagram for the determination of relative distal from hydrothermal vent site for iron formation (IF) samples from the Heath Steele E zone (modified from Wills et al. 2006).

At the Heath Steele E zone, there is a spatial relationship between the iron formation and the base-metal sulphide deposit. The iron formation intervals occur in the immediate vicinity of the mineralization and apparently at the same stratigraphic horizon. The Al, Fe, and Mn contents of the massive sulphide and iron formation from the E zone are consistent with a significant hydrothermal input (Figure 5.8).

66

Figure 5.8. Al-Fe-Mn discrimination diagram of Boström (1973) showing whole-rock composition of exhalite rocks from the Heath Steele E zone. Fields and average

compositions of modern sediments (from Peter and Goodfellow 1996, and references therein). DSDP=Deep Sea Drilling Project, EPR=East Pacific Rise, IF=iron formation, MS=massive sulphide.

The Fe/Ti versus Al/(Al+Fe+Mn) discrimination diagram (Figure 5.9) assess the variable hydrothermal (Fe and Mn) and terrigenous (Ti and Al) input in sedimentary rocks. Sedimentary samples from E zone fall in the terrigenous field, whereas exhalites and massive sulphide samples plot toward the hydrothermal end-member primarily because of their elevated Fe contents. This plot shows that the iron formation samples span continuous mixing trend between hydrothermal sediment, terrigenous, and pelagic sediment.

67

Figure 5.9. Plot of Fe/Ti versus Al/(Al+Fe+Mn) for sedimentary rocks from the Heath Steele E zone. Field boundaries are from Boström (1973). Curved line represents the ideal mixing between Al-free hydrothermal sediment and pelagic or terrigenous

sediments. SED HW= hanging wall sedimentary rocks, SED FW= footwall sedimentary rock, IF= iron formation (oxide-carbonate facies), and MS=massive sulphide.

The MnO versus Fe/Mn discrimination diagram reflects the redox (Eh) condition and subsequent precipitation of mineral phases during exhalative activity in a stratified basin: a high Fe/Mn indicates anoxic conditions, whereas low Fe/Mn reflects an oxidizing setting (Whitehead, 1973). The redox conditions may be controlled by water depth, e.g., with the deepest parts of the basin having the most reducing (lowest Eh) conditions. The carbonate and silicate facies iron formation of the Heath Steele E zone likely formed in deep water under moderately reducing conditions (Figure 5.10).

68

Figure 5.10. MnO versus Fe/Mn discrimination diagram for clastic and hydrothermal sedimentary rocks from Heath Steele E zone. SED HW= hanging wall sedimentary rocks, SED FW= footwall sedimentary rock, IF= iron formation (oxide-carbonate

facies), and MS=massive sulphide (field boundaries from Whitehead 1973).

Based on the chemostratigraphic analysis in this study, volcanic rocks of the hanging-wall and footwall have similar compositions implying that they represent the same volcanic unit, which implies that the ore body was folded creating two lenses (the

South and North lenses). Given their relative base-metal contents, the South lens represents the more proximal part of the deposit, whereas the North lens represents mineralization deposited distal to the vent zone. This hypothesis is supported by the distribution of iron formation and its association with sulphide deposits, elsewhere in the

BMC iron formation typically overlies and extends laterally for several kilometers beyond the massive sulphide lenses.

69

The model proposed for the formation of the E zone deposit is as follows:

1) A small-volume magma body supplied heat and metal-bearing fluids that were channeled to surface via a synvolcanic fault (feeder zone). Hydrothermal venting on the sea floor a massive sulphide mound (black smoker), which formed under anoxic conditions (Figure 5.11A). The hydrothermal fluids formed plume from which minerals precipitated locally and partially into the seawater column and then settled on the sea floor, forming the massive sulphide lens and the iron formation stratigraphically above and distal to the sulphide lens.

2) The E zone host sequence underwent a poly-phase deformation history, which locally resulted in complex folding (Figure 5.11C). The distribution of the iron formation intervals relative to massive sulphide intervals suggests that the sequence is overturned towards the north.

3) The massive sulphide was concentrated in the northern limb of the fold, which was later eroded to its present level of exposure (Figure 5.11D).

70

Figure 5.11. Schematic evolution of the Heath Steele E zone. A) Conceptual control of Brunswick-type deposits (from McCutcheon and Walker 2019 and references therein). B-D) Simplified schematic representation of the E zone through time during its post formation tectonic evolution.

71

Chapter 6

Hydrothermal Alteration trends

Hydrothermal alteration is associated with most deposits in the BMC, although hydrothermal alteration has been documented in detail for only a few deposits. These include the Brunswick No. 12 (Luff et al. 1992, Lentz and Goodfellow 1993, 1994a,

1996), Brunswick No. 6 (Nelson 1983, Lentz 1996a, 1997, Yang et al. 2003), Heath

Steele (Wahl 1978, McDonald 1983, Lentz et al. 1997), Halfmile Lake (Yang et al.

2003), and Caribou deposits (Goodfellow 2007).

In most deposits, the vent complex is underlain by a sulphide feeder zone that may extend up to several hundreds of metres into the footwall. This feeder zone is characterized by disseminated and vein sulphides (primarily pyrite, chalcopyrite and pyrrhotite ± subordinate sphalerite, galena etc.), silicates, and carbonates hydrothermally altered volcanic and (or) sedimentary rocks of the footwall (Goodfellow 2007). Footwall hydrothermal alteration can be widespread (1-5 km laterally and hundreds of metres vertically) and zoned with respect to the paleo-seafloor and the hydrothermal vent. From core to periphery these zones (Figure 6.1) are: Zone 1 is characterized by quartz + Fe- rich chlorite, Zone 2 is characterized by Fe-rich chlorite + phengite +/- pyrite, Zone 3 is dominated by Fe-Mg-chlorite + phengite + albite, and Zone 4 is characterized by albite

+ Mg-rich chlorite. In general, sulphides, chlorite, and phengite increase, while feldspars decrease with increasing proximity to the core of the hydrothermal fluid upflow zone

(Luff et al. 1992, Lentz and Goodfellow 1993, 1994a, 1994b, Goodfellow 2007).

72

Figure 6.1. Schematic diagram showing distribution of hydrothermal alteration zones related to volcanogenic massive sulphide deposits in the Bathurst district (from Lentz and Goodfellow 1994a).

The volcanic and sedimentary rocks at the Heath Steele, like elsewhere in the

BMC, have been subjected to low- to medium-grade greenschist facies regional metamorphism, and have locally been modified by hydrothermal alteration processes.

Both of these processes may lead to partial or complete modification of the primary rock mineral assemblage. In general, all of the host sequences of the Heath Steele E zone have undergone variably intense K and Mg metasomatism, resulting in alteration assemblages dominated by chlorite, quartz, and sericite. In addition, silicification and zones of clay alteration are also prominent and grade away from the mineralization into regional greenschist metamorphic assemblages (Hamilton et al. 1993).

Studies on the alteration have been focused mostly on the structural hanging wall and footwall using lithogeochemical data to identify key element associations and alteration assemblages. In this study several alteration indices including select ore-

73

related elements (i.e., Cu, S, Ba, etc.), and element ratios (i.e., Fe/Mn, Fe/K, K/Al, and

Ba/Al, the chlorite-carbonate-pyrite index (CCPI)

(CCPI=100(MgO+Fe2O3)/(MgO+Fe2O3+CaO+Na2O), and the Ishikawa Alteration index

[AI=100(K2O+MgO)/(K2O+Na2O+CaO+MgO) were tested in evaluating alteration of

VMS host rocks at the Heath Steele E zone (see Appendix F).

The major-element Fe/Mn ratio of a sedimentary rock is a good relative measure of the degree of oxygenation of the basin in which the sediment was deposited, with higher Fe/Mn indicative of more anoxic conditions, low Fe/Mn indicative of oxic conditions, and intermediate Fe/Mn ratios reflecting a transitional environment (Lentz

1996a and references therein). At Heath Steele, the Fe/Mn ratio has been shown to be an effective lithogeochemical index (Whitehead 1973) within the host rocks alteration

system, as well as within the exhalative sedimentary rocks (Lentz 1996a).

In cross section A-A’ (Figure 6.2) the high Fe/Mn ratio in sedimentary rocks of the Nepisiguit Falls Formation is an indicator of anoxic conditions during the deposition of sedimentary rocks. The greatest concentrations of Fe/Mn are present in the sedimentary rocks that host the massive sulphide (e.g., Fe/Mn increase with proximity to massive sulphide); this means that both sedimentary rocks and massive sulphides formed under reducing conditions.

74

75

Figure 6.2. Fe/Mn values of the Nepisiguit Falls Formation. From select drill cores at the Heath Steele E zone.

To define and visualize alteration trends in the Heath Steele E zone, the alteration box plot of Large et al. (2001) was modified to incorporate elements deemed to be reliably determined with the pXRF used in this study. Although Na2O and MgO are commonly used to fingerprint alteration types in VMS system (Sack and Lewis

2013), these could not be analysed with the pXRF used in this study.

The AI (Ishikawa Alteration index) and the CCPI (chlorite-carbonite-pyrite index) were modified by removing Na and Mg (not detected with the pXRF used in this study) according to the methodology of Sack and Lewis (2013): AI=100(K)/(K+Ca) and

CCPI=100(Fe+Mn)/(Fe+Mn+Ca+K). Removal of Na means the index under estimates albitic alteration, whereas removal of Mg likely results in an under estimation of chlorite alteration (Large et al. 2001). Consequently, this modified formula tends to slightly overestimate the Al value of a given sample (Figure 6.3a and 6.3b; Sack and Lewis

2013).

76

Figure 6.3. Alteration box plot modified for use with the suite of elements acquired with the pXRF used during this study. Field boundaries modified from Large et al. (2001). a) hydrothermal trends and b) diagenetic trends. Numbered trends 1 through 10 are explained in text (ser=sericite, chl=chlorite, py=pyrite, carb=carbonate, ab=albite, kfeld= K-feldspar, calc=calcite, ep=epidote).

Trends 1 to 6 on Figure 6.3a are hydrothermal trends; Trend 1 relates to a weak sericite alteration at the margins of a hydrothermal system in felsic volcanic host rocks

(both hanging wall and marginal footwall to ore). Trend 2 reflects intense sericite- chlorite ± pyrite alteration typical of the proximal footwall alteration system to a VHMS

77

deposit, within both felsic and mafic volcanic host rocks, whereas Trend 3 reflects chlorite ± sericite ± pyrite alteration typical of chlorite-dominated footwall alteration either in felsic or mafic volcanic rocks. Trend 4 corresponds to chlorite-carbonate alteration typically developed immediately adjacent to massive sulphide lenses in felsic or mafic footwall rocks. Trend 5 reflects sericite-carbonate alteration in the immediate hanging wall of massive sulphide or along the favourable stratigraphic host unit. Trend 6 corresponds to K feldspar-sericite alteration developed locally within felsic footwall volcanic rocks, but is not particularly common (Large et al. 2001).

The diagenetic trends (7 to 10 on Figure 6.3b) are; Trend 7 is albite-chlorite alteration typical of seawater interaction at low temperatures, whereas Trend 8 is an epidote-calcite ± albite alteration commonly developed in intermediate and mafic volcanic rocks. Trend 9 is K-feldspar-albite alteration, where K feldspar is replaced by albite by an exchange reaction. Finally, Trend 10 is a paragonitic sericite-albite alteration, which is a diagenetic trend commonly recorded in the hanging-wall volcaniclastic rocks (Large et al. 2001).

The alteration box plot applied to data from the Heath Steele E zone is presented in Figure 6.4. The hydrothermal alteration varies from K feldspar-sericite developed in the more distal felsic volcanic footwall rocks to intense sericite-chlorite ± pyrite alteration typical of the proximal volcanic and sedimentary footwall rocks. In the case of the sedimentary hanging and volcanic hanging wall rocks, strong chlorite-pyrite to chlorite-carbonate alteration typically developed immediately adjacent to massive sulphide lenses the oxide facies iron formation.

78

Figure 6.4. Alteration box plot modified for use with the suite pXRF acquired data collected during this study.). It shows hydrothermal trends for drill core samples from the Heath Steele E zone (SED HW = hanging wall sedimentary rocks, SED FW = footwall sedimentary rocks, VOL HW = hanging wall volcanic felsic rocks, VOL FW = footwall volcanic felsic rocks, IF=iron formation, ser=sericite, chl=chlorite, py=pyrite, carb=carbonate, ab=albite, kfeld= K-feldspar, calc=calcite, ep=epidote). Modified from Large et al. (2001).

79

Chapter 7

Conclusions and Recommendations

The results presented in the present study supports previous observations that lithogeochemical and chemostratigraphy techniques are important tools in mining exploration and stratigraphic correlation. This information provides a vast amount of geochemical information useful for recognizing different rock types, the magmatic affinity for volcanic rocks or provenance of sedimentary rocks, and helps with stratigraphic correlations even if the rocks have been subject to strong alteration and metamorphism. Finally, these techniques are useful in the exploration field since it aids with a detailed characterization of the hydrothermal zoning of the footwall and hanging wall rocks that host massive sulphide deposits.

Conclusions

1. Lithogeochemistry of felsic volcanic rocks (crystal tuff) of the Nepisiguit Falls

Formation hosting the Heath Steele E zone deposit are compositionally uniform, according to the Zr/TiO2 versus Nb/Y discrimination diagram where they plot mostly in the rhyodacite/dacite field. According to the Y/TiO2 versus Zr/TiO2 discrimination diagram the crystal tuff samples have a tholeiitic affinity. The tectonic setting of these rocks is interpreted to be transitional between volcanic arc and within-plate settings, which is consistent with studies from elsewhere in the BMC.

2. Sedimentary rocks of the Nepisiguit Falls Formation at the E zone deposit have variation in of Al2O3 and TiO2 contents (12 to15 wt.% Al2O3 and 0.8 to 1.1 wt.%

TiO2), this is consistent with a mixed protolith consisting of mixed or reworked

80

terrigenous from the Miramichi Group with hydrogenous/hydrothermal or pelagic sedimentation synchronous with deposition of the tuffaceous rocks. This interpretation is supported by the high concentration of Cr and V (e.g., Cr>80 ppm and V>160 ppm) reflecting a mature provenance (Lentz and Wilson 1997, Lentz et al 1996), the presence of thin layers of exhalites further supports hydrothermal input.

3. Chemostratigraphy shows that the footwall and hanging wall crystal tuff (cross section A-A’; Figure 3.4) at the E zone are essentially identical, suggesting the possibility that the volcanic units on both side of the sulphide body may be fold-repeats of the same unit. Immobile element analysis has identified at least 5 volcanic units in the host sequence; however, two of these units appear in both the footwall and hanging wall.

The units are characterized by their unique Al2O3/TiO2, Al2O3/Zr, TiO2/Nb, TiO2/Y,

Zr/TiO2, and Zr/Nb ratios (Table 5.2). Of these the most important and useful is Zr/TiO2.

4. The recognition of footwall units repeated in the hanging wall supports the interpretation that the two massive sulphide lenses originate as a single folded massive sulphide lens. The South lens (closest to the surface) corresponds to the most vent proximal of the two, whereas the North lens (located at the deepest section) corresponds to the more distal part of the sulphide lens.

5. The iron formation is separated by felsic units 2 and 3 and corresponds to a typical carbonate-oxide-silicate iron formation facies. Geochemical vectoring, using hydrothermal components (Fe/Mn versus Zn+Pb) above the stockwork stringer sulphide complex. The iron formation likely formed in deep water under moderately reducing conditions.

81

6. Hydrothermal alteration in the footwall is dominated by K feldspar-sericite alteration distal to mineralization, and intense sericite-chlorite ± pyrite alteration in areas more proximal to the stringer sulphide (stockwork) veins. Hanging wall rocks are variably altered from chlorite-pyrite to chlorite-carbonate in areas immediately adjacent to massive sulphide lenses or exhalative rocks (iron formation).

82

Recommendations

For future exploration in the Heath Steele E zone area it is recommended that:

1) An expanded lithogeochemical and chemostratigraphy program be conducted using more drill cores. For example, lengthening the section A-A’ to the south and north

(mainly north towards F zone; Figure 7.1), this would enable better correlation and boundary of the massive sulphide lenses. This research would also improve define the chemical zonation in exhalite sediments (iron formation), and would allow a better understanding of the massive sulphide distribution with respect to the hydrothermal vent site.

2) Carry out a drill program in the gap between the drill holes HS17-005 and the

HS18-001 drill (Figure 7.1) to deduce if there is a continuity of either lens.

Figure 7.1. Geological map of the Heath Steele E zone deposit (modified from Wilson et al. 2015).

83

3) Further studies should be conducted to the north (between the E and F zone;

Figure 7.1) to help resolve some of the stratigraphic and structural complexities in the E zone: to determine whether the chemostratigraphic framework remains intact, and to further refine the correlations among deposits throughout the Heath Steele belt.

4) From the research presented in this study, the pXRF has demonstrated significant accuracy and precision for the suite of elements required to assess rock type and hydrothermal alteration. Future use of the pXRF in rock geochemistry surveys (rock, soil, and stream sediment surveys) is recommended in the search for elevated contents of base metals and favourable indicators of VMS-type hydrothermal alteration.

84

References Barrett, T.J. and MacLean, W.H., 1994. Chemostratigraphy and hydrothermal alteration

in exploration for VHMS deposits in greenstones and younger volcanic rocks. In

Alteration and Alteration Processes Associated with Ore-forming Systems.

Edited by D.R. Lentz. Geological Association of Canada, Short Course Notes,

vol. 11, p. 433-467.

Barrett, T.J. and MacLean, W.H., 1998. Volcanic sequences, lithogeochemistry and

hydrothermal alteration in some bimodal VMS systems. In, Volcanic-Associated

Massive Sulphide Deposits: Processes and Examples in Modern and Ancient

Settings. Edited by C.T. Barrie and M.D. Hannington. Reviews in Economic

Geology, vol. 8, p. 101-131.

Barrie, C.T. and Hannington, M.D., 1999. Classification of volcanic-associated massive

sulfide deposits based on host-rock composition. Reviews in Economic Geology,

vol. 8, p. 1-11.

Boström, K., 1973. The origin and fate of ferromanganoan active ridge sediments.

Stockholm Contributions to Geology, vol. 27, p. 147-243.

Currie, K., and van Staal, C.R., 1999. The assemblage stilpnomelane-chlorite-phengitic

mica: A geothermobarometer for blueschist and associated greenschist terranes.

Journal of Metamorphic Geology, vol. 17, p. 613-620.

Currie, K.L., Van Staal, C.R., Peter, J., and Rogers, M., 2003. Conditions of

metamorphism of the main massive sulfide deposits and surrounding host rocks

in the Bathurst Mining Camp. Economic Geology Monograph, vol. 11, p. 65-78.

85

de Roo, J.A. and van Staal, C.R., 1991. The structure of the Half Mile Lake region,

Bathurst Camp, New Brunswick. In Current Research, Part D. Geological Survey

of Canada, Paper 91-1D, p. 179-186. de Roo, J.A., and van Staal, C.R., 2003. Sulfide remobilization and sulfide breccias in

the Heath Steele and Brunswick deposits, Bathurst Mining Camp, New

Brunswick. Economic Geology Monograph, vol. 11, p. 479-496. de Roo, J.A., Moreton, C., Williams, P.F., and van Staal, C.R., 1990. The structure of

the region, Bathurst Camp, New Brunswick. Atlantic

Geology, vol. 26, p. 27-41. de Roo, J.A., Williams, P.F., and Moreton, C., 1991. Structure and evolution of the

Heath Steele base metal sulfide orebodies, Bathurst Camp, New Brunswick,

Canada. Economic Geology, vol. 86, p 927-943.

Downey, W.S., 2005. The geological setting, petrology and facies analysis of the

Nepisiguit Falls Formation, Bathurst Mining Camp: an example of a deep

submarine pyroclastic eruptive sequence. Unpublished M.Sc. thesis, University

of New Brunswick, Fredericton, NB, 294 p.

Downey, W.S. and Lentz, D.R., 2006. Modeling of deep submarine pyroclastic

volcanism: a review and new results. Geoscience Canada, vol. 33, p. 5-24.

Downey, W.S., McCutcheon, S.R., and Lentz, D.R., 2006. A physical volcanological,

chemostratigraphic, and petrogenetic analysis of the Little Falls member,

86

Tetagouche Group, Bathurst Mining Camp, New Brunswick. Exploration and

Mining Geology, vol. 15, p. 77-98.

Fisher, L., Gazley, M.F., Baensch, A., Barnes, S.J., Cleverley, J., Duclaux, G., 2014.

Resolution of geochemical and lithostratigraphic complexity: a workflow for

application of portable X-ray fluorescence to mineral exploration. Geochemistry:

Exploration, Environment, Analysis, vol. 14, p. 149-159.

Floyd, P.A., and Winchester, J.A., 1978. Identification and discrimination of altered and

metamorphosed volcanic rocks using immobile elements. Chemical Geology,

vol. 21, p. 291-306.

Franklin, J.M., 1997. Lithogeochemical and mineralogical methods for base metal and

exploration. In Proceedings of Exploration 97: Fourth Decennial

International Conference on Mineral Exploration. Edited by A.G. Gubins.

Exploration Geochemistry, vol. 28, p. 191-208.

Franklin, J.M., Lydon, J.W., and Sangster, D.F., 1981. Volcanic-associated massive

sulfide deposits. Economic Geology 75th Anniversary Volume, p. 485-627.

Franklin, J.M., Gibson, H.L., Galley, A.G., Jonasson, I.R., 2005. Volcanogenic massive

sulfide deposits. Economic Geology, 100th Anniversary Volume, p. 523-560.

Galley, A.G., 1993. Characteristics of semi-conformable alteration zones associated with

volcanogenic massive sulphide districts. Journal of Geochemical Exploration,

vol. 48, p. 175-200.

87

Galley, A.G., Hannington, M., Jonasson, I., 2007. Volcanogenic massive sulphide

deposits. Geological Association of Canada, Mineral Deposits Division, Special

Publication num. 5, p. 141-161.

Garcelon, E.A., Lentz, D.R., and Walker, J.A., 2016. Portable X-ray fluorescence

spectroscopy applied to volcanic rocks hosting VMS deposits of the California

Lake Group, Bathurst Mining Camp, New Brunswick. New Brunswick

Department of Energy and Resource Development, Geological Surveys Branch,

Geoscience Report 2016-5 (Online), 25 p.

Gates, W., 1972. Heath Steele mine: 24th International Geological Congress. Field

Excursion A58 guidebook, p. 68-72.

Gibson, H.L., and Galley, A.G., 2007. Volcanogenic massive sulphide deposits of the

Archean, Noranda district, Québec. In A Synthesis of Major deposit-types,

district Metallogeny, the Evolution of Geological Provinces, and Exploration

Methods. Edited by W.D. Goodfellow. Mineral Deposits of Canada-Geological

Association of Canada, Mineral Deposits Division, Special Publication vol.5, p.

533-552.

Gibson, H.L., Morton, R.L., and Hudak, G.J., 1999. Submarine volcanic processes,

deposits, and environments favorable for the location of volcanic-associated

massive sulfide deposits. Reviews in Economic Geology, vol. 8, p. 13-51.

Gibson, H.L., Allen, R.L., Riverin, G., Lane, T.E., 2007. The VMS: advances and

application to exploration target. In Proceedings of Exploration 07: Fifth

88

Decennial International Conference on Mineral Exploration. Edited by B.

Milkereit, p. 713-730.

Glanzman, R.K., and Closs, L.G., 2007. Field portable X-ray fluorescence geochemical

analysis its contribution to onsite real-time project evaluation. In Proceedings of

Exploration 07: Fifth Decennial International Conference on Mineral Exploration

Edited by B. Milkereit, p. 291-301.

Goodfellow, W.D., 2007. Metallogeny of the Bathurst Mining Camp, northern New

Brunswick. In A Synthesis of Major Deposit-Types, District Metallogeny, the

Evolution of Geological Provinces, and Exploration Methods. Edited by W.D.

Goodfellow. Mineral Deposits of Canada-Geological Association of Canada,

Mineral Deposits Division, Special Publication num. 5, p. 449-469.

Goodfellow, W.D. and McCutcheon, S.R., 2003. Geologic and genetic attributes of

volcanic sediment-hosted massive sulfide deposits of the Bathurst Mining Camp,

northern New Brunswick-a synthesis. Economic Geology Monograph, vol. 11, p.

245-301.

Goodfellow, W.D., McCutcheon, S.R., and Peter, J.M., 2003. Massive sulfide deposits

of the Bathurst Mining Camp, New Brunswick, and Northern Maine:

introduction and summary of findings. Economic Geology Monograph, vol. 11,

930 p.

Goodfellow, W.D., and Peter, J.M., 1996. Sulphur isotope composition of the Brunswick

No. 12 massive sulphide deposit, Bathurst Mining Camp, N.B.:Implications for

89

ambient environment, sulphur source and ore genesis. Canadian Journal of Earth

Sciences, vol. 33, p. 231-251.

Hall, G., Buchar, A., and Bonham-Carter, G., 2011. Quality control assessment of

portable XRF analysers: development of standard operating procedures,

performance on variable media and recommended uses. CAMIRO Project 10E01

Phase 1.

Hall, G., Buchar, A., Bonham-Carter, G., 2013. Quality Control Assessment of Portable

XRF Analysers: Development of Standard Operating Procedures, Performance

on Variable Media and Recommended Uses. CAMIRO Project 10E01 Phase 2.

Hamilton, A., 1992. Geology of the Stratmat boundary and Heath Steele N-5 zones,

Bathurst Camp, northern New Brunswick. Exploration and Mining Geology, vol.

1, p. 131-135.

Hamilton, A., Park, A., and Moreton, C., 1993. Geology of Heath Steele Mines, Bathurst

Camp, New Brunswick. In Guidebook to the Metallogeny of the Bathurst Camp.

Edited by S.R. McCutcheon and D.R. Lentz. Bathurst ’93:Third Annual Field

Conference, Geological Society of the Canadian Institute of Mining, Metallurgy,

and Petroleum, p. 50-65.

Harley, D.N., 1979. A mineralized Ordovician resurgent caldera complex in the

Bathurst-Newcastle mining district, New Brunswick, Canada. Economic

Geology, vol. 74, p. 786-796.

90

Hibbard, J.P., van Staal, C.R., and Rankin, D.W., and Williams, H., 2006. Lithotectonic

map of the Appalachian Orogen (North), Canada-United States of America;

Geological Survey of Canada, Map 02042A, scale 1:1,500,000.

Ishikawa, Y., Sawaguchi, T., Ywaya, S., and Horiuchi, M., 1976. Delineation of

prospecting targets for Kuroko deposits based on modes of volcanism of

underlying dacite and alteration haloes. Mining Geology, vol. 26, p. 105-117.

Jambor, J.L., 1979. Mineralogical evaluation of proximal-distal features in New

Brunswick massive-sulphide deposits. Canadian Mineralogist, vol. 17, p. 649-

664.

Jenner, G.A., 1996. Trace element geochemistry of igneous rocks: geochemical

nomenclature and analytical geochemistry. In Trace Element Geochemistry of

Volcanic Rocks: Applications for Massive Sulphide Exploration. Edited by D.A.

Wyman. Geological Association of Canada, Short Course Notes, vol. 12, p. 51-

77.

Large, R.R., Gemmel, J.B., Paulick, H., and Huston, D.L., 2001. The alteration box plot:

a simple approach to understanding the relationship between alteration

mineralogy and lithogeochemistry associated with volcanic-hosted massive

sulfide deposits. Economic Geology, vol. 96, p. 957-971.

Langton, J.P., 1996. Relative age, stratigraphic position, and provenance of new

sedimentary formations in the eastern Bathurst camp, New Brunswick. In New

Brunswick Department of Natural Resources and Energy, Minerals and Energy

Division Mineral Resource Report 96-1, p. 61-71.

91

Langton, J.P., and McCutcheon, S.R., 1996. The EXTECH-II project (1994-1996),

Bathurst Mining Camp, northern New Brunswick. Atlantic Geology, vol. 32, p.

77-98.

Lemiere, B., 2018. A review of applications of pXRF (field portable X-ray fluorescence)

for applied geochemistry. Journal of Geochemical Exploration, vol. 188, p. 350-

363.

Lemiere, B., and Uvarova, Y.A., 2019. New developments in field-portable geochemical

techniques and on-site technologies and their place in mineral exploration.

Geochemistry: Exploration, Environment, Analysis, vol. 20, p. 205-216.

Lentz, D.R., 1994. A gamma-ray spectrometric study of the footwall felsic volcanic and

sedimentary rocks around Brunswick No. 6 massive sulphide deposit, northern

New Brunswick. In Geological Survey of Canada, Current Research 1994-D, p.

135-141.

Lentz, D.R., 1995. Preliminary evaluation of six in-house rock geochemical standards

from the Bathurst Camp, New Brunswick. In New Brunswick Department of

Natural Resources and Energy, Current Research 1994, Miscellaneous Report

18, p. 81-89.

Lentz, D.R., 1996a. Recent advances in lithogeochemical exploration for massive-

sulfide deposits in volcano-sedimentary environments: petrogenetic,

chemostratigraphic, and alteration aspects with examples from the Bathurst

Camp, New Brunswick. New Brunswick Department of Natural Resources and

Energy, Current Research 1995, Mineral Resources Report 96, p. 73-119.

92

Lentz, D.R., 1996b. Trace-element systematics of felsic volcanic rocks associated with

massive-sulfide deposits in the Bathurst Mining Camp: petrogenetic, tectonic,

and chemostratigraphic implications for VMS exploration. In Trace Element

Geochemistry of Volcanic Rocks: Applications for Massive Sulphide

Exploration. Edited by D.A. Wyman. Geological Association of Canada, Short

Course Notes, vol. 12, p. 359-402.

Lentz, D.R., 1996c. Geology and geochemistry of Tetagouche Group volcanic and

sedimentary rocks and related debris flows north of the Brunswick No. 12

massive-sulphide deposit, Bathurst, New Brunswick. In Geological Survey of

Canada, Current Research 1996-D, p. 81-92.

Lentz, D.R., 1997. Re-analysis of a section through the Heath Steele B-B5 zone area,

Bathurst Mining Camp, New Brunswick: exploration implications. In Current

Research 1996. Edited by B.M.W Carroll. New Brunswick Department of

Natural Resources and Energy, Minerals and Energy Division, Mineral Resource

Report 97-4, p. 113-127.

Lentz, D.R., 1998. Petrogenetic evolution of felsic volcanic sequences associated with

Phanerozoic volcanic-hosted massive sulphide systems: the role of extensional

geodynamics. Ore Geology Reviews, vol. 12, p. 289-327.

Lentz, D.R., 1999. Petrology, geochemistry, and oxygen isotope interpretation of felsic

volcanic rocks and related rocks hosting the Brunswick No. 6 and No. 12

massive sulfide deposits, Bathurst Mining Camp, New Brunswick, Canada.

Economic Geology, vol. 94, p. 57-86.

93

Lentz, D.R. and Goodfellow, W.D., 1993. Stringer sulphide mineralization and

alteration around the FAB base-metal occurrence, Bathurst Mining Camp, New

Brunswick. In New Brunswick Department of Natural Resources and Energy,

Current Research 1992, Information Circular 93-1, p. 91-103.

Lentz, D.R. and Goodfellow, W.D., 1994a. Character, distribution and origin of zoned

hydrothermal alteration features at the Brunswick No. 12 massive sulphide

deposit, Bathurst Mining Camp, New Brunswick. New Brunswick Department of

Natural Resources and Energy, Minerals Resources, Information Circular 94-1,

p. 94–119.

Lentz, D.R. and Goodfellow, W.D., 1994b. Petrology and geochemistry of altered

volcanic and sedimentary rocks associated with the FAB stringer sulphide zone,

Bathurst, New Brunswick. In Geological Survey of Canada, Current Research

1994-D, p. 123-133.

Lentz, D.R., Goodfellow, W.D, and Brooks, E., 1996. Chemostratigraphy and

depositional environment of an Ordovician sedimentary section across the

Miramichi Group-Tetagouche Group contact, northeastern New Brunswick.

Atlantic Geology, vol. 32, p. 101–122.

Lentz, D.R., Hall, D.C., and Hoy, L.D., 1997. Chemostratigraphic, alteration, and

oxygen isotopic trends in a profile through the stratigraphic sequence hosting the

Heath Steele B zone massive sulfide deposit, New Brunswick. Canadian

Mineralogist, vol. 35, p. 841-874.

94

Lentz, D.R., and Wilson, R.A., 1997. Chemostratigraphic analysis of the volcanic and

sedimentary rocks in the Heath Steele B-B5 zone area, Bathurst Camp, New

Brunswick: stratigraphic and structural implications In: Geological Survey of

Canada Current Research 1997-D, p. 21–33.

Lesher, C.M., Gibson, H.L., and Campbell, I.H., 1986. Composition-volume changes

during hydrothermal alteration of andesite at Buttercup Hill, Noranda District,

Quebec. Geochimica et Cosmochimica Acta, vol. 50, p. 2693-2705.

Luff, W.M., 1995. A history of mining in the Bathurst area, northern New Brunswick,

Canada. Canadian Institute of Mining and Metallurgy Bulletin, vol. 88, p. 62-68.

Luff, W.M., Goodfellow, W.D., Juras, S.J., 1992. Evidence for a feeder pipe and

associated alteration at the Brunswick no.12 massive sulphide deposit.

Exploration Mining Geology, vol. 1, p 167-185

Lusk, J., 1969. Base metal zoning in the Heath Steele B-1 orebody. New Brunswick,

Canada. Economic Geology, vol. 64, p. 506-518.

Lydon, J.W., 1988a. Volcanogenic massive sulphide deposits Part 1: Descriptive models.

Geoscience Canada, 11, p. 195-202. In Roberts, R.G. and Sheahan, P.A. eds.,

Geoscience Canada Reprint Series 3, p. 145–154.

Lydon, J.W., 1988b. Volcanogenic massive sulphide deposits Part 2: Genetic models.

Geoscience Canada, 11, p. 195-202. In Roberts, R.G. and Sheahan, P.A. eds.,

Geoscience Canada Reprint Series 3, p. 155–181.

95

MacLean, W.H., and Barrett, T.J., 1993. Lithogeochemical techniques using immobile

elements. Journal of Geochemical Exploration, vol. 48, p. 109-133.

McAllister, A.L., 1960. Massive sulphide deposits in New Brunswick. CIM Bulletin,

vol. 53, p. 88-98.

McBride, D.E., 1976. Geology of Heath Steele Mines, New Brunswick. Unpublished

Ph.D thesis. University of New Brunswick, Fredericton, NB, 227 p.

McClenaghan, S.H., Lentz, D.R., and Beaumont-Smith, C.J., 2006. The gold-rich

Louvicourt volcanogenic massive sulfide deposit, New Brunswick; a Kuroko

analogue in the Bathurst mining camp. Exploration and Mining Geology, vol. 15,

p. 127-154.

McCutcheon, S.R., 1992. Base-metal deposits of the Bathurst-Newcastle district:

characteristics and depositional models. Exploration and Mining Geology, vol. 1,

p. 105-119.

McCutcheon, S.R., Langton, J.P., van Staal, C.R., and Lentz, D.R., 1993. Stratigraphy,

tectonic setting and massive sulphide deposits of the Bathurst mining camp,

northern New Brunswick. In Guidebook to the Metallogeny of the Bathurst

Camp. Edited by S.R. McCutcheon and D.R. Lentz. Bathurst ’93:Third Annual

Field Conference, Geological Society of the Canadian Institute of Mining,

Metallurgy, and Petroleum, p. 1-28.

96

McCutcheon, S.R., Luff, W.M., and Boyle, R.W., 2003. The Bathurst Mining Camp,

New Brunswick, Canada: history of discovery and evolution of geologic models.

Economic Geology Monograph, vol. 11, p. 17-35.

McCutcheon, S.R., and Walker, J.A., 2009. Stratigraphic Setting of Base Metal Deposits

in the Bathurst Mining Camp (BMC), New Brunswick, Field Trip Guidebook:

International Applied Geochemistry Symposium (IAGS) 2009, p. 22.

McCutcheon, S.R., and Walker J.A., 2019. Great mining camps of Canada 7. The

Bathurst Mining Camp, New Brunswick, Part 1: Geology and Exploration

History. Geoscience Canada, vol. 46, p. 137-154.

McCutcheon, S.R., Walker, J.A., and McClenaghan, S.H., 2001. The geological settings

of massive sulphide deposits in the Bathurst Mining Camp: a synthesis. In

Current Research 2000. Edited by B.M.W. Carroll. New Brunswick Department

of Natural Resources and Energy, Minerals and Energy Division, Mineral

Resource Report 2001-4, p. 64-95.

McDonald, D.W.A., 1983. Fragmental pyrrhotite ore at Heath Steele Mines, New

Brunswick. M.Sc. thesis, University of New Brunswick, Fredericton, NB, 177 p.

Moreton, C., 1994. The stratigraphy, structure, and geometry of the B, B5 and E zone

massive sulphide deposits, Heath Steele Mines, Newcastle, New Brunswick,

Canada. Unpublished Ph. D. thesis, University of New Brunswick, Fredericton,

NB, 398 p.

97

New Brunswick Department of Natural Resources and Energy Development., 2018.

https://www2.gnb.ca/content/gnb/en/departments/erd/energy/content/minerals.ht

ml.

Nelson, G.E., 1983. Alteration of footwall rocks at Brunswick No. 6 and Austing Brook

deposits, Bathurst, New Brunswick, Canada. Unpublished M.Sc. thesis,

University of New Brunswick, Fredericton, NB, 236 p.

Ohmoto, H., 1996. Formation of volcanogenic massive sulfide deposits: The Kuroko

perspective. Ore Geology Reviews, 10, p. 135–177.

Owsiacki, L., and McAllister, A.L., 1979. Fragmental massive sulphides at the Heath

Steele Mine, New Brunswick. C.I.M vol. 72, no. 811, p. 93-100.

Pearce, J.A., Harris, N.B.W., and Tindle, A.G., 1984. Trace element discrimination

diagrams for the tectonic interpretation of granitic rocks. Journal of Petrology,

vol. 25, p. 956-983.

Peter, J.M., 2003. Ancient iron formations: their genesis and use in the exploration for

stratiform base-metal sulphide deposits, with examples from the Bathurst Mining

Camp. In Geochemistry of Sediments and Sedimentary Rocks: Evolutionary

Considerations to Mineral Deposit-Forming Environments. Edited by D.R.

Lentz. Geological Association of Canada, GeoText 4, p. 145-176.

Peter, J.M. and Goodfellow, W.D., 1996. Mineralogy, bulk and rare earth element

geochemistry of massive sulfide-associated hydrothermal sediments of the

Brunswick Horizon, Bathurst mining camp, New Brunswick. Canadian Journal

of Earth Sciences, vol. 33, p. 252-283.

98

Peter, J.M., and Goodfellow, W.D., 2003. Hydrothermal sedimentary rocks of the Heath

Steele belt, Bathurst mining camp, New Brunswick-Part 3. Application of

mineralogy and mineral and bulk compositions to massive sulfide exploration.

Economic Geology Monograph, vol.11, p. 417-433.

Peter, J.M., Goodfellow, W.D., and Doherty, W., 2003a. Hydrothermal sedimentary

rocks of the Heath Steele belt, Bathurst mining camp, New Brunswick-Part 2.

Bulk and rare earth element geochemistry and implications for origin. Economic

Geology Monograph, vol.11, p. 391-415.

Peter, J.M., Kjarsgaard, I.M., and Goodfellow, W.D., 2003b. Hydrothermal sedimentary

rocks of the Heath Steele belt, Bathurst mining camp, New Brunswick-Part 1.

Mineralogy and mineral chemistry. Economic Geology Monograph, vol.11, p.

361-390.

Piercey, S.J., 2007. An overview of the use of petrochemistry in the regional exploration

for volcanogenic massive sulfide (VMS) deposits. In Proceedings of Exploration

07. Edited by B. Milkereit. Fifth Decennial International Conference on Mineral

Exploration, Toronto, ON, p. 223-246.

Piercey, S.J., 2010. An overview of petrochemistry in the regional exploration for

volcanogenic massive sulphide (VMS) deposits. Geochemistry: Exploration,

Environment, Analysis, vol. 10, p. 119-136.

Piercey, S.J., 2011. The setting, style, and role of magmatism in the formation of

volcanogenic massive sulfide deposits. Mineralium Deposita, vol. 46, p. 449-

471.

99

Piercey, S.J., 2014. A review of quality assurance and quality control (QA/QC)

procedures for lithogeochemical data. Geoscience Canada, vol. 41, p. 75-88.

Piercey, S.J., and Devine, M.C., 2014. Analysis of powdered reference materials and

known samples with a benchtop, field portable X-ray fluorescence (pXRF)

spectrometer: evaluation of performance and potential applications for

exploration lithogeochemistry. Geochemistry: Exploration, Environment,

Analysis, vol. 14, p. 139-148.

Potts, P.J., and West, M., 2008. Portable X-ray Fluorescence Spectrometry: Capabilities

for In Situ Analysis. The Royal Society of Chemistry, Cambridge, 148 p.

Rogers, N., 1994. The geology and geochemistry of the felsic volcanic rocks of the

Acadians Range Complex, Tetagouche Group, New Brunswick. Unpublished

Ph.D. thesis, Keele University, Keele, UK, 420 p.

Rogers, N., 1995. The petrological variations of the Ordovician felsic volcanic rocks of

the Tetagouche Group, New Brunswick. In Geological Survey of Canada,

Current Research 1995-E, p. 279-286.

Rogers, N., and van Staal, C.R., 2003. Volcanology and tectonic setting of the northern

Bathurst Mining Camp: Part 2. Mafic volcanic constraints on back-arc opening.

Economic Geology Monograph, vol. 11, p. 181-201.

Rogers, N., van Staal, C.R., McNicoll, V. and Thériault, R., 2003a. Volcanology and

tectonic setting of the northern Bathurst Mining Camp: Part 1. Extension and

100

rifting of the Popelogan Arc. Economic Geology Monograph, vol. 11, p. 157-

179.

Rogers, N., van Staal, C.R., Winchester, J.A., and Fyffe, L.R., 2003b. Provenance and

chemical stratigraphy of the sedimentary rocks of the Miramichi, Tetagouche,

California Lake, and Fournier groups, northern New Brunswick. Economic

Geology Monograph, vol. 11, p. 111-128.

Rogers, N., Wodicka, N., McNicoll, V., and van Staal, C.R., 1997. U-Pb ages of

Tetagouche Group felsic volcanic rocks, northern New Brunswick. In Geological

Survey of Canada, Current Research 97-F, p. 61-69.

Rollinson, H., 1993. Using Geochemical Data: evaluation, presentation, interpretation.

London, Longman, 352p.

Ross, P.S. and Bedard, J.H., 2009. Magmatic affinity of modern and ancient subalkaline

volcanic rocks determined from trace-element discriminant diagrams. Canadian

Journal of Earth Sciences, vol. 46, p. 823-839.

Ross, P.S., Bourke, A., and Fresia, B., 2014. Improving lithological discrimination in

exploration drill-cores using portable X-ray fluorescence measurements: testing

three Olympus Innov-X analysers on unprepared cores. Geochemistry:

Exploration, Environment, Analysis, vol. 14, p. 171-185.

Sack, P.J., and Lewis, L., 2013. Field-portable x-ray fluorescence spectrometer use in

volcanogenic massive sulphide exploration with examples from the Touleary

occurrence (MINFILE Occurrence 115O 176) in west-central Yukon. In Yukon

101

Geological Survey. Edited by K.E. MacFarlane, M.G. Nordling, and P.J. Sack.

Exploration and Geology, p. 115-131.

Saif, S.I., 1980. Petrographic and geochemical investigation of iron formation and other

iron-rich rocks in Bathurst district, New Brunswick. In Geological Survey of

Canada Paper 80-1A, p. 309-317.

Saif, S.I., 1983. Petrographic and geochemical characteristics of iron-rich rocks and their

significance in exploration for massive sulfide deposits, Bathurst, New

Brunswick, Canada. Journal of Geochemical Exploration, vol. 19, p. 705-721.

Slack, J.F., 2012. Exhalites. In Volcanogenic Massive Sulfide Occurrence Model. Edited

by W.C. Shanks, R. Thurston. U.S. Geological Survey Scientific Investigations

Report 2010-5070-C, chap. 10, p. 155-163.

Sullivan, R.W. and van Staal, C.R., 1990. Age of a metarhyolite from the Tetagouche

Group, Bathurst, New Brunswick, from U-Pb isochron analyses of zircons

enriched in common Pb. Radiogenic Age and Isotopic Studies Report 3,

Geological Association of Canada Paper 89-2, p. 109-117.

Sullivan, R.W. and van Staal. C.R., 1996. Preliminary chronostratigraphy of the

Tetagouche and Fournier groups in northern New Brunswick. In Geological

Survey of Canada, Current Research 1995-F, Radiogenic Age and Isotopic

Studies Report 9, p. 43-56.

Valverde-Vaquero, P., van Staal, C.R., McNicoll, V., and Dunning, G., 2006. Middle

Ordovician magmatism and metamorphism along the Gander margin in central

102

Newfoundland. Journal of the Geological Society of London, vol. 163, p. 347-

362. van Staal, C.R., 1994. The Brunswick subduction complex in the Canadian

Appalachians: Record of the Late Ordovician to Late Silurian collision between

Laurentia and the Gander margin of Avalon. Tectonics, vol. 13, p. 946-962. van Staal, C.R., Currie, K.L., Rowbotham, G., Rogers, N., and Goodfellow, W., 2008.

Pressure-temperature paths and exhumation of Late Ordovician-Early Silurian

blueschists and associated metamorphic nappes of the Salinic Brunswick

subduction complex, northern Appalachians. Bulletin of the Geological Society

of America, vol. 120, p. 1455-1477. van Staal, C.R. and de Roo, J.A., 1995. Mid-Paleozoic tectonic evolution of the

Appalachian Central Mobile belt in northern New Brunswick, Canada: collision,

extension collapse and dextral transpression. In Current Perspectives in the

Appalachian-Caledonian Orogen. Edited by J.P. Hibbard, C.R. van Staal, and

P.A. Cawood. Geological Association of Canada, Special Paper 41, p. 367-389.

van Staal, C.R., Dewey, J.F., MacNiocaill, C., and McKerrow, W.S., 1998. The

Cambrian-Silurian tectonic evolution of the northern Appalachians and British

Caledonides: History o f a complex, west and southwest Pacific-type segment of

Iapetus: Geological Society of London Special Publication 143, p. 199-242. van Staal, C.R. and Fyffe, L.R., 1991. Dunnage and Gander zones, New Brunswick,

Canadian Appalachian region. New Brunswick Department of Natural Resources

and Energy, Mineral Resources Branch Geoscience Report 91-2, 39 p.

103

van Staal, C.R., Fyffe, L.R., Langton, J.P., and McCutcheon, S.R., 1992. The Ordovician

Tetagouche Group, Bathurst Camp, northern New Brunswick, Canada: history,

tectonic setting and distribution of massive sulphide deposits. Exploration and

Mining Geology, vol. 1, p. 93-103. van Staal, C.R., and Rogers, N., 2000, The northern half of the Bathurst Mining Camp:

Geological Survey of Canada Open File 3839. van Staal, C. R., and Wilson, R. A., 2014. The Popelogan arc - Tetagouche backarc

basin, the Brunswick subduction complex and the Salinic orogeny in northern

New Brunswick: Fredericton, New Brunswick, Canada. Field Trip Guidebook,

Geological Association of Canada/Mineralogical Association of Canada Joint

Annual Meeting 2014, 41 p. van Staal, C.R., Wilson, R.A., Rogers, N., Fyffe, L.R., Gower, S.J., Langton, J.P.,

McCutcheon, S.R., and Walker, J.A., 2002. Geology, Bathurst Mining Camp and

surrounding areas, New Brunswick., Geological Survey of Canada Open File

4182, 1:100 000 scale map sheet. van Staal, C.R., Wilson, R.A., Rogers, N., Fyffe, L.R., Langton, J.P., McCutcheon, S.R.,

McNicoll, V., and Ravenhurst, C.E., 2003. Geology and tectonic history of the

Bathurst Supergroup, Bathurst Mining Camp, and its relation to coeval rocks in

southwestern New Brunswick and adjacent Maine a synthesis. Economic

Geology Monograph, vol. 11, p. 37-60.

104

van Staal, C.R., Winchester, J.A., and Bedard, J.H., 1991. Geochemical variations in

Middle Ordovician volcanic rocks of the northern Miramichi Highlands and their

tectonic significance. Canadian Journal of Earth Sciences, vol. 28, p. 1031-1049.

Wahl, J.L., 1978. Distribution of major, minor, and trace elements in rocks aroun Heath

Steele and Key Anacon Mines. Unpublished Ph.D. thesis. University of New

Brunswick, Fredericton, NB, 429 p.

Walker, J.A. and McCutcheon, S.R., 2011. A chemostratigraphic assessment of core

from the discovery hole of the Halfmile Lake Deep VMS Zone, Bathurst Mining

Camp, northeastern New Brunswick. In Geological Investigations in New

Brunswick for 2010. Edited by Martin, G.L. New Brunswick Department of

Natural Resources; Lands, Minerals and Petroleum Division, Mineral Resource

Report 2, p. 1–49.

Whalen, J.B., Rogers, N., van Staal, C.R., Longstaffe, F.J., Jenner, G.A., and

Winchester, J.A., 1998. Geochemical and isotopic (Nd, O) data from Ordovician

felsic plutonic and volcanic rocks of the Miramichi Highlands; petrogenetic and

metallogenic implications for the Bathurst Mining Camp. Canadian Journal of

Earth Sciences, vol. 35, p. 237-252.

Whitehead, R.E., 1973. Environment of stratiform sulphide deposition: Variation in

Mn:Fe ratio in host rocks at Heath Steele mine, New Brunswick, Canada.

Mineralium Deposita, vol. 8, p. 148-160.

105

Whitehead, R.E.S. and Goodfellow, W.D., 1978. Geochemistry of volcanic rocks from

the Tetagouche Group, Bathurst, New Brunswick, Canada. Canadian Journal of

Earth Sciences, vol. 15, p. 207-219.

Whitehead, R.E., and Govett, G.J.S., 1974. Exploration rock geochemistry detection of

trace element halos at Heath Steele Mines (New Brunswick, Canada) by

discriminant analysis. Journal of Geochemical Exploration, vol. 3, p. 371-386.

Wills, A.O., 2014. Chemostratigraphy and hydrothermal alteration of the Flat Landing

Brook Formation, Brunswick belt, Bathurst Mining Camp. Unpublished Ph.D.

thesis, University of New Brunswick, Fredericton, NB, 572 p.

Wills, A.O., Lentz, D.R., and Roy, G., 2006. Chemostratigraphy at the Brunswick No. 6

volcanic-sediment-hosted massive sulfide deposit, New Brunswick: resolving

geometry from drill core in deformed felsic volcanic rocks. Exploration and

Mining Geology, vol. 15, p. 221-240.

Wilson, R.A., 1993. Geology of the Heath Steele-South Little River Lake-Island Lake

area, Bathurst Camp, New Brunswick. In Guidebook to the Metallogeny of the

Bathurst Camp. Edited by S.R. McCutcheon and D.R. Lentz. Bathurst ’93:Third

Annual Field Conference, Geological Society of the Canadian Institute of

Mining, Metallurgy, and Petroleum, p. 40-45.

Wilson, R.A., 2003. Geochemistry and petrogenesis of Ordovician arc-related mafic

volcanic rocks in the Popelogan Inlier, northern New Brunswick. Canadian

Journal of Earth Sciences, vol. 40, p. 1171-1189.

106

Wilson, R.A., Fyffe, L.R., McNicoll, V., and Wodicka, N., 1999. Lithogeochemistry,

petrography and geochronology of Ordovician rocks in the Big Bald Mountain

area (NTS 21 O/1), Bathurst Mining Camp, New Brunswick. In New Brunswick

Department of Natural Resources and Energy, Minerals and Energy Division,

Current Research 1998, Mineral Resources Report 99-4, p. 89-142.

Wilson, R.A. and Kamo, S.L., 2008. New U–Pb ages from the Chaleurs and Dalhousie

groups: Implications for regional correlations and tectonic evolution of northern

New Brunswick. In Geological Investigations in New Brunswick for 2007.

Edited by G.L. Martin. New Brunswick Department of Natural Resources;

Minerals, Policy and Planning Division, Mineral Resource Report 2008-1, p. 55–

77.

Wilson, R.A., van Staal, C.R., and McClelland, W.C., 2015. Synaccretionary

sedimentary and volcanic rocks in the Ordovician Tetagouche backarc basin,

New Brunswick, Canada: evidence for a transition from foredeep to forearc basin

sedimentation. American Journal of Science, vol. 315, p. 958-1001.

Wilson, R.A., van Staal, C.R., McCutcheon, S.R., and Fyffe, L.R. 1998. Revised

stratigraphic nomenclature for the Bathurst Mining Camp, northern New

Brunswick. In Abstracts 23rd annual Review of Activities. Edited by B.M.W.

Carroll. New Brunswick Department of Natural Resources and Energy, Minerals

and Energy Division, Information Circular 98-3, pp. 59-62.

107

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

series and their differentiation products using immobile elements. Chemical

Geology, vol. 20, p. 325-347.

Yang, K., Scott, S. D., and Goodfellow, W. D., 2003. Footwall alteration associated with

massive sulfide deposits in the Bathurst Mining Camp, New Brunswick:

implication for seafloor hydrothermal mixing processes. Economic Geology

Monograph, vol. 11, p. 435-456.

108

Appendix A Drill logs sheet

109

110

111

112

113

Appendix B Portable XRF correction curves for selected elements used in this project. CRM= certificate reference material values, pXRF=analytical results from portable XRF

114

115

Appendix C

Major-element (wt.%) and trace-element (ppm) pXRF data from the Heath Steele E zone

ID Depth from Depth to Rock SiO2 TiO2 AL2O3 Fe2O3 MNO CaO K2O P2O5 SO3 V Cr Cu Zn As S-362 5.5 10.1 VOL 71.1 0.35 2.5 1.9 0.04 0.06 4.5 1.15 1.1 94 26 30 268 6 S-362 10.1 20.3 SED 60.7 1.38 6.5 6.3 0.10 0.14 9.0 0.07 2.3 244 101 50 301 31 S-362 20.3 22.9 VOL 75.4 0.40 11.9 1.5 0.04 0.06 5.8 0.00 1.0 88 23 21 189 9 S-362 22.9 30.5 SED 73.4 1.06 13.1 4.7 0.07 0.10 8.3 0.00 2.2 163 95 40 296 43 S-362 30.5 57.5 VOL 77.4 0.31 9.0 1.8 0.04 0.05 5.5 0.00 0.9 80 21 23 178 8 S-362 57.5 68.2 SED 74.2 1.50 14.8 7.1 0.07 0.09 8.7 0.00 1.9 197 124 55 253 47 S-362 68.2 121.5 VOL 83.5 0.35 10.6 1.8 0.04 0.05 4.9 0.00 0.8 91 27 23 168 14

116 S-362 121.5 123.1 SED 75.0 1.09 10.6 5.3 0.03 0.05 5.7 0.00 1.7 136 82 79 307 44 S-362 123.1 125.6 VOL 88.8 0.38 10.3 2.5 0.05 0.07 4.3 0.00 1.0 67 27 28 248 15 S-362 125.6 146.6 SED 69.3 1.14 12.3 6.7 0.07 0.09 6.3 0.03 3.2 159 87 59 283 39 S-362 168.4 170.1 SED 74.2 1.56 13.0 9.8 0.24 0.33 8.2 0.00 6.0 1236 177 37 363 50 S-362 170.1 170.1 VOL 80.6 0.98 11.5 8.9 0.28 0.39 6.2 0.00 1.2 972 112 10 191 50 S-362 170.1 173.7 IF 50.1 0.35 1.3 101.5 1.05 1.47 1.5 2.92 12.3 244 149 13 2141 0 S-362 173.7 178.3 VOL 68.0 0.62 8.4 40.3 0.54 0.75 4.0 0.91 13.7 240 108 58 264 65 S-362 181.3 194.5 VOL 82.8 0.36 9.2 9.9 0.15 0.21 4.6 0.86 3.3 127 48 35 187 37 S-806 3.7 32.6 VOL 84.8 0.28 9.8 1.5 0.03 0.05 5.5 0.00 2.4 80 17 16 327 8 S-806 32.6 35.8 SED 73.9 0.60 11.3 4.6 0.07 0.10 6.2 0.00 0.6 132 47 23 259 30 S-806 35.8 54.2 VOL 84.9 0.29 9.2 1.5 0.03 0.05 3.9 0.00 3.0 75 15 7 239 7 S-806 54.2 74.1 SED 76.2 1.27 11.5 5.9 0.07 0.10 7.4 0.00 4.4 181 94 24 649 45 S-806 120.7 124.4 SED 84.0 0.42 7.9 7.1 0.07 0.10 5.0 0.00 10.2 216 45 24 145 27 S-806 124.4 136.2 VOL 82.2 0.49 10.4 1.5 0.05 0.07 6.4 0.00 2.0 302 40 16 378 8 S-832 5.5 28.2 VOL 84.4 0.32 10.1 1.5 0.04 0.05 6.5 0.00 1.9 93 20 15 227 8 S-832 28.2 28.4 SED 94.0 0.72 8.4 1.6 0.02 0.03 4.8 0.00 0.5 106 34 18 172 9 S-832 28.4 29.8 VOL 82.5 0.28 9.6 3.5 0.05 0.07 2.8 0.00 0.6 64 28 12 35 13

ID Depth from Depth to Rock Rb Sr Y Zr Nb Mo Ag Ba La Ce Pb Th U S-362 5.5 10.1 VOL 99 31 14 90 8 3 0 0 0 7 6 17 16 S-362 10.1 20.3 SED 171 15 34 171 18 12 0 63 0 4 0 34 1 S-362 20.3 22.9 VOL 184 42 41 149 14 1 0 510 0 33 1 14 3 S-362 22.9 30.5 SED 248 26 27 147 19 1 0 511 8 82 0 9 0 S-362 30.5 57.5 VOL 165 52 38 113 11 1 0 587 5 52 2 11 1 S-362 57.5 68.2 SED 225 20 33 292 22 1 0 479 21 99 0 11 0 S-362 68.2 121.5 VOL 141 69 37 123 12 2 0 602 6 53 4 12 1 S-362 121.5 123.1 SED 181 12 30 234 16 2 0 371 0 62 0 9 0 S-362 123.1 125.6 VOL 152 37 31 145 11 2 0 285 12 44 12 9 1 S-362 125.6 146.6 SED 183 20 31 210 18 1 0 496 12 73 0 9 0 S-362 168.4 170.1 SED 159 32 32 117 15 4 158 9213 0 36 24 10 0 117 S-362 170.1 170.1 VOL 112 91 38 138 12 3 149 8864 0 20 0 13 1 S-362 170.1 173.7 IF 52 180 55 30 15 10 30 824 0 0 5616 0 0 S-362 173.7 178.3 VOL 95 52 39 98 13 3 18 1050 1 34 366 8 1 S-362 181.3 194.5 VOL 114 43 41 89 10 2 3 696 3 30 85 7 1 S-806 3.7 32.6 VOL 139 51 31 107 10 1 2 683 0 40 7 10 1 S-806 32.6 35.8 SED 194 51 42 225 16 1 7 656 0 0 5 14 2 S-806 35.8 54.2 VOL 127 48 31 114 9 1 0 626 0 45 5 10 1 S-806 54.2 74.1 SED 219 18 30 219 20 0 1 657 25 97 0 9 0 S-806 120.7 124.4 SED 120 15 42 114 10 3 23 1787 0 0 0 9 0 S-806 124.4 136.2 VOL 161 41 37 127 12 3 39 2958 0 44 24 12 0 S-832 5.5 28.2 VOL 159 55 34 106 11 2 2 730 7 61 13 10 1 S-832 28.2 28.4 SED 151 9 30 234 13 3 0 543 0 63 0 7 0 S-832 28.4 29.8 VOL 96 31 33 96 8 3 0 269 0 74 5 10 0

ID Depth from Depth to Rock SiO2 TiO2 AL2O3 Fe2O3 MNO CaO K2O P2O5 SO3 V Cr Cu Zn As S-832 29.8 30.0 SED 90.9 0.69 9.9 2.8 0.02 0.03 5.1 0.00 2.4 133 43 21 323 25 S-832 30.0 30.3 VOL 73.6 0.39 12.4 3.7 0.03 0.04 7.6 0.00 0.6 136 31 16 310 21 S-832 30.3 30.5 SED 74.8 1.05 12.8 3.6 0.03 0.04 7.7 0.00 0.5 201 62 54 174 29 S-832 30.5 32.2 VOL 77.3 0.75 11.4 5.0 0.03 0.04 6.3 0.00 1.2 131 57 28 182 32 S-832 32.2 33.9 SED 87.0 0.90 8.6 4.0 0.03 0.04 4.5 0.00 1.1 113 60 50 285 23 S-832 33.9 38.4 VOL 88.0 0.25 10.2 1.2 0.03 0.04 3.5 0.00 1.3 49 12 13 301 6 S-832 38.4 47.2 SED 78.9 2.13 11.1 6.4 0.12 0.17 6.1 0.00 9.1 194 68 45 224 25 S-832 47.2 47.8 VOL 74.4 0.47 15.7 2.9 0.06 0.08 8.8 0.00 2.0 100 28 18 297 14 S-832 47.8 88.3 SED 75.4 2.02 11.1 8.5 0.11 0.16 4.1 0.00 4.0 239 74 27 305 66 S-832 99.2 103.0 VOL 89.3 0.28 10.5 1.5 0.02 0.03 4.6 0.00 1.6 87 20 14 430 10 S-832 103.0 111.3 SED 79.1 0.90 11.2 11.9 0.10 0.14 6.0 0.00 2.3 192 68 18 259 45

118 S-832 131.0 139.3 VOL 83.3 0.31 12.2 2.2 0.07 0.09 4.9 0.00 3.1 97 20 9 264 10 S-908 90.2 93.9 VOL 89.1 0.26 7.0 2.4 0.04 0.05 3.9 0.00 0.0 61 27 0 20 7

S-908 93.9 96.8 SED 85.0 0.86 7.9 4.6 0.06 0.08 5.7 0.00 0.2 123 80 9 29 32 S-908 96.8 101.7 VOL 83.4 0.26 10.7 1.7 0.02 0.03 4.2 0.00 0.0 52 21 0 15 8 S-908 101.7 117.4 SED 73.7 1.09 13.4 6.4 0.06 0.09 8.6 0.00 0.4 156 101 3 53 35 S-908 117.4 164.1 VOL 81.6 0.26 10.0 1.6 0.03 0.04 5.5 0.00 0.0 79 26 0 14 7 S-908 164.1 184.6 SED 75.5 1.44 12.2 6.7 0.07 0.10 6.5 0.00 2.7 342 133 39 61 40 S-908 187.0 208.0 VOL 68.8 0.51 8.2 25.8 0.53 0.73 4.9 0.27 13.4 191 104 64 91 38 S-908 208.0 209.5 IF 38.9 0.66 1.9 76.4 1.07 1.50 1.7 10.44 2.7 261 253 0 134 70 S-908 209.5 214.1 VOL 82.4 0.44 10.6 4.1 0.09 0.12 8.0 0.00 0.0 165 45 0 24 15 S-908 214.1 214.6 IF 46.7 0.78 2.4 112.0 1.11 1.56 1.3 0.00 1.8 280 273 0 138 0 S-908 214.6 231.3 VOL 67.8 0.46 10.2 26.6 0.30 0.43 4.9 0.00 0.7 95 117 8 171 46 S-908 231.3 232.9 SED 45.2 3.03 12.5 24.2 0.31 0.43 2.4 0.00 0.8 390 257 44 171 63 S-908 232.9 237.4 VOL 67.5 0.36 9.2 28.1 0.32 0.45 2.0 0.00 2.9 56 104 6 197 59

ID Depth from Depth to Rock Rb Sr Y Zr Nb Mo Ag Ba La Ce Pb Th U S-832 29.8 30.0 SED 164 11 21 160 12 3 0 788 0 72 0 6 0 S-832 30.0 30.3 VOL 231 16 40 170 14 3 10 832 0 89 0 16 0 S-832 30.3 30.5 SED 240 20 39 331 19 0 11 944 0 0 18 12 0 S-832 30.5 32.2 VOL 203 16 34 273 16 0 3 592 0 37 12 14 6 S-832 32.2 33.9 SED 143 11 30 218 15 1 0 348 25 78 0 8 0 S-832 33.9 38.4 VOL 135 30 26 85 9 1 0 327 0 44 12 7 1 S-832 38.4 47.2 SED 209 29 32 232 21 2 0 369 10 63 0 8 1 S-832 47.2 47.8 VOL 317 40 60 177 18 4 0 537 46 88 0 17 3 S-832 47.8 88.3 SED 142 55 36 172 21 3 8 751 12 66 8 6 1 S-832 99.2 103.0 VOL 134 46 30 99 9 3 0 637 0 57 2 8 2 S-832 103.0 111.3 SED 150 15 37 183 15 4 5 890 13 91 3 8 0 119 S-832 131.0 139.3 VOL 153 36 27 104 10 2 3 761 6 70 0 10 1 S-908 90.2 93.9 VOL 115 39 24 77 7 0 0 514 0 22 4 8 0 S-908 93.9 96.8 SED 174 8 23 202 17 2 0 441 23 42 0 4 0 S-908 96.8 101.7 VOL 147 34 31 110 9 0 0 343 0 57 0 10 0 S-908 101.7 117.4 SED 225 18 27 212 16 1 0 489 22 97 0 10 0 S-908 117.4 164.1 VOL 166 54 32 102 10 1 1 621 0 43 4 10 1 S-908 164.1 184.6 SED 186 23 33 290 20 0 27 2162 0 106 0 11 0 S-908 187.0 208.0 VOL 123 31 37 112 12 5 13 1064 0 46 108 9 0 S-908 208.0 209.5 IF 44 435 103 40 8 3 13 626 86 113 0 0 0 S-908 209.5 214.1 VOL 163 28 36 158 14 1 10 1310 0 50 0 14 1 S-908 214.1 214.6 IF 41 163 42 53 14 6 24 513 0 0 136 0 0 S-908 214.6 231.3 VOL 105 8 53 146 15 1 1 394 8 70 0 11 1 S-908 231.3 232.9 SED 40 101 36 98 10 1 0 212 0 0 0 0 0 S-908 232.9 237.4 VOL 33 3 34 95 10 2 0 109 0 34 0 7 0

ID Depth from Depth to Rock SiO2 TiO2 AL2O3 Fe2O3 MNO CaO K2O P2O5 SO3 V Cr Cu Zn As S-908 237.4 238.1 IF 73.7 0.32 8.6 214.2 0.35 0.49 3.4 0.00 6.1 233 285 0 238 0 S-908 238.1 248.8 VOL 72.7 0.40 10.5 17.3 0.24 0.34 6.9 0.00 0.6 106 74 8 119 30 S-908 248.8 249.0 SED 66.6 1.43 14.8 6.7 0.14 0.20 11.0 0.00 2.3 261 139 16 80 28 S-908 249.0 249.3 VOL 77.0 0.37 11.2 3.6 0.06 0.08 7.7 0.00 1.0 118 40 0 33 16 S-908 249.4 249.9 SED 69.6 1.33 13.8 6.8 0.13 0.19 9.8 0.00 3.3 222 129 25 51 26 S-908 250.7 253.4 VOL 43.2 2.09 9.5 15.1 0.27 0.38 7.3 0.00 2.1 263 186 28 63 57 S-908 253.4 254.8 SED 71.5 1.38 16.6 6.3 0.19 0.26 11.7 0.00 7.7 217 137 29 70 33 S-908 254.8 263.0 VOL 90.2 0.32 10.0 1.3 0.03 0.04 4.6 0.00 2.8 58 22 0 10 16 S-908 263.0 273.0 SED 77.1 1.04 13.5 4.5 0.11 0.16 8.0 0.00 3.9 152 92 8 37 25 S-908 273.0 274.4 VOL 88.2 0.27 8.6 2.1 0.02 0.03 5.2 0.00 1.4 125 39 0 13 9 S-908 274.4 276.8 SED 66.9 1.37 16.6 11.3 0.15 0.21 10.7 0.00 1.7 224 193 5 92 38

120 S-908 276.8 288.3 VOL 89.4 0.25 9.2 1.5 0.03 0.04 4.8 0.00 0.4 57 19 0 7 7 S-912 5.5 49.6 VOL 82.0 0.29 8.0 1.5 0.03 0.04 4.9 0.00 1.1 65 16 1 20 19

S-912 49.6 49.8 SED 72.8 1.46 12.6 9.5 0.14 0.19 10.3 0.00 4.3 206 121 40 103 54 S-912 49.8 50.0 VOL 80.7 0.22 8.1 1.1 0.02 0.02 3.6 0.00 3.8 51 12 0 10 6 S-912 50.0 50.3 SED 72.3 1.27 10.8 5.4 0.11 0.15 7.8 0.00 4.1 164 94 23 98 45 S-912 50.3 50.7 VOL 84.1 0.18 7.3 0.9 0.02 0.02 2.9 0.00 0.0 66 8 0 8 5 S-912 50.7 51.3 SED 75.2 1.26 15.0 5.5 0.08 0.11 8.4 0.00 3.5 203 113 15 45 53 S-912 51.3 51.5 VOL 77.0 0.19 7.6 1.8 0.02 0.03 3.1 0.00 1.8 43 12 10 19 8 S-912 51.5 57.7 SED 74.0 1.24 11.4 6.4 0.06 0.08 7.5 0.00 2.2 179 113 29 62 39 S-912 57.7 58.4 VOL 76.0 0.34 9.0 1.6 0.03 0.04 5.0 0.00 0.3 73 23 0 19 6 S-912 58.4 59.4 SED 79.7 1.34 11.4 4.6 0.05 0.07 7.8 0.00 0.7 161 100 39 63 28 S-912 59.4 118.0 VOL 77.9 0.28 7.7 1.8 0.05 0.06 4.4 0.00 2.6 66 12 2 19 8 S-912 118.0 136.6 SED 81.5 1.21 9.9 4.8 0.04 0.06 4.8 0.00 2.7 403 92 19 55 31 S-912 142.0 149.2 VOL 70.6 0.59 7.1 5.6 0.21 0.29 4.4 0.00 1.4 503 60 11 41 19

ID Depth from Depth to Rock Rb Sr Y Zr Nb Mo Ag Ba La Ce Pb Th U S-908 237.4 238.1 IF 150 97 11 25 20 21 30 395 0 0 1183 0 0 S-908 238.1 248.8 VOL 178 9 31 84 11 2 0 633 14 21 3 9 0 S-908 248.8 249.0 SED 247 19 41 249 25 3 0 936 0 94 0 15 0 S-908 249.0 249.3 VOL 195 13 31 169 15 3 0 717 0 0 0 16 0 S-908 249.4 249.9 SED 229 23 32 274 25 3 0 767 0 88 0 13 5 S-908 250.7 253.4 VOL 173 154 32 118 11 3 0 279 0 0 0 0 0 S-908 253.4 254.8 SED 352 21 38 216 24 1 0 874 0 60 0 10 0 S-908 254.8 263.0 VOL 145 28 24 103 10 2 0 388 0 33 0 8 1 S-908 263.0 273.0 SED 234 40 30 145 17 1 0 581 0 51 0 9 0 S-908 273.0 274.4 VOL 140 31 20 57 8 2 5 949 0 0 0 7 0 S-908 274.4 276.8 SED 264 13 40 220 21 1 4 824 0 62 0 10 0 121 S-908 276.8 288.3 VOL 158 39 27 93 9 1 0 481 0 0 0 7 0 S-912 5.5 49.6 VOL 156 37 31 99 9 1 0 490 4 44 0 9 0 S-912 49.6 49.8 SED 304 13 27 207 24 0 0 621 0 65 0 10 0 S-912 49.8 50.0 VOL 121 33 34 87 11 2 0 327 43 64 0 9 0 S-912 50.0 50.3 SED 267 23 29 276 23 4 0 590 0 83 0 10 0 S-912 50.3 50.7 VOL 100 51 27 73 11 0 0 443 0 0 0 8 0 S-912 50.7 51.3 SED 238 21 33 189 21 0 0 628 0 73 0 10 0 S-912 51.3 51.5 VOL 106 36 37 89 8 0 0 253 0 0 33 11 0 S-912 51.5 57.7 SED 209 16 29 243 22 1 0 451 0 72 0 8 0 S-912 57.7 58.4 VOL 163 35 41 133 13 3 0 472 0 99 5 14 0 S-912 58.4 59.4 SED 230 12 28 199 21 0 0 432 0 0 0 8 0 S-912 59.4 118.0 VOL 132 61 30 97 9 1 2 543 5 30 1 10 0 S-912 118.0 136.6 SED 142 37 27 179 18 2 41 2895 13 72 4 9 1 S-912 142.0 149.2 VOL 99 50 30 118 11 3 66 4265 0 11 0 11 0

ID Depth from Depth to Rock SiO2 TiO2 AL2O3 Fe2O3 MNO CaO K2O P2O5 SO3 V Cr Cu Zn As S-912 149.2 150.6 SED 65.4 0.22 2.6 15.5 0.40 0.56 1.0 0.28 3.2 68 44 27 44 40 S-912 150.6 155.5 VOL 74.7 0.39 8.3 3.5 0.10 0.14 6.2 0.00 0.4 215 25 7 25 14 S-912 155.5 156.5 IF 24.5 0.24 0.0 200.1 1.18 1.65 2.1 0.24 3.8 293 197 0 57 0 S-912 156.5 157.7 VOL 57.1 0.46 7.7 35.4 0.54 0.75 2.1 0.00 10.1 139 76 56 108 31 S-912 157.7 158.6 IF 52.9 0.29 1.6 120.8 0.52 0.73 2.3 0.31 1.8 237 135 0 43 0 S-912 158.6 163.3 VOL 69.9 0.37 7.8 15.4 0.31 0.43 3.4 0.00 0.6 112 45 12 69 30 S-912 163.3 164.7 1F 44.6 0.51 2.9 105.3 0.72 1.00 4.9 0.37 1.5 226 189 0 104 26 S-912 164.7 168.5 VOL 62.2 0.35 7.3 17.6 0.36 0.50 4.5 0.00 2.0 105 40 3 85 31 S-912 168.5 169.9 SED 20.6 0.36 2.0 95.6 1.98 2.77 1.6 0.31 16.1 164 163 329 72 94 S-912 169.9 184.9 VOL 72.3 0.34 8.0 11.4 0.23 0.32 5.3 0.00 3.9 97 38 13 67 28 S-912 184.9 187.3 IF 41.2 0.45 4.0 107.8 0.59 0.83 2.9 0.21 6.9 198 140 35 139 20

122 S-912 187.3 192.5 VOL 67.5 0.32 7.2 29.3 0.44 0.61 1.0 0.05 2.8 66 54 62 285 45 S-912 192.5 193.9 SED 51.1 2.88 10.6 17.1 0.34 0.47 6.1 0.00 13.6 368 181 481 102 75

S-912 193.9 208.6 VOL 76.6 0.42 7.7 5.3 0.11 0.15 5.7 0.00 6.9 110 42 14 47 21 S-912 208.6 213.9 SED 76.8 1.19 12.6 6.5 0.17 0.24 8.2 0.00 4.8 210 92 13 39 41 S-912 213.9 236.5 VOL 79.4 0.23 7.4 2.0 0.04 0.05 4.7 0.00 3.2 73 10 2 12 9 S-913 4.3 61.0 VOL 86.5 0.28 9.1 1.7 0.03 0.05 4.7 0.00 1.5 59 14 7 25 8 S-913 61.0 72.2 SED 76.6 1.14 11.4 6.4 0.08 0.11 7.1 0.00 0.7 149 84 19 65 39 S-913 72.2 78.9 VOL 81.4 0.29 8.9 1.8 0.04 0.06 3.6 0.00 0.4 59 17 9 26 9 S-913 78.9 80.1 SED 84.6 1.02 9.8 4.3 0.07 0.09 6.6 0.00 0.5 155 73 12 37 41 S-913 80.1 126.5 VOL 80.5 0.38 9.7 2.3 0.05 0.07 5.5 0.00 0.6 88 21 6 23 10 S-913 126.5 149.3 SED 79.8 1.53 12.2 5.8 0.06 0.09 6.2 0.00 1.2 590 118 28 56 36 S-913 160.6 164.1 VOL 37.1 1.45 15.6 40.3 1.65 2.30 3.9 0.02 1.4 1369 180 13 332 80 S-913 164.6 168.4 VOL 79.5 0.60 8.8 9.6 0.34 0.48 4.0 0.00 1.1 414 54 0 59 24 S-913 168.4 170.6 VOL 52.8 0.49 8.4 44.7 0.89 1.25 1.1 0.00 2.4 91 76 8 286 61

ID Depth from Depth to Rock Rb Sr Y Zr Nb Mo Ag Ba La Ce Pb Th U S-912 149.2 150.6 SED 17 39 39 41 6 2 0 221 0 21 7 3 0 S-912 150.6 155.5 VOL 135 45 36 100 11 1 23 2001 0 27 0 10 1 S-912 155.5 156.5 IF 76 148 16 12 12 8 38 965 0 0 1110 0 0 S-912 156.5 157.7 VOL 44 33 39 99 11 3 9 577 0 0 0 11 1 S-912 157.7 158.6 IF 69 253 49 30 9 6 30 530 0 0 845 0 0 S-912 158.6 163.3 VOL 77 29 37 99 10 1 3 678 11 45 1 9 1 S-912 163.3 164.7 1F 166 176 44 44 12 6 25 557 0 0 192 0 0 S-912 164.7 168.5 VOL 121 12 35 110 12 4 8 819 0 25 0 8 0 S-912 168.5 169.9 SED 27 170 54 28 8 2 13 95 38 0 275 0 0 S-912 169.9 184.9 VOL 137 20 33 98 11 1 1 632 8 34 0 10 1 S-912 184.9 187.3 IF 100 117 43 57 12 5 15 309 0 24 369 2 0 123 S-912 187.3 192.5 VOL 21 30 23 63 8 1 0 106 0 0 3 5 0 S-912 192.5 193.9 SED 135 137 38 90 10 1 5 555 0 0 0 0 0 S-912 193.9 208.6 VOL 157 27 43 150 13 3 2 681 0 57 0 13 0 S-912 208.6 213.9 SED 251 20 29 173 20 0 7 1076 0 47 0 9 0 S-912 213.9 236.5 VOL 142 24 35 78 9 2 0 685 0 21 0 7 0 S-913 4.3 61.0 VOL 145 42 25 93 9 0 0 403 0 18 0 8 0 S-913 61.0 72.2 SED 183 14 38 298 20 1 0 416 38 81 0 10 2 S-913 72.2 78.9 VOL 119 58 39 145 10 1 0 440 0 70 0 12 0 S-913 78.9 80.1 SED 183 11 21 184 17 2 0 546 0 78 0 8 0 S-913 80.1 126.5 VOL 155 71 39 125 13 1 2 679 5 36 1 12 1 S-913 126.5 149.3 SED 167 44 31 227 19 2 64 4052 7 67 0 10 0 S-913 160.6 164.1 VOL 27 84 57 175 20 0 184 10826 0 84 0 15 0 S-913 164.6 168.4 VOL 86 52 37 128 11 2 62 4025 0 61 23 12 0 S-913 168.4 170.6 VOL 14 5 31 105 14 1 0 136 0 0 0 9 0

ID Depth from Depth to Rock SiO2 TiO2 AL2O3 Fe2O3 MNO CaO K2O P2O5 SO3 V Cr Cu Zn As S-913 170.6 171.0 IF 49.8 0.89 9.8 58.4 0.89 1.24 3.9 0.00 18.5 258 128 0 600 79 S-913 171.0 173.0 SED 74.3 1.14 9.1 19.4 0.41 0.58 7.0 0.00 1.4 382 192 20 101 56 S-913 173.0 193.1 VOL 67.6 0.42 9.4 13.5 0.31 0.43 6.2 0.03 1.6 144 54 10 112 26 S-913 193.1 195.0 IF 23.3 0.43 1.7 210.2 0.38 0.53 3.0 0.32 6.4 274 193 0 320 0 S-913 195.0 203.4 VOL 77.2 0.32 8.3 15.5 0.33 0.46 4.5 0.00 1.4 92 48 15 89 43 S-913 203.4 206.6 SED 54.9 3.23 11.1 17.2 0.30 0.42 4.1 0.00 2.7 380 193 67 89 62 S-913 206.6 212.3 VOL 84.3 0.30 6.4 8.4 0.25 0.35 4.0 0.02 5.5 100 34 12 172 18 S-913 212.3 237.8 SED 69.6 1.66 15.2 8.3 0.11 0.15 9.7 0.00 4.2 372 136 21 65 186 S-913 237.8 242.6 VOL 70.3 0.34 11.0 3.5 0.03 0.04 6.5 0.00 13.2 80 20 30 38 13 S-913 242.6 245.4 SED 81.0 1.05 10.7 4.5 0.07 0.09 6.6 0.00 3.4 127 71 11 40 22 S-916 5.5 28.1 VOL 62.5 0.06 12.8 2.6 0.00 0.00 3.9 0.01 0.0 0 0 0 43 4

124 S-916 28.1 34.7 SED 58.6 0.26 14.4 3.7 0.00 0.00 3.8 0.01 0.0 0 0 0 48 9 S-916 34.7 85.5 VOL 65.6 0.00 11.5 1.7 0.00 0.08 2.7 0.01 0.0 0 0 0 19 1

S-916 85.5 96.9 SED 58.0 0.37 15.2 4.4 0.07 0.04 4.3 0.10 0.2 0 0 0 61 10 S-916 96.9 107.6 VOL 63.1 0.07 12.3 2.6 0.00 0.03 1.7 0.01 0.0 0 0 0 41 2 S-916 107.6 131.8 SED 58.6 0.76 13.1 5.6 0.00 0.19 2.8 0.18 0.0 0 0 0 77 19 S-916 150.6 153.2 VOL 38.9 0.00 5.5 5.3 0.00 0.45 2.8 0.15 0.2 0 0 0 25 4 S-916 156.4 157.0 IF 42.2 0.26 4.4 22.6 1.28 1.74 1.2 1.19 11.2 0 0 0 309 280 S-916 157.0 159.4 VOL 54.1 0.01 9.0 13.2 0.24 0.05 1.6 0.07 0.6 0 0 0 100 3 S-916 159.4 188.2 SED 56.0 0.22 11.6 11.0 0.09 0.70 2.5 0.65 3.7 0 0 274 70 10 S-916 197.0 202.8 SED 62.4 0.00 13.4 5.1 0.02 0.00 3.4 0.05 0.3 0 0 0 73 7 S-916 202.8 209.1 VOL 69.6 0.00 11.1 1.9 0.00 0.15 2.5 0.00 1.1 0 0 0 21 2 HS17-005 4.0 35.0 VOL 80.2 0.29 9.9 1.8 0.04 0.05 5.4 0.00 0.9 63 12 5 16 11 HS17-005 35.0 36.3 SED 52.7 4.09 13.0 20.0 0.35 0.49 6.8 0.00 7.5 464 249 21 128 121 HS17-005 36.3 78.6 VOL 82.5 0.52 9.9 2.7 0.06 0.08 5.1 0.00 2.2 86 33 7 22 12

ID Depth from Depth to Rock Rb Sr Y Zr Nb Mo Ag Ba La Ce Pb Th U S-913 170.6 171.0 IF 112 5 19 54 10 3 15 496 0 0 0 0 0 S-913 171.0 173.0 SED 154 7 100 94 11 5 35 1824 0 0 0 3 3 S-913 173.0 193.1 VOL 162 31 45 117 12 2 9 913 14 27 88 9 1 S-913 193.1 195.0 IF 131 153 26 24 19 9 39 910 0 0 2811 0 0 S-913 195.0 203.4 VOL 94 8 29 74 9 1 2 442 0 13 0 7 1 S-913 203.4 206.6 SED 94 112 40 103 10 1 7 450 0 0 0 0 0 S-913 206.6 212.3 VOL 112 37 24 81 9 1 2 625 0 23 103 7 1 S-913 212.3 237.8 SED 263 31 44 221 23 0 22 1899 5 100 0 11 0 S-913 237.8 242.6 VOL 210 23 31 90 11 2 0 534 0 0 0 7 0 S-913 242.6 245.4 SED 206 11 29 207 16 0 0 433 0 100 0 7 0 S-916 5.5 28.1 VOL 167 42 28 127 10 1 0 879 17 95 6 11 1 125 S-916 28.1 34.7 SED 174 21 34 178 12 2 0 688 30 119 3 11 2

S-916 34.7 85.5 VOL 117 63 32 111 9 1 0 789 5 70 11 11 3 S-916 85.5 96.9 SED 238 23 30 133 13 1 0 598 11 87 8 11 1 S-916 96.9 107.6 VOL 87 59 29 121 9 0 0 515 11 67 5 11 4 S-916 107.6 131.8 SED 155 26 32 230 15 1 0 569 29 108 7 10 1 S-916 150.6 153.2 VOL 42 135 17 65 5 2 2 7825 0 60 12 5 0 S-916 156.4 157.0 IF 54 61 25 43 4 4 0 231 0 37 786 1 0 S-916 157.0 159.4 VOL 72 15 26 91 7 1 1 1643 4 21 6 8 1 S-916 159.4 188.2 SED 139 47 32 132 10 2 1 575 24 71 25 8 2 S-916 197.0 202.8 SED 170 14 24 256 16 4 0 999 0 139 20 8 2 S-916 202.8 209.1 VOL 136 42 20 129 9 1 0 688 6 47 16 8 1 HS17-005 4.0 35.0 VOL 188 46 26 94 10 1 0 440 3 26 0 8 0 HS17-005 35.0 36.3 SED 221 40 76 165 17 3 9 221 0 0 0 0 0 HS17-005 36.3 78.6 VOL 174 50 30 104 11 2 0 374 6 37 2 7 0

ID Depth from Depth to Rock SiO2 TiO2 AL2O3 Fe2O3 MNO CaO K2O P2O5 SO3 V Cr Cu Zn As HS17-005 78.6 83.2 SED 72.2 1.39 14.7 9.4 0.11 0.16 9.4 0.00 0.7 195 130 49 129 58 HS17-005 83.2 113.0 VOL 80.7 0.39 7.6 2.1 0.04 0.05 4.2 0.00 0.1 74 26 4 19 11 HS17-005 113.0 141.9 SED 80.2 1.30 11.5 6.4 0.05 0.07 7.0 0.01 0.9 162 98 29 59 38 HS17-005 141.9 158.3 VOL 81.4 0.29 11.0 2.0 0.03 0.04 5.2 0.00 0.3 83 20 1 29 12 HS17-005 158.3 171.2 SED 86.1 1.07 8.7 5.6 0.04 0.06 5.4 0.00 0.3 129 84 22 45 34 HS17-005 171.2 186.7 VOL 88.6 0.29 9.2 2.1 0.05 0.06 4.4 0.00 0.5 60 17 6 28 9 HS17-005 186.7 208.2 SED 83.1 1.14 10.8 5.8 0.05 0.07 5.6 0.00 0.2 154 88 17 56 38 HS17-005 228.4 248.1 VOL 81.8 0.29 10.6 2.0 0.07 0.10 6.8 0.00 0.5 115 22 5 23 9 HS17-005 248.1 251.4 SED 87.2 0.43 11.5 1.1 0.03 0.04 3.8 0.00 0.0 75 18 9 44 7 HS17-005 276.8 304.9 SED 80.9 0.79 10.7 5.3 0.06 0.08 5.7 0.00 0.8 124 58 9 35 30 HS17-005 304.9 321.0 VOL 84.9 0.36 9.7 1.6 0.03 0.04 5.9 0.00 0.8 90 13 6 13 6

126

ID Depth from Depth to Rock Rb Sr Y Zr Nb Mo Ag Ba La Ce Pb Th U HS17-005 78.6 83.2 SED 256 17 39 165 23 1 0 455 16 88 0 10 0 HS17-005 83.2 113.0 VOL 137 43 31 183 10 2 0 390 4 45 1 9 0 HS17-005 113.0 141.9 SED 186 18 33 227 21 0 0 419 24 91 0 10 0 HS17-005 141.9 158.3 VOL 153 48 35 110 10 1 1 640 0 57 8 11 1 HS17-005 158.3 171.2 SED 145 14 25 200 18 1 0 331 48 79 0 7 0 HS17-005 171.2 186.7 VOL 164 36 28 91 10 1 0 348 0 29 5 8 0 HS17-005 186.7 208.2 SED 172 17 29 195 19 1 0 451 5 89 0 9 0 HS17-005 228.4 248.1 VOL 192 29 26 78 10 1 8 1034 0 41 3 7 0 HS17-005 248.1 251.4 SED 99 113 47 198 18 0 0 561 0 0 9 21 4 HS17-005 276.8 304.9 SED 164 22 32 188 16 2 0 532 21 72 0 8 0 HS17-005 304.9 321.0 VOL 188 29 28 104 12 1 0 721 0 0 0 8 0

Appendix D Drill core profiles showing major- and trace-element and element ratios discriminators used in this study See Appendix C for data.

127

128

129

130

Appendix E

Section A-A’with average signatures per unit with pXRF data performed on samples taken from drill cores of the Heath Steele E zone

131

132

133

134

Appendix F

Drill core profiles showing the geochemical alteration signature for Nepisiguit Falls Formation at the Heath Steele E zone

135

136

137

138

Curriculum Vitae

Candidate’s full name: Josue Jimenez-Gonzalez

Universities attended: National Polytechnic Institute, Mexico City, Mexico (2018)

B.Sc. in Geological Engineering

University of New Brunswick (2020)

Masters of Science, Earth Science

Conferences:

Jimenez-Gonzalez, J., Lentz, D.R., and Walker, J.A., 2019. Chemostratigraphic correlation between drill cores S362 and S916, Heath Steele E zone, New Brunswick, Canada. In Abstracts 2019: Exploration, Mining and Petroleum New Brunswick. Edited by E.A. Keith. New Brunswick Department of Natural Resources and Energy Development, Geoscience Report 2019-1, p. 11.

Jimenez-Gonzalez, J., Lentz, D.R., Walker J.A., 2019. Summer Fieldwork in Heath Steele E Zone Area, Bathurst Mining Camp, New Brunswick. In Geoscience Project Summaries and Other Activities 2019. Edited by E.A. Keith. New Brunswick Department of Energy and Resource Development, Information Circular 2019-1, p. 49- 53.

Jimenez-Gonzalez, J., Lentz, D.R., Walker J.A., and Day, J.J., 2019. Advances in the Lithogeochemical Study of Drill Core from the Heath Steele E-Zone, Bathurst Mining Camp, New Brunswick. Geological Association of Canada-Mineralogical Association of Canada, Volume of Abstracts, vol. 42, p. 117.

Jimenez-Gonzalez, J., Lentz, D.R., Walker J.A., and Day, J.J., 2019. Chemostratigraphic Assessment of Drill Core S916 from the Heath Steele E Zone, Bathurst Mining Camp, New Brunswick. Atlantic Geoscience Society, vol. 55, p. 182.

Jimenez-Gonzalez, J., Lentz, D.R., Walker J.A., 2018. Lithogeochemical data from drill core S-916, the Heath Steele E Zone, Bathurst Mining Camp, New Brunswick: A preliminary pXRF analysis. In Abstracts 2018: Exploration, Mining and Petroleum New Brunswick. Edited by E.A. Keith. New Brunswick Department of Energy and Resource Development, Geoscience Report 2018-1, p. 14.