The source and evolution of paleofluids responsible for secondary minerals in low-permeability Ordovician limestones of the Michigan Basin

The source and evolution of paleofluids responsible for secondary minerals in low-permeability Ordovician limestones of the Michigan Basin

Accepted Manuscript The source and evolution of paleofluids responsible for secondary minerals in lowpermeability Ordovician limestones of the Michiga...

12MB Sizes 1 Downloads 39 Views

Accepted Manuscript The source and evolution of paleofluids responsible for secondary minerals in lowpermeability Ordovician limestones of the Michigan Basin D.C. Petts, J.K. Saso, L.W. Diamond, L. Aschwanden, T.A. Al, M. Jensen PII:

S0883-2927(17)30282-2

DOI:

10.1016/j.apgeochem.2017.09.011

Reference:

AG 3952

To appear in:

Applied Geochemistry

Received Date: 7 July 2017 Revised Date:

14 September 2017

Accepted Date: 16 September 2017

Please cite this article as: Petts, D.C., Saso, J.K., Diamond, L.W., Aschwanden, L., Al, T.A., Jensen, M., The source and evolution of paleofluids responsible for secondary minerals in lowpermeability Ordovician limestones of the Michigan Basin, Applied Geochemistry (2017), doi: 10.1016/ j.apgeochem.2017.09.011. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Title: The source and evolution of paleofluids responsible for secondary minerals in low-

RI PT

permeability Ordovician limestones of the Michigan Basin

Authors:

a

M AN U

SC

Petts, D.C.a, *, Saso, J.K.b, Diamond, L.W.c, Aschwanden, L.c, Al, T.A.a and Jensen, M.d

Department of Earth and Environmental Sciences, University of Ottawa, Ottawa, ON, K1N

6N5, Canada b

Department of Earth Sciences, University of New Brunswick, Fredericton, New Brunswick,

c

TE D

Canada

Rock-Water Interaction Group, Institute of Geological Sciences, University of Bern,

Baltzerstrasse 3, 3012 Bern, Switzerland

Nuclear Waste Management Organization, Toronto, ON, M4T 2S3, Canada

EP

d

AC C

* Corresponding author: [email protected]

Version: September 19, 2017

1

ACCEPTED MANUSCRIPT

1 2

Abstract In this study we report on the source and evolution of fluids associated with secondary vein and replacement minerals in low-permeability carbonate units in the Michigan Basin. This

4

petrogenetic information was collected using thermometric data from fluid inclusions combined

5

with C-, O- and Sr-isotope data, and focuses on mm- or cm-wide vein and vug minerals from

6

Ordovician limestones of the Trenton and Black River groups and Cambrian sandstones in SW

7

Ontario, Canada.

SC

9

Primary fluid inclusions in dolomite represent fluid stage I and have the highest trapping temperatures (Ttrap) in the sedimentary succession, between 88 and 128 °C. Primary inclusions in

M AN U

8

RI PT

3

calcite (stage II), and celestine and anhydrite (stage III), represent the final stages of secondary

11

mineral formation and have Ttrap values of 54–78 °C. All three stages of vein minerals formed

12

from brines with salinities of 31 to 37 wt.% [CaCl2+NaCl]eq that were saturated in halite and

13

methane gas at the time of mineral growth. Three subsequent stages of secondary fluid

14

inclusions were observed in the samples (stages IV–VI), however, no secondary vein minerals

15

formed during these stages and they are interpreted as re-mobilization and/or minor fluid ingress

16

along grain boundaries and micro-fractures. Notably, stage IV secondary inclusions are gas-

17

undersaturated with salinities of 32 to 34 wt.% [CaCl2+NaCl]eq and minimum Ttrap values of 57–

18

106 °C, and are interpreted to have formed during a Late Devonian–Mississippian regional

19

heating event.

EP

AC C

20

TE D

10

Secondary minerals from the Cambrian units and the Shadow Lake Formation have δ13C

21

values of −6.1 to −2.5‰ (VPDB), δ18O values of +14.6 to +24.2‰(VSMOW) and 87Sr/86Sr of

22

0.70975 to 0.71043. Relative shifts of approximately +2 to +3‰ in δ13C, >+4‰ in δ18O and

23

−0.002 in 87Sr/86Sr are observed upward across the boundary between the Cambrian–Shadow

2

ACCEPTED MANUSCRIPT

Lake units and the overlying Gull River Formation. Samples from the Gull River and Coboconk

25

formations and the Trenton Group (Kirkfield, Sherman Fall and Cobourg formations) have δ13C

26

values of −1.0 to +1.9‰, δ18O values of +18.9 to +28.1‰ and 87Sr/86Sr values of 0.70790 to

27

0.70990.

RI PT

24

The combined microthermometric and isotopic data for secondary minerals in the

29

Cambrian–Shadow Lake units suggest they formed from hydrothermal brine with a geochemical

30

signature obtained by interaction with the underlying Precambrian shield, or shield-derived

31

minerals in the Cambrian sandstones. Previous U-Pb dating of vein calcite and Rb-Sr dating of

32

secondary K-feldspar from the region suggests that brine ingress occurred during the Late

33

Ordovician–Silurian. The O- and Sr-isotope variability in vein samples from the Gull River and

34

Coboconk formations is interpreted as localized mixing of 18O-enriched, connate fluids with

35

hydrothermal brine. In comparison, isotopic data for the Trenton Group indicate secondary

36

mineral formation from connate fluids, sourced from 18O-enriched, evolved seawater and/or

37

modified hydrothermal brine that experienced fluid–rock interaction during transit through the

38

underlying stratigraphy.

M AN U

TE D

EP

39

SC

28

Keywords: Michigan Basin; low-permeability rocks; secondary minerals; fluid–rock interaction;

41

fluid inclusions; strontium isotopes; stable isotopes

AC C

40

3

ACCEPTED MANUSCRIPT

42 43

1. Introduction Low-permeability sedimentary rocks have increasingly become a global focus for a wide range of environmental and geoscience studies, including nuclear waste management (Russell

45

and Gale, 1982; Hendry et al., 2015), CO2 sequestration (Shafeen et al., 2004; Benson and Cole,

46

2008) and extraction of gas from shale (Arthur and Cole, 2014). In the cases of CO2

47

sequestration and extraction of gas with hydraulic fracturing, the low-permeability rocks play a

48

critical role in limiting the upward migration of fluids that are injected or displaced. Low-

49

permeability sedimentary rocks have been of particular interest in countries such as Canada,

50

Switzerland, France, Belgium and Germany for the long-term storage of nuclear waste because

51

solute transport is dominated by diffusion and the clay-rich mineralogy provides large surface

52

areas for solute retention by sorption mechanisms. Internationally, detailed paleofluid studies

53

have provided valuable insight into fluid sources and solute transport mechanisms (Altmann et

54

al., 2012; Clark et al., 2013; Al et al., 2015; Mazurek and de Haller, 2017) that are essential for

55

evaluating the long-term viability of a site for nuclear waste storage.

SC

M AN U

TE D

56

RI PT

44

Low-permeability argillaceous limestones of the Trenton Group (10-16 to 10-12 m s-1; Beauheim et al., 2014), in the Bruce region of the Michigan Basin (southwestern Ontario,

58

Canada) have been identified as a potential host rock for the construction of a deep geologic

59

repository (DGR) for low- and intermediate-level nuclear waste. The Cobourg Formation, a

60

low-permeability, argillaceous limestone at the top of the Trenton Group occurs >650 m beneath

61

the Bruce nuclear site and has been proposed as the repository host rock. These limestones are

62

overlain by up to 200 m of regionally-extensive, low-permeability Ordovician shale (Intera,

63

2011; Beauheim et al., 2014). Detailed studies on the long-term physical and chemical stability

64

of the surrounding rocks have focused on a wide range of stratigraphic, structural, geophysical

AC C

EP

57

4

ACCEPTED MANUSCRIPT

and geochemical methods (Intera, 2011 and references therein). Previous investigations at the

66

Bruce site have contributed valuable information on the residence time and transport

67

mechanisms of fluids in these rocks, which include studies of porewater geochemistry and solute

68

transport (Clark et al., 2013; Al et al., 2015), hydraulic conductivity (Beauheim et al., 2014), and

69

mechanisms responsible for anomalous hydraulic-head profiles observed in the Trenton and

70

Black River group limestones (Normani and Sykes, 2012; Khader and Novakowski, 2014;

71

Neuzil and Provost, 2014; Neuzil, 2015).

SC

72

RI PT

65

Secondary minerals from veins in sedimentary rocks provide direct information on the movement of deep basin fluids. Detailed isotopic and petrographic information of secondary

74

minerals combined with fluid inclusion temperature constraints have been utilized in paleofluid

75

studies to differentiate between fluid sources that relate to either diagenesis/burial (Ayalon and

76

Longstaffe, 1988), or the influx of fluids from allochthonous sources (Spencer, 1980; Davies and

77

Smith, 2006). Detailed petrography (e.g., Taylor and Sibley, 1986) combined with thermometric

78

data from fluid inclusions (e.g., Roedder, 1979; Goldstein 2001) provide an understanding of the

79

relative timing of fluid infiltration and temperature constraints during crystallization. Here, we

80

report on the source and evolution of fluids associated with secondary mineral formation in low-

81

permeability Ordovician limestones and Cambrian siliciclastic units in the Michigan Basin from

82

southwestern Ontario, Canada.

TE D

EP

AC C

83

M AN U

73

Ordovician limestones of the Trenton and Black River groups, which are laterally

84

traceable across much of the Michigan Basin and occur in sections of the Appalachian Basin,

85

have long been a focus of paleofluid studies due to their association with hydrocarbon reservoirs

86

in the Albion-Scipio/Stoney Point fields in south-central Michigan (e.g., Hurley and Budros,

87

1990; Allan and Wiggins, 1993), the Lima-Indiana field in Ohio/Indiana (e.g., Wickstrom et al.,

5

ACCEPTED MANUSCRIPT

1992) and fields of southern Ontario (Powell et al., 1984; Middleton et al., 1993; Coniglio et al.,

89

1994; Obermajer et al., 1999). In these fields, hydrocarbon migration/reservoir formation is

90

closely tied to the formation of hydrothermal dolomite (Allan and Wiggins, 1993), which has

91

been suggested to form during high-temperature, fluid–rock interaction between the Ordovician

92

limestones and upwelling hydrothermal fluids (Davies and Smith, 2006). Furthermore, C-, O-

93

and Sr-isotope compositions of secondary minerals in southern Ontario (Harper et al., 1995;

94

Zeigler and Longstaffe, 2000; Haeri-Ardakani et al., 2013) and central Michigan (Allan and

95

Wiggins, 1993) have provided valuable information on the source of deep-seated fluid

96

migrations in these regions of the basin. In the Bruce region, however, detailed paleofluid

97

studies have largely focused on the geochemical and isotopic evolution of porewaters in the

98

Michigan Basin sedimentary succession (Clark et al., 2010a; 2010b; 2013; 2015; Al et al., 2015).

99

Knowledge gaps remain in the understanding of paleofluid contributions to the formation of

M AN U

SC

RI PT

88

secondary vein minerals in the Ordovician limestones, and in particular, the source and transport

101

mechanisms that are responsible for vein-related fluid migrations in these low-permeability

102

sedimentary rocks.

The objective of this study is to enhance the understanding of fluid residence time and

EP

103

TE D

100

migration that has been developed from previous research at the Bruce site. Detailed

105

petrography and microthermometric data from fluid inclusions are combined with C-, O- and Sr-

106

isotope data to determine the source and relative timing of paleofluid migrations that resulted in

107

the formation of secondary minerals, including vein and vug infill and replacement phases.

108

Samples were obtained from drill cores that section the entire succession of Cambrian to

109

Devonian sedimentary rocks and the underlying Precambrian gneissic basement.

AC C

104

110

6

ACCEPTED MANUSCRIPT

111

2. Geology and Stratigraphy of the Michigan Basin The Michigan Basin represents a broadly circular, ~500 km diameter intracratonic

113

sedimentary basin located in central North America. The study area is located along the eastern

114

margin of the basin, at the Bruce nuclear site (Fig. 1). A detailed summary of the Paleozoic

115

bedrock units in southern Ontario is discussed in Armstrong and Carter (2010) (and references

116

therein), a brief description of these units is provided below. At the Bruce site, the sedimentary

117

succession reaches a maximum thickness of ~ 860 m. The base of this succession comprises ~17

118

m of Cambrian sandstone which unconformably overlies Precambrian gneissic basement. Upper

119

Ordovician rocks disconformably overlie the Cambrian sandstones and are subdivided into the

120

Black River and Trenton group limestones and a thick sequence of shale. The Black River

121

Group consists of a 4–5 m thick basal unit of siltstone and minor dolostone (Shadow Lake

122

Formation) that is overlain by ~ 75 m of variably bioturbated and fossiliferous lithographic

123

limestone (Gull River and Coboconk formations). Overlying the Black River Group are ~ 110 m

124

of mostly argillaceous limestones of the Trenton Group (subdivided into the Kirkfield, Sherman

125

Fall and Cobourg formations). The Cobourg Formation is further subdivided into lower and

126

upper members, which, respectively comprise ~ 28 m of argillaceous limestone, with minor

127

interbeds of calcareous shale, and ~ 7.5 m of calcareous shale. The Cobourg Formation is

128

capped by ~ 200 m of Ordovician shale which is subdivided into the Blue Mountain, Georgian

129

Bay and Queenston formations. Overall, the Ordovician limestone–shale succession has been

130

interpreted as initial deposition in supra-tidal/tidal flat to lagoonal marine or offshore shallow

131

and deep-shelf settings (Black River and Trenton groups), and open marine to shallow intertidal

132

or supratidal settings for deposition of the Blue Mountain–Queenston shales. Overlying the

133

Ordovician shales are 415 m of Silurian and Devonian interlayered dolostone, shale and

AC C

EP

TE D

M AN U

SC

RI PT

112

7

ACCEPTED MANUSCRIPT

134

evaporitic units, which represent deposition in a carbonate-shelf setting with periods of restricted

135

marine circulation.

136

Herein we refer to “Ordovician limestones”, “Trenton Group limestones”, “Black River Group limestones” and “Cambrian sandstones” as generalized nomenclature for each of the

138

respective units, regardless of the minor occurrence of units/formations of differing lithology

139

(e.g., Cambrian dolostone units, Shadow Lake siltstones/dolostones in the Black River Group,

140

Collingwood shale member at the top of the Cobourg Formation).

SC

141

RI PT

137

Hydrogeologic conditions at the Bruce site have been investigated by Intera (2011), Normani and Sykes (2012), Beauheim et al. (2014), Khader and Novakowski (2014), Neuzil and

143

Provost (2014) and Neuzil (2015). The objective of these studies was to understand the

144

hydraulic properties of the sedimentary rocks through modelling (Normani and Sykes, 2012), in

145

situ hydraulic conductivity measurements (Beauheim et al., 2014) and interpretation of observed

146

anomalies in hydraulic-head distributions (Khader and Novakowski, 2014; Neuzil and Provost,

147

2014; Neuzil, 2015). Overall, the hydraulic data indicate that the permeability of the Upper

148

Ordovician limestones and shales is very low, such that solute transport is limited to diffusion.

149

In situ measurements of horizontal hydraulic conductivity (Kh) by Beauheim et al. (2014)

150

indicate that the Ordovician shales and the argillaceous limestones of the Trenton Group

151

represent an aquiclude (Fig. 1), with Kh values ranging from 10-16 to 10-12 m s-1. Limestones of

152

the Black River Group have Kh values of 10-12 to 10-10 m s-1 and define an aquitard (Fig. 1). The

153

Cambrian sandstone unit has a Kh value of 10-6 m s-1 and represents a basal aquifer (Fig. 1).

TE D

EP

AC C

154

M AN U

142

155

3. Methods

156

3.1 Sampling and Petrography

8

ACCEPTED MANUSCRIPT

157

Samples were obtained from six drill cores (DGR-2 to DGR-6; DGR-8) at the Bruce site, and were initially collected on the basis of visible secondary minerals (calcite, dolomite,

159

anhydrite, celestine, pyrite, halite, etc.) in veins, vugs and replacement/alteration zones, largely

160

within the Trenton and Black River group limestones and the Cambrian sandstones. Sample

161

depths are reported here as metres below ground surface (mBGS) relative to drill hole DGR-2.

162

RI PT

158

Polished thin sections of the secondary minerals were prepared at the University of New Brunswick, and characterized using optical microscopy and electron-beam techniques

164

(Microscopy and Microanalysis Facility at the University of New Brunswick). Secondary

165

minerals were imaged with a JEOL 6400 scanning electron microscope (SEM) using a

166

backscattered electron detector (BSE) and identified using SEM-based energy-dispersive X-ray

167

spectrometry (EDS). Quantitative measurements of major and trace element concentrations (Ca,

168

Mg and Sr) were made using wavelength-dispersive X-ray spectrometry on a JEOL 733 Super-

169

probe electron probe microanalyzer (EPMA), operating at 15 kV with a beam current of 15-20

170

nA. Integration times for each element ranged from 30–80 s and the detection limits were

171

generally lower than 160 ppm for Ca and Mg, and 260 ppm for Sr.

174

M AN U

TE D

EP

173

3.2 Fluid Inclusion Microthermometry Fluid inclusion measurements were undertaken in the Institute of Geological Sciences at

AC C

172

SC

163

175

the University of Bern, and follow the general fluid inclusion methodology outlined by Roedder

176

(1984), Goldstein and Reynolds (1994) and Samson et al. (2003). Double-polished thick

177

sections (150 µm) were prepared for each sample to permit characterization by optical

178

microscopy using transmitted, plane-polarized light and reflected, ultraviolet (UV) light. UV

179

light microscopy was utilized to identify fluid inclusions that contain hydrocarbons.

9

ACCEPTED MANUSCRIPT

180

Microthermometric measurements were made using a Linkam MSD-600 heating-cooling stage mounted on an Olympus BX51 polarizing microscope. The stage was calibrated against

182

synthetic CO2–H2O inclusions and synthetic H2O inclusions with critical density, such that the

183

accuracy is ± 0.1 °C for measurements made below room temperature and ± 1.0 °C for

184

measurements above room temperature. For the microthermometric measurements of inclusions

185

in calcite, careful attention was paid to avoid spurious data and/or anomalously high

186

homogenization temperatures that resulted from inclusion “stretching” (i.e. deformation of the

187

inclusion walls during heating/pressurization; Roedder and Bodnar, 1980; Goldstein and

188

Reynolds, 1994). “Stretched” inclusions were identified by increases in the diameter of the

189

contained vapour bubbles after heating to homogenization and cooling down to room

190

temperature, as visible in photographs taken prior to and after heating.

SC

M AN U

191

RI PT

181

Identification of gas and mineral phases trapped within the inclusions was undertaken using laser Raman spectroscopy. Either red (632.8 nm) He-Ne or green (532.12 nm) Nd-YAG

193

lasers were focused onto the sample through an Olympus 100/0.95 UM PlanFI objective lens on

194

a Olympus BX41 microscope, and Raman spectra were collected for 10–20 s using a Horiba

195

Jobin-Yvon LabRam HR-800 spectrometer calibrated against the emission lines of a neon lamp.

198 199

EP

197

AC C

196

TE D

192

3.3 Stable Isotope Analysis

C- and O-isotope measurements were conducted in the G.G. Hatch Stable Isotope

200

Laboratory at the University of Ottawa using isotope ratio mass spectrometry (IRMS). Calcite

201

and dolomite samples were acquired from thin section chips using a Merchantek microdrill (300

202

µm drill bit). Drill locations were pre-selected in portions of veins, vugs and/or

10

ACCEPTED MANUSCRIPT

alteration/replacement zones using a combination of optical microscopy and SEM-BSE imaging

204

of the associated thin section. The powdered calcite and dolomite samples were weighed into

205

glass reaction vessels and flushed with helium. Orthophosphoric acid was introduced into the

206

vials and reacted for 24 hours at 25 °C for calcite and 50 °C for dolomite. The C- and O-isotope

207

ratios of the resulting CO2 gas were measured using a Thermo-Finnigan Delta XP and Gas

208

Bench II in continuous flow mode, and normalized against NBS-18, NBS-19 and LSVEC

209

(carbon only). C- and O-isotope compositions are reported in δ-notation relative to Vienna

210

Peedee belemnite (VPDB) for carbon and Vienna standard mean ocean water (VSMOW) for

211

oxygen, in permil (‰). Analytical uncertainties associated with the δ13C and δ18O data are ± 0.1

212

‰ (2σ).

M AN U

SC

RI PT

203

213

215

3.4 87Sr/86Sr Analysis

Sr-isotope ratios (87Sr/86Sr) for carbonate samples were obtained in the Isotope

TE D

214

Geochemistry and Geochronology Research Centre (IGGRC) at Carleton University using

217

thermal ionization mass spectrometry (TIMS). Calcite, dolomite, anhydrite and celestine

218

samples were collected by micro-drilling following the same methodology described in Section

219

3.3. Sr-concentrations were determined using mean EPMA analyses for the target mineral.

220

Samples were weighed out between 0.5 and 15 mg (depending on Sr-concentration) and

221

dissolved in 2.5 N nitric acid. Micro-chromatography columns were prepared using 100 µL of

222

Eichrom Sr Spec resin (5 to 100 µm) in a 14 mL borosilicate glass column. The resin was

223

washed with 1200 µL of ultrapure water and then conditioned with 400 µL of 7 N nitric acid.

224

The samples were added to the column, then washed using 7 N nitric acid. Sr ions were eluted

225

from the resin using 1600 µL of ultrapure water, and dried down on a hot plate. The dried Sr

AC C

EP

216

11

ACCEPTED MANUSCRIPT

226

samples were loaded onto Ta filaments using 0.3 N orthophosphoric acid and analyzed for

227

87

228

blanks for the procedure are < 250 picograms. The three-year average for analyses of NIST

229

SRM987 is 87Sr/86Sr = 0.710254 ± 0.000019. Analytical uncertainties associated with the Sr-

230

isotope data are reported at 2σ.

RI PT

Sr/86Sr using a Thermo-Finnigan Triton TIMS, at temperatures of 1300-1500 °C. Total Sr

231 4. Results

233

4.1 Paragenetic sequence and chemistry of secondary minerals

Secondary minerals in the Cambrian sandstones and the Black River and Trenton group

M AN U

234

SC

232

limestones occur as matrix replacement and/or veins and vugs (Fig. 2). The mineral assemblages

236

predominantly consist of calcite and dolomite, with varying proportions of anhydrite, celestine,

237

pyrite and quartz (Fig. 3). Varying amounts of clay minerals and halite also occur as secondary

238

minerals (Intera, 2011). Relative proportions of dolomite versus calcite (as wt.% dolomite) are

239

exhibited in Fig. 4A using bulk XRD data from Intera (2011). Elemental concentrations (Ca, Mg

240

and Sr) are shown in Table 1S (Supporting Material). Saddle dolomite is observed as linings in mm- or cm-wide veins and vugs within: (1)

EP

241

TE D

235

stratabound dolostone in the Sherman Fall, Coboconk, Gull River and Cambrian units, (2) Gull

243

River limestone and (3) Shadow Lake siltstone (e.g., Figs. 2C–E; 2G–H; Fig. 1S – Supporting

244

Material). Replacement dolomite occurs as fine-grained and dolomite-rich, mm- or cm-wide

245

zones in limestone within both the Trenton and Black River groups (e.g., Fig. 2B). Uniform

246

MgO abundances of 18−22 wt.% were obtained for saddle/vein dolomite and dolostone matrix

247

(Fig. 4B). Sr abundances are commonly less than 1000 ppm for all three types of dolomite (Fig.

AC C

242

12

ACCEPTED MANUSCRIPT

248

4C). Dolomitic growth zones in calcite crystals (MgO up to 22 wt.%) are identified in

249

Coboconk, Gull River and Shadow Lake veins and vugs (Fig. 4B). Secondary calcite commonly consists of medium- or coarse-grained, euhedral crystals

251

within mm- to cm-sized veins and vugs, and are the principal vein/vug forming mineral in the

252

Trenton and Black River groups and the Cambrian sandstones (Fig. 1S – Supporting Material).

253

Low MgO abundances (< 2 wt.%) are common for most secondary calcite. Sr abundances in

254

vein and vug calcite exhibit a decreasing trend with depth, from as high as 4300 ppm in the

255

Cobourg to <950 ppm in the Cambrian (Fig. 4C).

SC

Anhydrite and celestine are observed as late infill after saddle dolomite and calcite (Figs.

M AN U

256

RI PT

250

2A; 2D; 2E; Saso, 2013; Fig. 1S – Supporting Material). The abundance of Sr in anhydrite is

258

comparable to calcite in the Cobourg and Sherman Fall formations, but it is higher than calcite in

259

the Gull River Formation (Fig. 4C). Authigenic quartz was identified in one sandstone and one

260

dolostone in the Cambrian, as euhedral quartz rims mantling rounded, detrital quartz cores (Fig.

261

2I), and druse quartz lining the wall of a calcite vein (Fig. 2G). At both of these occurrences,

262

authigenic quartz appears to have formed either prior to or synchronous with calcite formation.

263

In DGR-4-844.78, the quartz lining appears to postdate the growth of saddle dolomite (Fig. 2G).

264

Pyrite is observed throughout the succession, commonly as an early matrix mineral in all

265

limestone units (except the Sherman Fall Formation) and stratabound dolostones in the

266

Coboconk, Gull River, and Cambrian units, and as a late vug mineral, post-dating calcite in

267

DGR-2-843.83 (Fig. 2F).

269

EP

AC C

268

TE D

257

4.2 Petrography and microthermometry of fluid inclusions

13

ACCEPTED MANUSCRIPT

270

Six paleofluid stages were identified from the primary inclusions in dolomite (stage I), primary inclusions in calcite (stage II), primary inclusions in celestine and anhydrite (stage III),

272

and secondary inclusions mostly in calcite (stage IV to VI), and are summarized in Figs. 3–8 and

273

Table 1. Photomicrographs of representative primary and secondary inclusions are shown in

274

Figures 5 and 6, respectively. All inclusion assemblages trapped during stages I to III contain

275

variable proportions of liquid and methane vapour (e.g., Figs. 5B, E, H), indicating

276

heterogeneous fluid entrapment. Therefore, trapping temperatures (Ttrap) for these stages are

277

defined by the minimum homogenization temperature (Th) observed in each inclusion

278

assemblage (Diamond et al., 2015). Uniform proportions of liquid and vapour were observed in

279

assemblages of stage IV inclusions (e.g., Figs. 6B–C, E), thus the maximum Th for each

280

assemblage represents a minimum constraint on Ttrap (i.e. Ttrap > Th,max). All assemblages in

281

stages I to IV contain variable proportions of halite, indicating they were saturated with respect

282

to halite at the time of trapping (principles of these interpretations are explained in Diamond,

283

2003). Stage V inclusions consist of saline brine without vapour bubbles at room temperature

284

(e.g., Fig. 6G). In this case, the only constraint on Ttrap is the empirical rule that they were

285

trapped below 70 °C (Goldstein and Reynolds, 1994; Diamond, 2003). The Ttrap results are

286

summarized in Fig. 7 and Table 1 (detailed Th data are shown in Table 2S – Supporting Material,

287

after Diamond et al., 2015). The relative timing of fluid inclusion entrapment associated with

288

stages IV vs. VI was determined by cross-cutting relationships as shown in Figures 6H and 6I.

SC

M AN U

TE D

EP

AC C

289

RI PT

271

Temperatures of final melting of ice, Tm(ice), in the presence of halite + liquid + vapour

290

were measured in assemblages of inclusions in stages I to IV. No observations of eutectic

291

melting could be made but the span of Tm(ice) values between –52.7 and –34°C suggests the

292

principal salts are NaCl and CaCl2. As commonly observed in CaCl2-rich inclusions, nominally

14

ACCEPTED MANUSCRIPT

stable hydrohalite nucleated only rarely upon cooling and so the measured Tm(ice) transitions

294

represent metastable equilibria. To enable salinity determinations from these data, the locus of

295

the metastable cotectic involving ice + halite + liquid + vapour was calculated for this study (red

296

curve Figs. 8A, B) from the stable phase relations for the H2O–NaCl–CaCl2 ternary system given

297

by Oakes et al. (1990) and Steele-MacInnis et al. (2011). Fluid compositions and salinities in

298

terms of the model ternary components were then obtained for each assemblage by constructing

299

a halite-melting path beginning at Tm(ice) on the cotectic and moving towards the NaCl apex of

300

the ternary. The point at which this path intersects Ttrap on the halite liquidus defines the ternary

301

composition and the salinity of the inclusions (i.e. green arrow in Fig. 8A). For the halite-

302

undersaturated inclusions of stage V, salinity could only be broadly constrained to lie along the

303

relevant isotherms of Tm (ice) (dark bands in Fig. 8B; question marks show range of uncertainty

304

in salinity). Table 1 lists the resulting ternary compositions and salinities. Microthermometric

305

data of all individual inclusions measured are given in Diamond et al. (2015).

SC

M AN U

TE D

306

RI PT

293

Primary, stage I inclusions in saddle dolomite (e.g., Figs. 5A–C) from the Coboconk and Gull River formations yielded the highest Ttrap values; 122 to 128 °C (n = 3) and a Ttrap value of

308

88 °C was obtained for primary inclusions in rhombic dolomite from the Cambrian. Stage I

309

inclusions have Tm (ice) values between −44.1 and −41.5 °C, implying that the fluid had a

310

salinity of 33 to 34 wt.% [CaCl2+NaCl]eq at halite saturation (at the time of Ttrap).

AC C

311

EP

307

Primary stage II inclusions in calcite (e.g., Figs. 5D–F) yielded Ttrap values between 54

312

and 78 °C (mean and S.D. of 65 ± 9 °C; n = 7) and Tm (ice) values between −47.9 and −35.9 °C.

313

Inclusions in two Coboconk samples yielded Tm (hydrohalite) values from −22.9 to −16.5 °C.

314

The calculated salinity of the halite-saturated stage II fluid at Ttrap varies between 30.7 and 33.2

315

wt.% [CaCl2+NaCl]eq.

15

ACCEPTED MANUSCRIPT

316

Stage III is represented by primary inclusions in celestine and anhydrite (e.g., Figs. 5G–I) and by early secondary inclusions in stage II calcite. Crystallites of anhydrite were occasionally

318

identified in stage III inclusions in calcite (Diamond et al., 2015). The inclusions in celestine

319

and calcite yielded Ttrap values between 42 and 80 °C and Tm (ice) values from −46.5 to −33.2

320

°C, corresponding to salinities of 32 to 33 wt.% [CaCl2+NaCl]eq for the halite-saturated stage III

321

fluid at Ttrap.

RI PT

317

Secondary, stage IV inclusions in calcite (e.g., Figs. 6A–E) have minimum Ttrap values

323

between 57 and 106 °C. Their Tm (ice) values vary from −52.7 to −26.4 °C, which indicates a

324

narrow range of halite-saturated salinities between 32 and 34 wt.% [CaCl2+NaCl]eq for this gas-

325

undersaturated fluid at minimum Ttrap.

M AN U

326

SC

322

Late secondary, liquid-only (metastable) inclusions associated with stage V (e.g., Figs. 6F–G) formed at temperatures <70 °C (Goldstein and Reynolds, 1994) and were identified in

328

calcite and dolomite throughout the sample suite. Stage V inclusions from the Gull River

329

Formation and the Cambrian show Tm(ice) values of −35.0 to −24.3 °C, indicating salinities of

330

23 to 26 wt.% [CaCl2+NaCl]eq.

Hydrocarbon inclusions (stage VI) in calcite (e.g., Figs. 6H–I) from the Coboconk and

EP

331

TE D

327

Gull River formations contain variable proportions of liquid oil and methane vapour, indicating

333

they were gas-saturated at the time of trapping. Their minimum Th values yield Ttrap values of 48

334

and 59 °C, respectively.

335 336 337 338

AC C

332

4.3 C- and O-isotope Compositions Calcite vein and vug samples from the Cambrian sandstones yielded the most 13Cdepleted C-isotope compositions in the sedimentary succession, with δ13C values of −6.1 to

16

ACCEPTED MANUSCRIPT

−3.5‰ (Fig. 9A; Table 3S – Supporting Material). Cambrian dolomite samples, including one

340

dolostone and one saddle dolomite, have δ13C values of −2.5‰ and −3.8‰, respectively.

341

Similar C-isotope compositions were obtained for two samples from a vug in the Shadow Lake

342

Formation, including δ13C values of −3.4‰ for saddle dolomite and −5.0‰ for calcite (Table 3S

343

– Supporting Material).

344

RI PT

339

A pronounced C-isotope shift of +2 to +3‰ occurs at the boundary between the Shadow Lake and Gull River Formations. Uniform C-isotope compositions are observed for most

346

samples from the base of the Gull River Formation to the Cobourg Formation, with δ13C values

347

of −1.0 to +1.9‰ and no identifiable differences between the Black River and Trenton group

348

samples. Similarly, no differences in δ13C are observed between samples of limestone matrix

349

and calcite veins and vugs, and samples of saddle dolomite, dolostone and matrix replacement

350

dolomite. Although most of the C-isotope data are constant versus depth, three calcite vein

351

samples from a 3 m section at the base of the Kirkfield Formation (DGR-6-864.60; DGR-6-

352

865.20; DGR-6-867.60) vary by 2.7‰ (−0.8 to +1.9‰; Fig. 9A). It should be noted, however,

353

these data fall within the broader range of C-isotope compositions defined by other samples from

354

the Ordovician succession, including a limestone from the Gull River Formation (−1.0‰) and

355

dolostone from the Coboconk Formation (+1.9‰).

M AN U

TE D

EP

AC C

356

SC

345

Calcite vein and vug samples from the Cambrian have δ18O values between +14.6‰ and

357

+24.2‰ and overall represent the most 18O-depleted compositions in the succession, with only

358

one of these Cambrian calcite samples higher than +17.8‰, (Fig 9B; Table 3S – Supporting

359

Material). Two dolomite samples from the Cambrian units have δ18O values of +23.0‰

360

(dolostone) and +23.2‰ (saddle dolomite). Two samples from a vug in the Shadow Lake

361

Formation (DGR-2-843.83) yielded similar O-isotope compositions to the other respective

17

ACCEPTED MANUSCRIPT

362

calcite and dolomite samples in the Cambrian sample suite, and include δ18O values of +15.9‰

363

for calcite and +20.6‰ for saddle dolomite.

364

A pronounced shift of ~ +3 to +7‰ is observed in the δ18O data along the boundary between the Shadow Lake and Gull River formations (Fig. 9B). Calcite vein and vug samples

366

from the Gull River and Coboconk formations have a wide range of δ18O values, from +18.9 to

367

+26.7‰. Notably, a ~ 10 m section at the top of the Coboconk Formation, along the Trenton–

368

Black River Group boundary, exhibits a spread in δ18O values of up to 5.4‰ (+21.3 to +26.7‰).

369

Calcite veins and vug samples from the Trenton Group have δ18O values from +23.0 to +28.1‰,

370

and are broadly consistent with calcite O-isotope data for the Black River Group. Most of the

371

Trenton Group calcite samples vary by up to 3.1‰, with only one exceptional δ18O value higher

372

than +26.1‰ (Fig. 9B). Six saddle dolomite samples from the combined Black River–Trenton

373

group units have uniform δ18O values of +20.3 to +21.3‰, and are > 1–2‰ lower in δ18O than

374

most calcite vein and vug samples in the Ordovician succession. In contrast, one sample of

375

dolostone matrix (DGR-2-778.44) and one replacement dolomite sample (DGR-4-772.10) from

376

the Coboconk Formation yielded δ18O values of +23.6‰ and +22.0‰, respectively. Eleven

377

limestone matrix samples from the Trenton and Black River groups exhibit a subtle trend of

378

increasing δ18O values towards the surface, from +23.8 to +26.8‰, and no identifiable shifts or

379

sections of O-isotope variability along the boundary between the two groups, as observed in the

380

calcite vein and vug data (Fig. 9B).

382 383 384

SC

M AN U

TE D

EP

AC C

381

RI PT

365

4.4 Sr-isotope ratios Secondary calcite and dolomite samples from the Cambrian units have the highest Srisotope ratios in the sedimentary succession. Four calcite vein/vug samples have 87Sr/86Sr values

18

ACCEPTED MANUSCRIPT

from 0.70975 to 0.71039, while one sample of dolostone matrix and one saddle dolomite yielded

386

similar 87Sr/86Sr values of 0.71043 and 0.71027, respectively (Fig 9C; Table 3S – Supporting

387

Material). Furthermore, two calcite/saddle dolomite samples from a vug in the Shadow Lake

388

Formation have 87Sr/86Sr values of 0.71037 and 0.70984, and overlap in Sr-isotope composition

389

with other samples from the Cambrian.

RI PT

385

A shift in 87Sr/86Sr of up to ~ –0.02 is observed across the boundary between the Shadow

390

Lake and the Gull River formations. Calcite vein and vug samples from the Gull River and

392

Coboconk formations have the highest Sr-isotope variability in the succession, with 87Sr/86Sr

393

values ranging from 0.70799 to 0.70990 (Fig. 9C). Similar variability is observed for three

394

samples of saddle dolomite, one dolostone and one sample of anhydrite from the Gull River–

395

Coboconk units, with combined 87Sr/86Sr values from 0.70854 to 0.70939. In contrast, three

396

limestone samples from these Black River Group units (two Gull River, one Coboconk) have

397

uniform 87Sr/86Sr values of 0.70810 to 0.70833, and overlap with the lowest Sr-isotope ratios of

398

calcite veins from these units.

TE D

M AN U

SC

391

399

For the Trenton Group, five calcite vein/vug samples and three limestone matrix samples have uniform 87Sr/86Sr values of 0.70790 to 0.70817 (Fig. 9C), and are consistent with the

401

87

402

values of 0.70818, 0.70808 and 0.70848 are observed for anhydrite samples from the Cobourg

403

and Sherman Fall formations, and a celestine sample from the Kirkfield Formation, respectively

404

(Fig. 9C). One sample of saddle dolomite from a vug in the Sherman Fall Formation (DGR-5-

405

743.10) has an exceptional 87Sr/86Sr value of 0.70856. However, this sample has an associated

406

2σ analytical uncertainty of ± 0.00083, which is 20–40 times larger than other analyses (Table

407

3S – Supporting Material).

EP

400

AC C

Sr/86Sr values of limestone samples from the underlying Black River Group. Similar 87Sr/86Sr

19

ACCEPTED MANUSCRIPT

408 409

5. Discussion

410

5.1 Paragenetic sequence Overall six distinct fluid events were identified in the microthermometric/petrographic

RI PT

411

results (Figs. 3 and 7), with the first two stages responsible for most of the mineral precipitation

413

in veins and vugs and only minor amounts of anhydrite and celestine that formed during stage

414

III. Based on the preservation of well-defined dolomite, calcite, anhydrite and celestine crystal

415

margins from precipitation events during stages I to III (Fig. 2), and on the absence of dissolution

416

features, fluid stages IV–VI are interpreted as migration of relatively minor quantities of fluids

417

along grain boundaries or micro-fractures.

M AN U

418

SC

412

Stage I is represented by formation of Fe- and Mn-enriched saddle dolomite in the Black River Group (Diamond et al., 2015) that resulted from the influx of a halite-saturated brine (34

420

wt.% [CaCl2+NaCl]eq) and free methane gas at temperatures of 122–128 °C. Elsewhere in the

421

Michigan Basin, high temperatures have been determined for fluid inclusions in saddle dolomite

422

from the Trenton–Black River groups, including up to ~160 °C in the Albion-Scipio oil field of

423

south-central Michigan (Allan and Wiggins, 1993) and temperatures as high as ~170 °C and 220

424

°C in dolomite samples from Manitoulin Island and southwestern Ontario, respectively (Coniglio

425

et al., 1994). Elevated temperatures of ~97–144 °C were also obtained by Haeri-Ardakani et al.

426

(2013) for saddle dolomite samples from the Trenton Group in southwestern Ontario. Davies

427

and Smith (2006) suggested that the formation of saddle dolomite in these fault-controlled

428

reservoirs is due to ascending fluid that is hotter than the ambient or burial temperatures of the

429

host formation (i.e. hydrothermal). In our study, a third Fe- and Mn-enriched saddle dolomite

430

sample from the Cambrian yielded a significantly lower Ttrap value of 88 °C, but a similarly high

AC C

EP

TE D

419

20

ACCEPTED MANUSCRIPT

431

salinity of 34 wt.% [CaCl2+NaCl]eq. It is not known whether these saddle dolomite samples are

432

contemporaneous.

433

Stage II primary inclusions in calcite have Ttrap values of 54–78 °C (mean 65 ± 9 °C; n = 7) and represent a halite-saturated brine with a salinity of 31 to 37 wt.% [CaCl2+NaCl]eq and

435

entrained free methane gas. Based on the predominance of calcite as a vein/vug mineral, it can

436

be presumed that stage II was the most significant vein- and vug-forming fluid event. The range

437

of Ttrap values from stage II inclusions is slightly lower than minimum paleotemperature

438

estimates of 80 °C at the base of the Silurian and 90 °C at the base of the Trenton Group reported

439

by Legall et al. (1981). Determinations of the absolute timing of vein and vug calcite formation

440

in the Trenton–Black River groups have been made using U-Pb dating methods with laser

441

ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) and thermal ionization

442

mass spectrometry (TIMS) (Davis et al., 2013). Davis et al. (2013; 2017) reported a LA-ICP-MS

443

U-Pb age of 434 ± 5 Ma, based on a calculated mean of three model age determinations from the

444

Cobourg and Coboconk formations, and a TIMS U-Pb date of 451 ± 38 Ma for a calcite vein

445

from the Cobourg Formation. The Davis et al. (2013; 2017) U-Pb data compare well with a Rb-

446

Sr isochron age of 425 ± 6 Ma (errorchron age of 440 ± 50 Ma) by Harper et al. (1995) and 7

447

combined K-Ar dates ranging from 453 ± 9 Ma to 412 ± 8 Ma by Harper et al. (1995) and

448

Ziegler and Longstaffe (2000) (weighted mean of 442 ± 16 Ma) for secondary K-feldspar

449

obtained along the Precambrian−Paleozoic boundary in southern Ontario. These data suggest the

450

influx of stage II, calcite-forming fluids in the Trenton–Black River groups likely occurred

451

between the Late Ordovician and Silurian.

452 453

AC C

EP

TE D

M AN U

SC

RI PT

434

Stage III fluid inclusions have Ttrap values of 42–80 °C and represent a stage of fluid migration during which anhydrite and celestine were formed. The mean Ttrap value of 62 ± 17 °C

21

ACCEPTED MANUSCRIPT

for four stage III inclusions is similar to the mean Ttrap value (65 ± 9 °C; n = 7) for stage II

455

primary inclusions in calcite. Salinities of stage III inclusions range from 32 to 33 wt.%

456

[CaCl2+NaCl]eq, and are comparable to the salinity estimates for stage I and stage II inclusions.

457

This suggests a paragenetic link between the stage I–III fluids, with saddle dolomite forming

458

during initial brine influx at 122–128 °C (stage I), then subsequent cooling and formation of

459

calcite (stage II) and penecontemporaneous formation of anhydrite and celestine (stage III). Secondary fluid inclusions defining stage IV have minimum Ttrap values between 57 and

SC

460

RI PT

454

106 °C, including one calcite vein from the Sherman Fall Formation (DGR-2-691.70; see Figs.

462

6D–E) that yielded minimum Ttrap values ranging from 77 to 106 °C. The absolute timing and

463

burial depth (i.e. thickness of overburden sediments) at the time of stage IV fluid inclusion

464

entrapment is unclear, therefore, definitive pressure corrections to determine Ttrap (true) cannot

465

be applied to the stage IV data. In order to determine an upper limit for stage IV fluid inclusion

466

entrapment, Ttrap (maximum) values were calculated using pressure corrected data for peak burial

467

conditions for the Ordovician–Cambrian succession. This is a conservative approach because we

468

do not know that stage IV coincides with peak burial, and because all corrections are made to the

469

depth at the base of the Paleozoic succession. After peak burial conditions, it can be inferred

470

from the work of Coniglio and Williams-Jones (1992) that approximately 1000 m of Paleozoic

471

sediments were eroded from the sedimentary succession at the Bruce site (NWMO, 2011).

472

Utilizing an overburden thickness of 1000 m and a current depth of 860 m for the base of the

473

Paleozoic succession, a maximum burial depth of 1860 m was used to calculate an upper limit

474

for stage IV fluid inclusion entrapment. Pressure corrected Ttrap (maximum) values for the stage

475

IV inclusions range from 77–127 °C (Fig. 7; Table 1).

AC C

EP

TE D

M AN U

461

22

ACCEPTED MANUSCRIPT

There are two models that explain the formation of stage IV fluids: (1) influx of

477

hydrothermal brines, or (2) regional heating of the sediments (i.e. through conduction), coupled

478

with fracturing and re-mobilization of fluids. For model 1, significant quantities of fluid and

479

high flow rates along permeable structures would be required to facilitate the transfer of heat, but

480

there is no evidence for fluid movement at this scale (e.g., large fracture/veins systems,

481

dissolution/re-precipitation features, formation of new secondary minerals with primary

482

inclusions analogous to stage IV). Alternatively, a regional heating event (model 2) is consistent

483

with an estimated minimum paleotemperature of 90 °C at the base of the Trenton Group (Legall

484

et al., 1981). Evidence for a regional heating event exists elsewhere in the Michigan Basin.

485

Temperatures of ~70–170 °C and ~40–260 °C have been obtained for secondary fluid inclusions

486

in calcite from Trenton–Black River rocks on Manitoulin Island and in southwestern Ontario,

487

respectively (Coniglio et al., 1994). Haeri-Ardakani et al. (2013) also reported temperatures of

488

~60–131 °C for fluid inclusions in calcite cement samples from the Trenton Group in southwest

489

Ontario. In central Michigan, fluid inclusions in Middle Ordovician St. Peter Formation

490

sandstones yielded temperatures of 76–169 °C for quartz overgrowths, and 94–149 °C for

491

dolomite cements (Girard and Barnes, 1995), and are consistent with temperatures of ~ 100–160

492

°C obtained for fluid inclusions in saddle dolomite from elsewhere in central Michigan (Allan

493

and Wiggins, 1993). Overall, these high temperatures (up to 260 °C) strongly suggest that

494

regional heat sources were present at depth below the Michigan Basin. Reactivation of the mid-

495

continent rift was suggested by Ma et al. (2009) as a source of heat and mantle-derived He and

496

Ne anomalies in the hydrothermal fluids that interacted with the basin rocks. Regional-scale

497

heating has also been linked to the formation of secondary illite at 367–322 Ma in the St. Peter

498

sandstone along the base of the Paleozoic succession (Girard and Barnes, 1995). Reactivation of

AC C

EP

TE D

M AN U

SC

RI PT

476

23

ACCEPTED MANUSCRIPT

499

the mid-continent rift during the Late Devonian–Mississippian could be responsible for initiating

500

regional-scale heating below the Michigan Basin and may be linked to the stage IV fluid

501

migration event observed in the Ordovician carbonates at the Bruce site. Stage V and VI inclusions represent the latest observed records of fluid migration in the

503

sedimentary succession, and for the latter stage VI fluids, include hydrocarbon inclusions in the

504

Coboconk and Gull River formations and the Cambrian units. Samples from the Coboconk and

505

Gull River formations yielded trapping temperatures of 48 and 59 °C (Fig. 7). The occurrence of

506

hydrocarbon fluids is consistent with the presence of bitumen layering or oil in the Coboconk,

507

Gull River and Shadow Lake formations (see Fig. 2.32 in Intera, 2011), suggesting that stage VI

508

represents an important period of hydrocarbon generation and/or migration in the Black River

509

Group.

510 5.2 Source and evolution of fluids

TE D

511

M AN U

SC

RI PT

502

By combining the petrographic and fluid inclusion results with the C- O- and Sr-isotope

513

data, inferences can be made about the sources and paleofluid evolution through the secondary-

514

mineral formation events of stages I to III. Secondary calcite (veins/vugs) and dolomite

515

(dolostone matrix and saddle) samples from the combined Cambrian–Shadow Lake units have

516

δ13C values of ~ −6.1 to −2.5‰ (Fig. 9A). Estimates of the average C-isotope composition of

517

Cambrian seawater include δ13C values between −0.9‰ and +0.0‰ (Fig. 9A; Veizer et al.,

518

1999), suggesting that the calcite and dolomite samples from the Cambrian–Shadow Lake units

519

were not sourced from connate fluids derived from Cambrian seawater. Alternatively, these low

520

δ13C values reflect carbonate mineral formation from fluids that reacted with organic carbon.

521

Harper et al. (1995) reported similar 13C-depleted data (−9.3 to −3.9‰) for secondary calcite and

AC C

EP

512

24

ACCEPTED MANUSCRIPT

522

dolomite associated with hydrothermally altered granitoid rocks along the

523

Precambrian−Paleozoic boundary in southern Ontario.

524

A 2−3‰ shift in δ13C occurs at the boundary between Cambrian−Shadow Lake units and the overlying Gull River Formation. The C-isotope compositions of calcite and dolomite

526

samples show little scatter through the Gull River to Cobourg formations (Fig. 9A) and are

527

consistent with estimates by Veizer et al. (1999) for the average C-isotope composition of

528

seawater during the Late Ordovician (−0.1 to +1.7‰), suggesting a C source dominated by

529

seawater-derived connate fluids or external fluids that experienced high degrees of fluid–rock

530

interaction with the host limestones (Fig. 9A).

SC

M AN U

531

RI PT

525

Calcite samples from the Shadow Lake and Cambrian units generally have the lowest δ18O values (+14.6 to +24.2‰), with most calcite veins and vugs plotting at ~ +16‰ (7 of 8

533

calcite veins form a calculated mean of +15.9 ± 1.0‰; S.D.). A prominent feature of the O-

534

isotope data through the Cambrian−Ordovician succession is the >4‰ shift that occurs along the

535

boundary between the Cambrian–Shadow Lake units and the overlying Gull River Formation

536

(Fig. 9B). Relatively high variability is observed in the δ18O data for the Coboconk Formation

537

just below the Trenton−Black River group boundary (Fig. 9B), with values from calcite vein and

538

vug samples exhibiting a spread of up to 5.4‰. This variability could be explained by: (1)

539

mixing between two or more fluids that had distinct O-isotope compositions (i.e. connate vs.

540

external hydrothermal fluids), or (2) temperature-related differences in ∆18Ocalcite−fluid during the

541

precipitation of calcite (e.g., Friedman and O'Neil, 1977). The fact that large temperature

542

differences are not observed in the primary fluid inclusion data for the Coboconk Formation

543

(Fig. 7) suggests that the δ18O variability may be explained by mixing with external fluids. The

544

Trenton−Black River boundary represents a significant change in hydraulic conductivity, with

AC C

EP

TE D

532

25

ACCEPTED MANUSCRIPT

rocks in the Trenton Group having Kh values up to six orders of magnitude lower than those of

546

the Black River Group (Normani and Sykes, 2012; Beauheim et al., 2014; Khader and

547

Novakowski, 2014). Furthermore, thin dolostone units (< 20 cm) and a bentonite unit (< 10 cm

548

thick) occur as stratabound layers in the Coboconk Formation (~ 790.50 m and 781.00 mBGS,

549

respectively). These units, particularly the dolostone, may have relatively high permeability and

550

porosity compared to the bounding limestones, and as such, could represent pathways of

551

enhanced lateral flow, allowing for ingress of external fluids.

SC

552

RI PT

545

The O-isotope compositions of source fluids responsible for secondary mineral formation were estimated using the microthermometric data and mineral−fluid O-isotope exchange

554

thermometers for calcite (Friedman and O'Neil, 1977; Fig. 10A) and dolomite (Horitia, 2014;

555

Fig. 10B). The O-isotope composition of Late Ordovician seawater is shown in Figure 10, with

556

δ18O values between −4 and −2‰ (Veizer et al., 1999; Shields et al., 2003). Calculated values

557

for the O-isotope composition of fluids in equilibrium with all limestones from the Gull River to

558

Cobourg formations are consistent with Ordovician seawater (Fig. 10A), indicating that there has

559

been limited diagenetic or hydrothermal alteration effects to their primary O-isotope

560

compositions.

TE D

EP

561

M AN U

553

Calcite vein and vug samples from the Cambrian−Shadow Lake units have calculated δ18Ofluid values between ~ −6‰ and +4‰, based on formation temperatures of 70 to 72 °C (Fig.

563

10A). Gull River and Coboconk samples have similar calculated δ18Ofluid values of ~ −6 to

564

+4‰, while samples from the Trenton Group have calculated fluid compositions of ~ +2 to +8‰

565

(Fig. 10A). These fluid compositions are suggestive of calcite precipitation from Late

566

Ordovician seawater or an 18O-enriched, seawater-derived basin fluid. Similar 18O-enrichments

567

relative to seawater have been described for fluids in many sedimentary basins (Holser, 1979;

AC C

562

26

ACCEPTED MANUSCRIPT

Knauth and Beeunas, 1986; Ayalon and Longstaffe, 1988; Kyser and Kerrich, 1990; Wilson and

569

Long, 1993a; 1993b; Hanor, 1994), and are interpreted to have undergone 18O-enrichment by

570

evaporation and/or fluid−rock interaction following burial. Modification of seawater-derived

571

connate fluids by influx of hydrothermal fluids that interacted with, or were derived from the

572

Precambrian shield (Spencer, 1980) could also explain such 18O-enrichments.

573

RI PT

568

Calculated fluid compositions for saddle dolomite have δ18Ofluid values of ~ 0 to +7‰ (Fig. 10B) and overlap with the calculated fluid compositions of calcite, suggesting that dolomite

575

precipitated from 18O-enriched hydrothermal fluids originating from evolved seawater. No

576

microthermometry data could be obtained from fluid inclusions in samples of matrix replacement

577

dolomite and dolostone. However, if the matrix replacement dolomite and dolostone samples

578

formed at temperatures similar to those of calcite veins and vugs (54 to 78 °C) or saddle

579

dolomite (88 to 128 °C), the calculated fluid compositions would be −6 to +1‰ or +1 to +7‰,

580

respectively (Fig. 10B). These calculated fluid compositions are suggestive of matrix

581

replacement dolomite and dolostone formation directly from Late Ordovician seawater or

582

hydrothermal fluids derived from evolved seawater, and are broadly consistent with fluid source

583

determinations for all other secondary minerals.

M AN U

TE D

EP

584

SC

574

Sr-isotope data for many of the samples from the Black River Group, and all samples from the Cambrian, exhibit 87Sr/86Sr enrichment well above the seawater curve of Veizer et al.

586

(1999). Enriched 87Sr/86Sr values (> 0.710) are a common signature of Precambrian brines (e.g.,

587

Frape et al., 1984; McNutt et al. 1984; Bottomley et al., 1999), and the observed enrichment

588

suggests vein and vug formation from shield-derived hydrothermal fluids, and/or fluid–rock

589

interaction with shield-derived detrital minerals in the Cambrian sandstones. In contrast, all vein

590

and vug samples from the Trenton Group, and approximately half of the samples from the Black

AC C

585

27

ACCEPTED MANUSCRIPT

River Group, overlap with the seawater curve of Veizer et al. (1999), suggesting they formed

592

from either connate fluids derived from Late Ordovician seawater and/or external hydrothermal

593

brines that underwent high degrees of fluid–rock interaction and associated modification of their

594

original isotopic signature during transit through the underlying stratigraphy.

RI PT

591

The combined microthermometry and isotopic data are consistent with a model of

596

secondary mineral formation by fluid mixing; with one end member being either a shield-derived

597

hydrothermal fluid, and/or a fluid that was isotopically influenced by detrital minerals in the

598

Cambrian sandstone, and the second end member being a connate pore fluid derived from the

599

Paleozoic carbonates.

M AN U

SC

595

600 601 602

5.3 Implications for the migration of high-temperature fluids in the Ordovician limestones An important feature of our fluid influx model for the Ordovician sedimentary succession is the formation of dolomite in the Black River Group at temperatures of 122–128 °C, which are

604

higher than the nearest available estimate of peak burial temperature, 55–90 °C from Coniglio

605

and Williams-Jones (1992) at Manitoulin Island, providing evidence for migration of

606

hydrothermal fluid.

EP

TE D

603

Hydrothermal dolomite is an important reservoir for oil and gas plays within the

608

Trenton−Black River groups in the Albion-Scipio/Stoney Point fields in south-central Michigan

609

(Hurley and Budros, 1990; Allan and Wiggins, 1993; Davies and Smith, 2006), the Lima-Indiana

610

field in Ohio/Indiana (Wickstrom et al., 1992) and southern Ontario (Powell et al., 1984;

611

Middleton et al., 1993; Coniglio et al., 1994; Obermajer et al., 1999). In these regions,

612

hydrothermal dolomite developed along large-scale extensional faults as a result of high-

613

temperature fluid−rock interaction (up to ~ 260 °C; Coniglio et al., 1994) between limestone and

AC C

607

28

ACCEPTED MANUSCRIPT

hydrothermal fluids that migrated upward from the underlying Cambrian aquifer. Hydrothermal

615

dolomite reservoirs are recognized in 2D seismic data aligned with vertical “sag” or “flower”

616

structures, often 10s of kilometers in strike length and 100s of meters in width (e.g., Davies and

617

Smith, 2006). At that scale, these structures represent regionally extensive zones of

618

dolomitization. At the Bruce site, however, the abundance and scale of hydrothermal dolomite

619

occurrence is much different. Saddle dolomite with associated high fluid trapping temperatures

620

was observed in veins within Ordovician limestone, stratabound dolostone in the Ordovician, and

621

in Cambrian dolostone. These features are typically mm or cm in thickness (e.g., Fig 2B–E; 2G;

622

2H; Fig. 1S – Supporting Material), and as such, do not represent reservoir-scale structures of

623

dolomitization as observed elsewhere in the Michigan Basin.

M AN U

SC

RI PT

614

624 625

5.4 Conceptual model

Models for the source and timing of paleofluid movements in the Cambrian and

627

Ordovician sedimentary units below the Bruce site, until now, have been largely based on direct

628

geochemical and isotopic constraints from porewater extracted from drill cores (Clark et al.,

629

2010a; 2010b; 2013; 2015; Al et al., 2015). In a detailed study, Clark et al. (2013) reported O-

630

and H-isotope compositions and Cl- and Br-concentration data for porewater obtained

631

throughout the Paleozoic sedimentary succession. Downward trends of decreasing Cl− and Br−

632

concentrations, coupled with decreases in δ18O, were observed in the Ordovician limestones

633

(Fig. 2 in Clark et al., 2013). These trends were interpreted to represent a mixing profile

634

between hypersaline evaporated seawater from the overlying Silurian Salina Formation and

635

shield-derived fluids (Clark et al., 2013; Al et al. 2015). The influence of a shield end member is

636

supported by a downward trend of 2H-enrichment – a common feature of shield fluids (Fritz and

AC C

EP

TE D

626

29

ACCEPTED MANUSCRIPT

637

Frape, 1982; Frape et al., 1984; Pearson, 1987; Bottomley et al., 1999, 2004; Douglas et al.,

638

2000; Gascoyne, 2004; Al et al., 2015).

639

Clark et al. (2013) also reported He-isotope ratios for porewater from the Ordovician shales and the Trenton and Black River group limestones. Porewater extracted from the shales

641

and the Cobourg Formation limestone yielded uniform R/RA values (0.020 ± 0.002; where R/RA

642

= (3He/4He)sample/(3He/4He)air, and (3He/4He)air = 1.38 x 10-6), which are consistent with an

643

authigenic origin, and the total He concentrations indicated a minimum period of 260 Ma for He

644

ingrowth. A positive shift in R/RA values was observed from the base of the Cobourg to the

645

upper portion of the Sherman Fall Formation, followed by uniform R/RA values of 0.035 ± 0.002

646

with increasing depth into the Cambrian (Fig. 4; Clark et al., 2013). The elevated R/RA values

647

below the Cobourg Formation limestone were interpreted by Clark et al. (2013) to reflect

648

allochthonous fluids of mixed basin and shield origin.

SC

M AN U

Furthermore, 87Sr/86Sr data were reported by Clark et al. (2010a; 2010b) for porewater

TE D

649

RI PT

640

samples obtained at varying depths in the sedimentary succession. Porewater samples from the

651

Ordovician shales yielded uniform Sr-isotope compositions (mean 87Sr/86Sr value of ~ 0.710)

652

that are enriched in radiogenic Sr relative to the seawater curve of Veizer et al. (1999). Similar

653

to the interpretation of the He isotope data, Clark et al. (2010a; 2010b) considered the 87Sr

654

enrichment to have resulted from in situ ingrowth by decay of 87Rb in the porewater. Porewater

655

samples from the Trenton and Black River group limestones yielded highly variable 87Sr/86Sr

656

values, between 0.7093 and 0.7103, and consistent with the He-isotope data, were interpreted as

657

resulting from mixing between basin- and shield-derived fluids.

658 659

AC C

EP

650

The isotopic and fluid inclusion data from secondary minerals are generally consistent with the conceptual model for porewater evolution presented by Clark et al. (2013) and Al et al.

30

ACCEPTED MANUSCRIPT

(2015), but our new data allow for some refinements to the model. Based on the porewater data,

661

Clark et al. (2013) and Al et al. (2015) suggest that mixing between basin- and shield-derived

662

fluids was driven by diffusion, but the secondary mineral data indicate that advection in small

663

fractures also contributed to upward transport of shield-type, halite-saturated fluids during stages

664

I to III. These events formed the secondary vein minerals and are broadly constrained by U-Pb

665

dating (Davis et al., 2013) to the Late Ordovician or Silurian. The isotopic signature of the

666

shield end member does not extend upward into the Trenton Group, likely because it was

667

eliminated by fluid−rock interaction. The fact that present-day porewater in the Black River

668

Group is below halite saturation is consistent with indications of allochthonous porewater from

669

87

670

fluid migration that are not recorded in the chemical and isotopic composition of the vein

671

minerals.

M AN U

SC

RI PT

660

Sr/86Sr and 3He/4He ratios, and likely reflects the later stages (V to VI) of halite-undersaturated

673 674

Conclusions

TE D

672

Overall, the combined microthermometric and isotopic data allow us to establish a model for the fluid migration history associated with the formation of secondary vein minerals in low-

676

permeability Ordovician limestones of the Michigan Basin. Six stages of fluid migration are

677

suggested by the fluid inclusion data. These comprise fluid stages I through III, which are

678

represented by primary inclusions trapped during the formation of dolomite (stage I: 88–128 °C),

679

calcite (stage II: 54–78 °C) and celestine/anhydrite (stage III: 42–80 °C). The ubiquitous

680

presence of variable amounts of halite in these fluid inclusions indicates that the secondary vein

681

minerals precipitated from halite-saturated brines with salinities of 31–37 wt.% [CaCl2+NaCl]eq,

682

and is suggestive of a paragenetic link between fluid stages I–III.

AC C

EP

675

31

ACCEPTED MANUSCRIPT

683

Secondary, stage IV inclusions in calcite have a similar, halite-saturated fluid composition as stages I–III, but were trapped above 57–106 °C. No new secondary vein

685

minerals formed during this stage and dissolution features are absent from pre-existing secondary

686

minerals. The stage IV inclusions are interpreted as re-mobilization of trapped fluids along grain

687

boundaries or micro-fractures during a Late Devonian–Mississippian regional heating event.

688

Stages V and VI represent the latest recognized fluid migrations in the sample suite, during

689

which a halite-undersaturated brine (23–26 wt.% [CaCl2+NaCl]eq) was trapped at temperatures <

690

70 °C.

Secondary minerals from the Cambrian and Shadow Lake units are characterized by the

M AN U

691

SC

RI PT

684

lowest δ13C and δ18O values in the succession and 87Sr/86Sr ratios that reflect formation from

693

radiogenic Sr sources. The combined Cambrian–Shadow Lake isotopic data are suggestive of

694

secondary mineral formation from hydrothermal brines that experienced high degrees of fluid–

695

rock interaction with the underlying shield, and/or with shield-derived minerals in the Cambrian

696

sandstone. For the Coboconk and Gull River formations, greater variability in the δ18O and

697

87

698

connate fluids and hydrothermal brines, with the latter having a shield signature sourced from

699

below. Uniform C- O- and Sr-isotope compositions for Trenton Group samples are suggestive of

700

secondary mineral formation from either connate fluids, which were sourced directly from Late

701

Ordovician seawater (or 18O-enriched, evolved seawater), or mixed with hydrothermal brines

702

that underwent high degrees of fluid–rock interaction.

703

TE D

692

AC C

EP

Sr/86Sr values of the secondary minerals is interpreted to reflect mixed fluid sources, including

Age constraints on the absolute timing of secondary mineral formation (stages I–III) are

704

gleaned from previous geochronology studies in the region, which indicate that brine influx

705

occurred during the Late Ordovician to Silurian. Secondary inclusions associated with stage IV–

32

ACCEPTED MANUSCRIPT

VI inclusions did not form new secondary vein minerals and these late fluid stages are

707

interpreted as re-mobilization and/or minor fluid migration along grain boundaries and micro-

708

fractures during a regional heating event. The timing of this regional heating event cannot be

709

easily constrained from our data, however, elsewhere in the Michigan Basin it has been

710

suggested that reactivation of the mid-continent rift during the Late Devonian–Mississippian

711

could be responsible for regional-scale heating and associated fluid migration. Despite evidence

712

elsewhere in the Michigan Basin for secondary-mineral formation during the Late Devonian–

713

Mississippian, the Ordovician limestones at the Bruce site show no evidence of large-scale fluid

714

ingress during this time period (e.g., dissolution/re-precipitation features, formation of new

715

secondary vein minerals); which suggests that the low permeability nature of these rocks had

716

been established before this time.

717

719

Acknowledgements

TE D

718

M AN U

SC

RI PT

706

We are grateful to Paul Middlestead, Wendy Abdi and Patricia Wickham for assistance in the G.G. Hatch Stable Isotope Laboratory (uOttawa), and to Douglas Hall and Steven Cogswell

721

for help with EPMA and SEM work at the University of New Brunswick. Shuangquan Zang,

722

E.A. Spencer and Rhea Mitchell are acknowledged for their assistance with Sr-isotope work in

723

the Isotope Geochemistry and Geochronology Research Centre (Carleton). Rita Caldas kindly

724

assisted with initial microthermometry. We gratefully acknowledge Ian Clark and Josué Jautzy

725

(uOttawa) for discussions on fluid migration models in the Michigan Basin. We are also grateful

726

to Don Davis and Chelsea Sutcliffe (U of T) for thoughtful discussions on the absolute timing of

727

calcite vein emplacement in the Ordovician carbonates. An anonymous reviewer and Laura

728

Kennell-Morrison and Richard Crowe (NWMO) are thanked for comments and suggestions on

AC C

EP

720

33

ACCEPTED MANUSCRIPT

729

an earlier version of the manuscript. Financial support for this project was provided by the

730

NWMO and NSERC, through an NSERC Industrial Postgraduate Scholarship to JS and an

731

NSERC/NWMO Collaborative Research and Development grant (477852-14) held by TA.

RI PT

732 733 Tables

735

Table 1. Microthermometric results for fluid inclusions in the secondary minerals.

SC

734

736

M AN U

737 Figure Captions

739

Figure 1. Bedrock geology, stratigraphy and hydrostratigraphy of the Michigan Basin below the

740

Bruce site, Southwestern Ontario, Canada. Also shown here are regional structural features

741

(Algonquin Arch, Chatham Sag) and contours lines for the base of the Paleozoic sediments (i.e.

742

elevation in metres to the top of the Precambrian basement rocks). The geologic map is based on

743

Ontario Geological Survey (1991) and Armstrong and Carter (2010).

EP

744

TE D

738

Figure 2. Photomicrographs of secondary minerals in the Trenton Group (A–B), Black River

746

Group (C–F) and the Cambrian units (G–I). (A) DGR-5-704.44, sharp-edged calcite vein in an

747

argillaceous limestone, with an infill of anhydrite that post-dates calcite formation (SL); (B)

748

DGR-2-691.70, irregular calcite vein in a fossiliferous limestone with zones of fine-grained

749

replacement dolomite (SL); (C) DGR-2-778.44, dolostone with zones of tan saddle dolomite and

750

discrete veins of colourless saddle dolomite (SL); (D) DGR-2-817.41, micritic limestone with a

751

saddle dolomite vug with infill anhydrite. Note the fine-grained dolomite lining the interior of

AC C

745

34

ACCEPTED MANUSCRIPT

vug (PPL); (E) DGR-4-829.53, dolostone with saddle dolomite vug, infill anhydrite and discrete

753

veins of colourless dolomite (SL); (F) DGR-2-843.83, micritic limestone with a sparry calcite

754

vug, infill minerals include rough-edged pyrite partly replaced and followed by crystallites of

755

magnetite suspended in an organic-rich matrix (RL); (G) DGR-4-844.78, irregular vein with

756

saddle dolomite and quartz lining and calcite infill within a fine-grained dolostone (RL); (H)

757

DGR-4-845.20, dolostone with saddle dolomite-lined calcite vug (PPL); (I) DGR-2-847.42,

758

sandstone with a calcite-lined vug. Note the euhedral authigenic quartz rims that mantle detrital

759

quartz grains (XPL). Abbreviations: Anh – anhydrite; Cal – calcite; Dol – dolomite; Mag –

760

magnetite; PPL – plane-polarized light image; Py – pyrite; RL – reflected light image; SDol –

761

saddle dolomite; SL – stereo-light image (combination of both transmitted and reflected light);

762

Qtz – quartz; XPL – cross-polarized light image.

M AN U

SC

RI PT

752

763

Figure 3. Relative timing of secondary mineral formation and trapping of fluid inclusion stages

765

I–VI, based on detailed petrography. The timing of authigenic quartz and pyrite formation

766

relative to fluid stages I–II and III–V, respectively, is uncertain and indicated by the dashed lines.

767

Note: matrix replacement dolomite predates the formation of saddle dolomite.

EP

768

TE D

764

Figure 4. Compositions of the secondary minerals in the DGR samples, including wt.% dolomite

770

relative to calcite (A), which was determined using XRD (Saso, 2013) and MgO (B) and Sr (C)

771

concentrations determined using EPMA. Note, the location of stratabound dolostone layers in

772

the succession are indicated in (B) by block symbols. Boundaries in the sedimentary succession

773

are also shown here (in descending order): Queenston–Collingwood/Cobourg (Trenton Group);

AC C

769

35

ACCEPTED MANUSCRIPT

774

Kirkfield (Trenton Group)–Coboconk (Black River Group); Shadow Lake (Black River Group)–

775

Cambrian; Cambrian–Precambrian.

776 Figure 5. Photomicrographs of fluid inclusions representative of paleofluid stages I–III, all

778

viewed in plane-polarized, transmitted light. (A) Stage I in DGR-4-829.53, saddle dolomite

779

crystal showing curved crystal faces lining the walls of a vug. The core is rich in primary fluid

780

inclusions and is overgrown by a clear, inclusion-free outer rim; (B–C) Details of A showing

781

variable proportions of liquid brine, CH4–CO2–vapour and halite within individual inclusions of

782

the same assemblage, indicating that all three phases were present at mutual saturation during

783

fluid entrapment; (D) Stage II in DGR-3-799.15, calcite crystals in a vug, containing clouds of

784

three-dimensionally distributed primary fluid inclusions. The rims of these crystals contain

785

linear arrays of fluid inclusions that radiate outwards, parallel to the growth direction of the

786

crystal, and are indicative of primary inclusions. All are younger than inclusions in A, because

787

calcite systematically overgrows saddle dolomite (Fig. 2H); (E–F) Details of D showing variable

788

proportions of liquid brine, CH4–vapour and halite within individual inclusions of the same

789

assemblage, indicating that all three phases were present at mutual saturation during fluid

790

entrapment; (G) Stage III in DGR-4-829.53, primary inclusions within fibrous anhydrite. The

791

inclusions are younger than in D, because anhydrite systematically overgrows calcite (Fig. 2A);

792

(H–I) Details of G showing variable proportions of liquid brine, CH4–vapour and halite within

793

individual inclusions of the same assemblage, indicating that all three phases were present at

794

mutual saturation during fluid entrapment.

AC C

EP

TE D

M AN U

SC

RI PT

777

795

36

ACCEPTED MANUSCRIPT

Figure 6. Photomicrographs of fluid inclusions representative of paleofluid stages IV–VI, all

797

viewed in plane-polarized, transmitted light. (A) Stage IV in DGR-2-847.42, planar array of

798

secondary fluid inclusions on a healed fracture in vein calcite; (B–C) Details of A showing

799

uniform proportions of liquid brine and H2O–vapour but variable proportions of halite within

800

individual inclusions of the same assemblage. The inclusions homogenize to liquid, therefore

801

only brine and halite were present (without free vapour) at mutual saturation during fluid

802

entrapment; (D) Stage IV in DGR-2-691.70, planar array of secondary inclusions on a healed

803

fracture in vein calcite, as in A; (E) Detail of D showing the same relations as in (B). This key

804

assemblage homogenizes to liquid at 106 °C; Sample (F) Stage V in DGR-2-844.31, planar

805

array of secondary liquid-only inclusions (metastable stretched water) on a healed fracture in

806

vein calcite. Such assemblages cross-cut stage IV assemblages (not shown); (H) Stages IV and

807

VI in DGR-6-892.99, viewed in simultaneous transmitted and UV epi-illumination.

808

Interpretation of relative timing of secondary fluid inclusion generations based on cross-cutting

809

relations. Traces of two healed fractures are visible, each marked by secondary fluid inclusions.

810

Green: Stage IV aqueous inclusions within a fracture trace that strikes to top-left of image and

811

dips steeply to top-right. Blue: Stage VI blue fluorescent liquid hydrocarbon inclusions with

812

CH4-vapour bubbles in a fracture trace that strikes to top-right and dips shallowly to bottom-

813

right. No inclusion is present at the intersection of the fracture traces (at shallow focus), so the

814

relative ages of the two fractures are ambiguous; (I) Same view as H but at a deeper focus level.

815

A hydrocarbon inclusion is present at the intersection of the fracture traces, indicating that the

816

hydrocarbon inclusions and their host fracture are younger than the aqueous inclusions and their

817

corresponding host fracture.

AC C

EP

TE D

M AN U

SC

RI PT

796

818

37

ACCEPTED MANUSCRIPT

Figure 7. Microthermometric data for assemblages of primary and secondary fluid inclusions in

820

the secondary minerals. For the stage IV inclusions, which have uniform liquid–vapour

821

proportions, each symbol represents a range of estimated temperatures between the minimum

822

Ttrap value (i.e. Th,max) and a pressure corrected Ttrap value, which was calculated for peak burial

823

conditions using an estimated burial depth of 1860 m (see section 5.1). Estimated Ttrap values of

824

30–70 °C are indicated for stage V inclusions. Boundaries in the sedimentary succession are

825

also shown here (in descending order): Queenston–Collingwood/Cobourg (Trenton Group);

826

Kirkfield (Trenton Group)–Coboconk (Black River Group); Shadow Lake (Black River Group)–

827

Cambrian; Cambrian–Precambrian.

M AN U

828

SC

RI PT

819

Figure 8. Salinity estimates for assemblages of primary and secondary fluid inclusions in the

830

secondary minerals, derived from Tm (ice) values for all fluid stages (A–B) shown in a H2O–

831

NaCl–CaCl2 ternary model. Stable phase relations of Oakes et al., (1990) and Steele-MacInnis et

832

al. (2011) are also shown in (A). The metastable cotectic in (A) and (B) was calculated in this

833

study.

EP

834

TE D

829

Figure 9. C-, O- and Sr-isotope results for the secondary minerals in the Ordovician–Cambrian

836

sedimentary succession (A–C, respectively). Average C-isotope compositions of seawater

837

during the Late Ordovician (−0.1 to +1.7‰) and the Cambrian (−0.9 to +0.0‰) are also shown

838

in (A) and are based on marine carbonate rocks from Veizer et al. (1999). Average 87Sr/86Sr

839

values for Late Ordovician seawater (0.70887 to 0.70941) were obtained from Veizer et al.

840

(1999). The 87Sr/86Sr values of Precambrian shield brines from the Sudbury region (0.71037 to

841

0.74009) were reported by McNutt et al. (1984). Boundaries in the sedimentary succession are

AC C

835

38

ACCEPTED MANUSCRIPT

842

also shown here (in descending order): Queenston–Collingwood/Cobourg (Trenton Group);

843

Kirkfield (Trenton Group)–Coboconk (Black River Group); Shadow Lake (Black River Group)–

844

Cambrian; Cambrian–Precambrian.

RI PT

845 846

Figure 10. Estimates of O-isotope compositions of fluids (contours) that precipitated carbonate

848

minerals based on mineral–fluid exchange thermometers of Friedman and O'Neil (1977) for

849

calcite (A) and Horitia (2014) for dolomite (B). The boxed fields represent formation

850

temperature estimates, based largely on fluid inclusion data discussed in Section 5.1, plotted

851

against δ18O values for each calcite/dolomite type. Formation temperatures of limestone matrix

852

samples were estimated at 20–30 °C based on Shields et al. (2003). Cambrian–Shadow Lake

853

calcite samples are plotted at 70–72 °C and are based on Ttrap values for stage II fluid inclusions

854

in these units (see Section 5.2; Table 3S – Supporting Material). Calcite vein and vug samples

855

from the Gull River–Coboconk formations and the Trenton Group are plotted using Ttrap

856

temperatures of 54–61 °C and 65–78 °C, respectively. For saddle and vein dolomite, Ttrap

857

temperatures of 88–128 °C were used. No fluid inclusion data were obtained for dolostone and

858

replacement dolomite samples but fluid estimates at 54–78 °C (based on calcite veins/vugs) and

859

88–128 °C (based on saddle dolomite) are shown by dashed boxes. The O-isotope composition

860

of late-Ordovician seawater is indicated by the shaded bands between −2 and −4‰ (Shields et

861

al., 2003).

AC C

EP

TE D

M AN U

SC

847

862 863

References

864

Al, T.A., Clark, I.D., Kennell, L., Jensen, M., Raven, K.G., 2015. Geochemical evolution and

39

ACCEPTED MANUSCRIPT

865

residence time of porewater in low-permeability rocks of the Michigan Basin, Southwest

866

Ontario. Chem. Geol. 404, 1–17. doi:10.1016/j.chemgeo.2015.03.005

867

Allan, J.R., Wiggins, W.D., 1993. Dolomite reservoirs: geochemical techniques for evaluating origin and distribution. American Association of Petroleum Geologists; Continuing

869

Education Course Note Series, 36.

870

RI PT

868

Altmann, S., Tournassat, C., Goutelard, F., Parneix, J.-C., Gimmi, T., Maes, N., 2012. Diffusiondriven transport in clayrock formations. Appl. Geochem. 27, 463–478.

872

doi:10.1016/j.apgeochem.2011.09.015

875 876 877

M AN U

874

Armstrong, D.K. and Carter, T.R. 2010. The Subsurface Paleozoic Stratigraphy of Southern Ontario. Ontario Geological Survey, Special Volume 7

Arthur, M.A., Cole, D.R., 2014. Unconventional Hydrocarbon Resources: Prospects and Problems. ELEMENTS 10, 257–264. doi:10.2113/gselements.10.4.257 Ayalon, A., Longstaffe, F.J., 1988. Oxygen isotope studies of diagenesis and pore-water

TE D

873

SC

871

evolution in the Western Canada sedimentary basin; evidence from the Upper Cretaceous

879

basal Belly River Sandstone, Alberta. Journal of Sedimentary Research 58, 489–505.

880

doi:10.1306/212F8DCD-2B24-11D7-8648000102C1865D

EP

878

Beauheim, R.L., Roberts, R.M., Avis, J.D., 2014. Hydraulic testing of low-permeability Silurian

882

and Ordovician strata, Michigan Basin, southwestern Ontario. Journal of Hydrology 509,

883

163–178. doi:10.1016/j.jhydrol.2013.11.033

884 885

AC C

881

Benson, S.M., Cole, D.R., 2008. CO2 Sequestration in Deep Sedimentary Formations. ELEMENTS 4, 325–331. doi:10.2113/gselements.4.5.325

886

Bottomley, D.J., Clark, I.D., 2004. Potassium and boron co-depletion in Canadian Shield brines:

887

evidence for diagenetic interactions between marine brines and basin sediments. Chem.

40

ACCEPTED MANUSCRIPT

888

Geol. 203, 225–236. doi:10.1016/j.chemgeo.2003.10.010 Bottomley, D.J., Katz, A., Chan, L.H., Starinsky, A., Douglas, M., Clark, I.D., Raven, K.G.,

890

1999. The origin and evolution of Canadian Shield brines: evaporation or freezing of

891

seawater? New lithium isotope and geochemical evidence from the Slave craton. Chem.

892

Geol. 155, 295–320. doi:10.1016/S0009-2541(98)00166-1

RI PT

889

Clark, I.D., Al, T., Jensen, M., Kennell, L., Mazurek, M., Mohapatra, R., Raven, K.G., 2013.

894

Paleozoic-aged brine and authigenic helium preserved in an Ordovician shale aquiclude.

895

Geology 41, 951–954. doi:10.1130/G34372.1

SC

893

Clark, I.D., Ilin, D., Jackson, R.E., Jensen, M., Kennell, L., Mohammadzadeh, H., Poulain, A.,

897

Xing, Y.P., Raven, K.G., 2015. Paleozoic-aged microbial methane in an Ordovician shale

898

and carbonate aquiclude of the Michigan Basin, southwestern Ontario. Organic

899

Geochemistry 83-84, 118–126. doi:10.1016/j.orggeochem.2015.03.006 Clark, I.D., Mohapatra, R., Mohammadzadeh, H., Kotzer, T., 2010a. Porewater and Gas

TE D

900

M AN U

896

901

Analyses in DGR-1 and DGR-2 Core. NWMO Technical Report TR-07-21. Nuclear Waste

902

Management Organization (NWMO), Toronto, pp. 25. https://www.nwmo.ca/en/Reports Clark, I.D., Liu, I., Mohammadzadeh, H., Zhang, P., Wilk, M., 2010b. Porewater and Gas

EP

903

Analyses in DGR-3 and DGR-4 Core. NWMO Technical Report TR-08-19. Nuclear Waste

905

Management Organization (NWMO), Toronto, pp. 29. https://www.nwmo.ca/en/Reports

906

Coniglio, M., Williams-Jones, A.E., 1992. Diagenesis of Ordovician Carbonates From the North-

AC C

904

907

East Michigan Basin, Manitoulin Island Area, Ontario - Evidence From Petrography, Stable

908

Isotopes and Fluid Inclusions. Sedimentology 39, 813–836. doi:10.1111/j.1365-

909

3091.1992.tb02155.x

910

Coniglio, M., Sherlock, R., Williams Jones, A.E., Middleton, K., Frape, S.K., 1994. Burial and

41

ACCEPTED MANUSCRIPT

Hydrothermal Diagenesis of Ordovician Carbonates from the Michigan Basin, Ontario,

912

Canada, in: Purser, B., Tucker, M., Zenger, D. (Eds.), Dolomites a Volume in Honour of

913

Dolomieu. Special Publications of the International Association of Sedimentologists 21.

914

Blackwell Publishing Ltd., Oxford, UK, pp. 231–254. doi:10.1002/9781444304077.ch14

915

Coplen, T.B., Hopple, J.A., Boehike, J.K., Peiser, H.S., Rieder, S.E., Krouse, H.R., Rosman,

RI PT

911

K.J.R., Ding, T., Vocke, R.D., Jr., Révész, K.M., Lamberty, A., Taylor, P., De Bièvre, P.,

917

2002. Compilation of minimum and maximum isotope ratios of selected elements in

918

naturally occurring terrestrial materials and reagents. U.S. GEOLOGICAL SURVEY Water

919

Resources Investigation Report 01-4222.

M AN U

SC

916

920

Davies, G.R., Smith, L.B., Jr., 2006. Structurally controlled hydrothermal dolomite reservoir

921

facies: An overview. American Association of Petroleum Geologists Bulletin 90, 1641–

922

1690. doi:10.1306/05220605164

Davis, D.W., 2013. Application of U-Pb Geochronology Methods to the Absolute Age

TE D

923

Determination of Secondary Calcite. NWMO Technical Report TR-2013-21. Nuclear Waste

925

Management Organization (NWMO), Toronto, pp. 31. https://www.nwmo.ca/en/Reports

926

Davis, D.W., Sutcliffe, C.N., Smith, P., Zajacz, Z., Thibodeau, A.M., Spalding, J., Schneider, D.,

EP

924

Adams, J., Cruden, A. and Parmenter, A., 2017. Secondary calcite as a useful U-Pb

928

geochronometer: An example from the Paleozoic sedimentary sequence in southern Ontario.

929

GAC-MAC 2017 Joint Annual Meeting, Kingston, ON, Canada, May 14–18. Abstracts

930

Volume.

931

AC C

927

Diamond, L.W., 2003. Systematics of H2O inclusions. In: Samson, I.M., Anderson, A.J.,

932

Marshall, D.D. (Eds.), Short Course 32. Fluid Inclusions: Analysis and Interpretation. Short

933

Course 32. Mineralogical Association of Canada, pp. 55-79.

42

ACCEPTED MANUSCRIPT

934

Diamond, L.W., Aschwanden, L. and Caldas, R., 2015. Fluid inclusion study of core and outcrop samples from the Bruce Nuclear Site, Southern Ontario, Canada. NWMO Technical Report

936

TR-2015-24. Nuclear Waste Management Organization (NWMO), Toronto, pp. 138.

937

https://www.nwmo.ca/en/Reports

938

RI PT

935

Douglas, M., Clark, I.D., Raven, K., Bottomley, D., 2000. Groundwater mixing dynamics at a Canadian Shield mine. Journal of Hydrology 235, 88–103. doi:10.1016/S0022-

940

1694(00)00265-1

941

SC

939

Frape, S.K., Fritz, P., McNutt, R.H., 1984. Water-rock interaction and chemistry of groundwaters from the Canadian Shield. Geochim. Cosmochim. Acta 48, 1617–1627. doi:10.1016/0016-

943

7037(84)90331-4

945 946 947

Friedman, I., O'Neil, J.R., 1977. Compilation of stable isotope fractionation factors of geochemical interest. USGS Professional Paper 440-KK.

Fritz, P., Frape, S.K., 1982. Saline groundwaters in the Canadian Shield — A first overview.

TE D

944

M AN U

942

Chem. Geol. 36, 179–190. doi:10.1016/0009-2541(82)90045-6 Gascoyne, M., 2004. Hydrogeochemistry, groundwater ages and sources of salts in a granitic

949

batholith on the Canadian Shield, southeastern Manitoba. Appl. Geochem. 19, 519–560.

950

doi:10.1016/S0883-2927(03)00155-0 Girard, J.-P., Barnes, D.A., 1995. Illitization and Paleothermal Regimes in the Middle

AC C

951

EP

948

952

Ordovician St. Peter Sandstone, Central Michigan Basin: K-Ar, Oxygen Isotope, and Fluid

953

Inclusion Data. American Association of Petroleum Geologists Bulletin 79, 49–69.

954 955 956

Goldstein, R.H., 2001. Fluid inclusions in sedimentary and diagenetic systems. Lithos 55, 159– 193. doi:10.1016/S0024-4937(00)00044-X Goldstein, R.H., Reynolds, T.J., 1994. Systematics of fluid inclusions in diagenetic minerals.

43

ACCEPTED MANUSCRIPT

957

SEPM (Society for Sedimentary Geology) Short Course 31. Tulsa, U.S.A. Haeri-Ardakani, O., Al-Aasm, I., Coniglio, M., 2013. Fracture mineralization and fluid flow

959

evolution: an example from Ordovician-Devonian carbonates, southwestern Ontario,

960

Canada. Geofluids 13, 1–20. doi:10.1111/gfl.12003

RI PT

958

961

Hanor, J.S., 2001. Reactive transport involving rock-buffered fluids of varying salinity.

962

Geochim. Cosmochim. Acta 65, 3721–3732. doi:10.1016/S0016-7037(01)00703-7

Harper, D.A., Longstaffe, F.J., Wadleigh, M.A., McNutt, R.H., 1995. Secondary K-feldspar at

964

the Precambrian–Paleozoic unconformity, southwestern Ontario. Can. J. Earth Sci. 32,

965

1432–1450. doi:10.1139/e95-116

M AN U

966

SC

963

Hendry, M.J., Solomon, D.K., Person, M., Wassenaar, L.I., Gardner, W.P., Clark, I.D., Mayer, K.U., Kunimaru, T., Nakata, K., Hasegawa, T., 2015. Can argillaceous formations isolate

968

nuclear waste? Insights from isotopic, noble gas, and geochemical profiles. Geofluids 15,

969

381–386. doi:10.1111/gfl.12132

TE D

967

970

Holser, W.T., 1979. Mineralogy of Evaporites, in: Burns, R.G. (Ed.), Marine Minerals. Reviews

971

in Mineralogy and Geochemistry, Vol. 6. Mineralogical Society of America. pp. 295–346. Horita, J., 2014. Oxygen and carbon isotope fractionation in the system dolomite-water-CO2 to

EP

972

elevated temperatures. Geochim. Cosmochim. Acta 129, 111–124.

974

doi:10.1016/j.gca.2013.12.027

975 976 977

AC C

973

Hurley, N.F., Budros, R., 1990. Albion-Scipio and Stoney Point Fields-U.S.A., Michigan Basin. American Association of Petroleum Geologists Special Volumes 23, 1–37. Intera, 2011. Descriptive Geosphere Site Model. NWMO Technical Report TR-2011-24. Nuclear

978

Waste Management Organization (NWMO), Toronto, pp. 426.

979

(http://www.nwmo.ca/dgrgeoscientificsitecharacterization).

44

ACCEPTED MANUSCRIPT

980 981

Khader, O., Novakowski, K., 2014. The study of diffusion dominant solute transport in solid host rock for nuclear waste disposal. Report to: Canadian Nuclear Safety Commission. Knauth, L.P., Beeunas, M.A., 1986. Isotope geochemistry of fluid inclusions in Permian halite

983

with implications for the isotopic history of ocean water and the origin of saline formation

984

waters. Geochim. Cosmochim. Acta 50, 419–433. doi:10.1016/0016-7037(86)90195-X

985

Kurtis Kyser, T., Kerrich, R., 1990. Geochemistry of fluids in tectonically active crustal regions.

RI PT

982

Short Course on Fluids in Tectonically Active Regimes of the Continental Crust,

987

Mineralogical Association of Canada, Short Course Handbook 18, 133–230. Legall, F.D., Barnes, C.R., Macqueen, R.W., 1981. Thermal Maturation, Burial History and

M AN U

988

SC

986

989

Hotspot Development, Paleozoic Strata of Southern Ontario - Quebec, from Conodont and

990

Acritarch Colour Alteration Studies. Bulletin of Canadian Petroleum Geology 29, 492–539. Ma, L., Castro, M.C., Hall, C.M., 2009. Atmospheric noble gas signatures in deep Michigan

992

Basin brines as indicators of a past thermal event. Earth Planet. Sci. Lett. 277, 137–147.

993

doi:10.1016/j.epsl.2008.10.015

994

TE D

991

Mazurek, M., de Haller, A., 2017. Pore-water evolution and solute-transport mechanisms in Opalinus Clay at Mont Terri and Mont Russelin (Canton Jura, Switzerland). Swiss J Geosci

996

110, 129–149. doi:10.1007/s00015-016-0249-9 McNutt, R.H., Frape, S.K., Fritz, P., 1984. Strontium isotopic composition of some brines from

AC C

997

EP

995

998

the Precambrian Shield of Canada. Chem. Geol. 46, 205–215. doi:10.1016/0009-

999

2541(84)90190-6

1000

Middleton, K., Coniglio, M., Frape, S.K., 1993. Dolomitization of Middle Ordovician Carbonate

1001

Reservoirs, Southwestern Ontario. Bulletin of Canadian Petroleum Geology 41, 150–163.

1002

Neuzil, C. E., and A. M. Provost. 2014. Ice sheet load cycling and fluid underpressures in the

45

ACCEPTED MANUSCRIPT

1003

Eastern Michigan Basin, Ontario, Canada, J. Geophys. Res. Solid Earth, 119, 8748–8769,

1004

doi:10.1002/ 2014JB011643.

1006 1007

Neuzil, C.E., 2015. Interpreting fluid pressure anomalies in shallow intraplate argillaceous formations. Geophys. Res. Lett. 42, 4801–4808. doi:10.1002/2015GL064140

RI PT

1005

Normani, S.D., Sykes, J.F., 2012. Paleohydrogeologic simulations of Laurentide ice‐sheet history on groundwater at the eastern flank of the Michigan Basin. Geofluids 12, 97–122.

1009

doi:10.1111/j.1468-8123.2012.00362.x

1011 1012

NWMO, 2011. Geosynthesis. NWMO Technical Report TR-2011-11. Nuclear Waste Management Organization (NWMO), Toronto, pp. 418. (https://www.nwmo.ca/en/Reports).

M AN U

1010

SC

1008

Oakes, C.S., Bodnar, R.J., Simonson, J.M., 1990. The system NaCl-CaCl2-H2O: I. The ice

1013

liquidus at 1 atm total pressure. Geochim. Cosmochim. Acta 54, 603–610.

1014

doi:10.1016/0016-7037(90)90356-P

Obermajer, M., Fowler, M.G., Snowdon, L.R., 1999. Depositional environment and oil

TE D

1015

generation in Ordovician source rocks from southwestern Ontario, Canada: Organic

1017

geochemical and petrological approach. American Association of Petroleum Geologists

1018

Bulletin 83, 1426–1453.

1020 1021 1022 1023 1024 1025

Ontario Geological Survey, 1991. Bedrock Geology of Ontario, Southern Sheet. Ontario Geological Survey, Map 2544, Scale 1:1,000,000.

AC C

1019

EP

1016

Pearson, F.J., Jr, 1987. Models of mineral controls on the composition of saline groundwaters of the Canadian Shield. Geological Association of Canada - Special Paper 33, 39–51. Powell, T.G., Macqueen, R.W., Barker, J.F., Bree, D.G., 1984. Geochemical character and origin of Ontario oils. Bulletin of Canadian Petroleum Geology 32, 289–312. Roedder, E., 1984. Fluid Inclusions, in: Ribbe, P.E. (Ed.), Reviews in Mineralogy and

46

ACCEPTED MANUSCRIPT

1026

Geochemistry, Vol. 12. Mineralogical Society of America. Reviews in Mineralogy and

1027

Geochemistry, Mineral Soc Am.

1028

Roedder, E., 1979. Fluid inclusion evidence on the environments of sedimentary diagenesis, a review. The Society of Economic Paleontologists and Mineralogists, Special Publication 89–

1030

107.

1033 1034 1035 1036

Annu. Rev. Earth Planet. Sci. 8, 263–301.

SC

1032

Roedder, E., Bodnar, R.J., 1980. Geologic Pressure Determinations from Fluid Inclusion Studies.

Russell, D.J., Gale, J.E., 1982. Radioactive waste disposal in the sedimentary rocks of southern Ontario. Geoscience Canada 9, 200–207.

M AN U

1031

RI PT

1029

Samson, I.M., Anderson, A.J., Marshall, D.D., 2003. Fluid inclusions: analysis and interpretation. Mineralogical Association of Canada, Short Course 32. Saso, J., 2013. Paleohydrogeological Investigation of Upper Ordovician and Cambrian Rocks

1038

from the Michigan Basin in Southwestern Ontario. University of New Brunswick. M.Sc.

1039

Thesis.

1040

TE D

1037

Shafeen, A., Croiset, E., Douglas, P. and Chatzis, I. 2004. CO2 sequestration in Ontario, Canada. Part I: Storage evaluation of potential reservoirs. Energy Conversion and Management 45,

1042

2645–2659.

EP

1041

Shields, G.A., Carden, G.A.F., Veizer, J., Meidla, T., Rong, J.-Y., Li, R.-Y., 2003. Sr, C, and O

1044

isotope geochemistry of Ordovician brachiopods: a major isotopic event around the Middle-

1045

Late Ordovician transition. Geochim. Cosmochim. Acta 67, 2005–2025. doi:10.1016/S0016-

1046

7037(02)01116-X

1047 1048

AC C

1043

Spencer, R.J., 1987. Origin of CaCl brines in Devonian formations, western Canada sedimentary basin. Appl. Geochem. 2, 373–384. doi:10.1016/0883-2927(87)90022-9

47

ACCEPTED MANUSCRIPT

1049

Steele-MacInnis, M., Bodnar, R.J., Naden, J., 2011. Numerical model to determine the composition of H2O–NaCl–CaCl2 fluid inclusions based on microthermometric and

1051

microanalytical data. Geochim. Cosmochim. Acta 75, 21–40. doi:10.1016/j.gca.2010.10.002

1052

Taylor, T.R., Sibley, D.F., 1986. Petrographic and geochemical characteristics of dolomite types

RI PT

1050

1053

and the origin of ferroan dolomite in the Trenton Formation, Ordovician, Michigan Basin,

1054

U.S.A. Sedimentology 33, 61–86. doi:10.1111/j.1365-3091.1986.tb00745.x

Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., 1999. 87Sr/86Sr, δ13C and δ18O

SC

1055

evolution of Phanerozoic seawater. Chem. Geol. doi:10.1016/S0009-2541(99)00081-9

1057

Wickstrom, L.H., Gray, J.D., Stieglitz, R.D., 1992. Stratigraphy, structure, and production

M AN U

1056

1058

history of the Trenton Limestone (Ordovician) and adjacent strata in northwestern Ohio.

1059

Rep. of investigations/State of Ohio. Dep. of natural resources, Div. of geol. survey.

1060

doi:10.1111/j.1502-3931.1996.tb01833.x/full

Wilson, T.P., Long, D.T., 1993a. Geochemistry and isotope chemistry of CaNaCl brines in

TE D

1061 1062

Silurian strata, Michigan Basin, U.S.A. Appl. Geochem. 8, 507–524. doi:10.1016/0883-

1063

2927(93)90079-V

Wilson, T.P., Long, D.T., 1993b. Geochemistry and isotope chemistry of Michigan Basin brines:

1065

Devonian formations. Appl. Geochem. 8, 81–100. doi:10.1016/0883-2927(93)90058-O Ziegler, K., Longstaffe, F.J., 2000. Multiple Episodes of Clay Alteration at the

AC C

1066

EP

1064

1067

Precambrian/Paleozoic Unconformity, Appalachian Basin: Isotopic Evidence for Long-

1068

Distance and Local Fluid Migrations. Clays and Clay Minerals 48, 474–493.

1069

doi:10.1021/ac60068a025

1070

48

ACCEPTED MANUSCRIPT

Table 1. Microthermometric results for fluid inclusions in the secondary minerals Sample ID

Formation

mBGS

Matrix

T m(ice) or T m(hydrohalite)d (ºC)

T trap (ºC)

Mineral I

II

IVb

III

V

VI

I

[77–96], [106–127]

<70

[57–77] [80–100] [61–78]

<70 <70 <70

48

II III IV min max min max min max –46.9 –40.9 –45.2 –44.7 –21.5 –18.2 –36.4 –33.2 –40.7 –39.0 –45.0 –34.0 –47.0 –42.0 –44.2 –35.9 –46.5 –35.2 –46.8 –42.1

[86–107] [94–113]

<70

59

–22.9 –18.4 –43.1 –40.2

Sherman Fall Coboconk Coboconk Coboconk Coboconk Coboconk Coboconk Gull River Gull River

691.70 764.65 767.81 769.57 778.29 778.44 799.15 818.36 829.53

cal-py dol dol cal-py dol

cal cal cal cal-cls cal

dol-py cal dol-py

dol (SD) cal cal

122a

dol-py

dol (SD)

125, 128a

65, 78 61 57 54

71

42

SC

DGR-2-691.70 DGR-6-883.75 DGR-6-886.91 DGR-4-769.57 DGR-6-892.99 DGR-2-778.44 DGR-3-799.15 DGR-4-818.36 DGR-4-829.53

max

RI PT

min

Vc min

max

–38.8 –44.1

–44.1 –42.7

M AN U

DGR-2-844.31 Cambrian 844.31 55 <70 dol-py dol-cal [75–96], [77–97] –42.7 –41.5 –40.3 –37.4 –43.9 –42.6 –35.0 –34.4 88a qtz cal [91–112], [94–115] –42.9 –41.8 –52.1 –50.5 DGR-2-847.42 Cambrian 847.42 80 qtz cal [88–109], [91–112] DGR-2-850.01 Cambrian 850.01 72 <70 –42.6 –41.5 –52.1 –49.5 –34.2 –32.1 qtz cal –47.9 –42.6 DGR-2-853.72 Cambrian 853.72 70 [102–123], [104–125] <70 –52.7 –50.3 –25.9 –24.3 Notes: Microthermometric measurements were made in calcite and dolomite (a), and are indicated by the Mineral column. Relative timing of fluid stages I–VI was determined petrographically based cross-cutting relationships between fluid inclusion assemblages. (b) T trap for stage IV can only be bracketed between T h (minimum possible value) and a maximum T calculated from an assumed lithostatic pressure of 1860 m (see text for explanation).

AC C

EP

TE D

T m(ice) values in stages I–IV represent the metastable equilibrium between halite+ice+liquid+vapour. (c) T m(ice) for stage V was obtained by artificially stretching the inclusions. (d) Hydrohalite T m values are shown in italics. Abbreviations: cal – calcite; cls – celestine; dol – dolomite; py – pyrite; SD – saddle dolomite; qtz – quartz.

ACCEPTED MANUSCRIPT

Stratigraphy Pleistocene Lucas Amherstburg

Karstic Aquifer

Bois Blanc

RI PT

100

Devonian

0

Hydrostratigraphy

Bass Islands

Silurian

Silurian Aquitard A1 Aquifer

Guelph Goat Island Gasport/Lions Head/ Fossil Hill Cabot Head Manitoulin Queenston

400

500

Georgian Bay

600 Blue Mountain Collingwood Cobourg Sherman Fall Kirkfield Coboconk

800

Gull River Shadow Lake Cambrian

900

Precambrian

Trenton Group

700

Guelph Aquifer

Ordovician Aquiclude

Black River Group

Depth (mBGS)

Salina

300

Ordovician

AC C

EP

TE D

M AN U

SC

200

Black River Group Aquitard Basal Aquifer

150

Temperature (°C)

125 100 75 50

–2

0

+4

+8

–4 –6

Trenton Group

ACCEPTED MANUSCRIPT

Cambrian–Shadow Lake

–10

–14

RI PT

A

Gull River– Coboconk

SC

25

M AN U

All Limestones

0 10.0

14.0

18.0

22.0

26.0

30.0

B

150

–4

–2

TE D

δ 18OVSMOW(‰) of calcite 0

+4

+8

100

–6

AC C

Temperature (°C)

125

EP

Saddle Dolomite

–10

Matrix Replacement and Dolostone

75 –14

50 25 0 10.0

14.0

18.0

22.0

26.0

δ 18OVSMOW(‰) of dolomite

30.0

A

ACCEPTED MANUSCRIPT

B

C

Dol

RDol Anh

Anh

Cal vein

RI PT

Dol vein SDol

Cal vein

Anh

2 mm

E

SDol

Anh

SDol

1 mm

Py

Organics+Mag

Cal

Dol vein

Cal Dol 500 µm

EP

SDol

AC C

H

Dol

Qtz

F

Dol

1 mm

G

TE D

Dol

4-829.53

SDol

reimage SDol

SC

D

2 mm

M AN U

2 mm

Dol

Cal vug

I

SDol

Cal vug

SDol SDol

Cal vein Dol

Dol 500 µm

SDol

1 mm

1 mm

Qtz

Fluid Stages

I

II

III

IV

ACCEPTED MANUSCRIPT

V

VI

Secondary Minerals Dolomite - saddle and veins

RI PT

Authigenic Quartz Calcite

SC

- veins and vugs

Anhydrite

M AN U

Celestine Pyrite Magnetite

EP

Fluids

TE D

Bitumen

AC C

Halite sat. brine I + CH4

Halite sat. brine II + CH4 Halite sat. brine III + CH4 Halite sat. brine IV Brine V Oil + CH4 early

late

650

A

B

C

ACCEPTED MANUSCRIPT

650

650

Calcite Veins and Vugs

RI PT

700 700

Dolomite Dolostone Matrix Matrix Replacement

M AN U

Anhydrite in Veins and Vugs

750

TE D

Depth (mBGS)

SC

Saddle and Veins

AC C

EP

800 800

850 0

20

40

60

80

Dolomite wt.% (Relative to Calcite)

100

0

4

8

12

MgO (wt.%)

16

20

24

0

1000

2000

3000

Sr (ppm)

4000

5000

ACCEPTED MANUSCRIPT

B

A

C

SDol

B

SDol

RI PT

C

Stage I

SC

D

M AN U

E

h) H

Anh

EP

F

Cal

Stage II

Stage II

H

I

Anh

Stage III

Stage I

Cal

AC C

G

TE D

Cal

Stage II

SDol

Stage I

Anh

Stage III

Stage III

A

ACCEPTED MANUSCRIPT

B

C

Cal

RI PT

B

Cal Stage IV

SC

E

Cal

M AN U

D

Cal Stage IV

Stage IV

F

Cal

E

Cal G

H

Cal

Cal

I

AC C

G

Cal Cal

Stage V

Stage V

Stage IV

EP

Stage IV

TE D

Uniform vapour fractions

ACCEPTED MANUSCRIPT Primary inclusions

650

Stage I (dolomite) Stage II (calcite) Stage III (cls+anh)

Stage IV Stage V Stage VI (oil+gas)

M AN U AC C

800 800

TE D

750

EP

Depth (mBGS)

SC

RI PT

700 700

Secondary inclusions

850 25

50

75

100

Temperature (Ttrap, °C)

125

A Ca Cl 2

H 2O

Ma ss%

10

10 Ice + L + V

te ) i s

Metastable cotectic Ice + Halite + L + V 30

otherms

SC

h a li T m(

Hydrohalite + L + V

RI PT

therms

°C 42 °C 70 °C 6 10 8 °C 12

40

Halite + L + V

M AN U

40

20

Tm(ice) iso

30 Eutectic -52 °C Peritectic -22.4 °C

Cl Na

20

-24 °C -26 °C -32 °C -35 °C

% ss Ma

ACCEPTED MANUSCRIPT

Antarcticite + L + V

CaCl2

TE D

B

NaCl

Ca

ss%

Ma

Ice + L + V

20

Cl

Na

AC C

10

%

10

ss

Ma

EP

Cl

2

H 2O

20

Stage I Stage II Stage III Stage IV Stage V

? ?

30

40

CaCl2

30

Halite + L + V

40

NaCl

650

A

B

C

ACCEPTED MANUSCRIPT

650

Calcite Limestone Matrix Veins and Vugs Dolomite Dolostone Matrix Matrix Replacement

RI PT

700

Saddle and Veins Anhydrite and Celestine in Veins and Vugs

M AN U

Late Ordovician Seawater

750

TE D

Depth (mBGS)

SC

700

AC C

EP

800 800

Precambrian Shield Brines

Late Ordovician Seawater

850

Cambrian Seawater

Cambrian Seawater

-8.0

-6.0

-4.0

-2.0

0.0

δ 13CVPDB(‰)

2.0

4.0

10.0

14.0

18.0

22.0

δ 18OVSMOW(‰)

26.0

30.0 0.707

0.708

0.709 87

Sr/ 86 Sr

0.710

0.711

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Highlights 1. Multi-stage paleofluid history recorded by primary and secondary fluid inclusions 2. Primary stages I to III formed the dolomite and calcite veins 3. Secondary fluid stages IV to VI record trapping of local and re-mobilized fluids 4. Mineral isotope data indicate variable fluid sources for the veins 5. Sources include connate fluids and mixed connate–shield-type brines