Mineral Reactions in Shale Gas Reservoirs: Barite Scale Formation

Jul 19, 2017 - Abstract: Basalt formations could enable secure long-term carbon storage by ... Abstract: For successful CO2 storage in underground res...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF NEWCASTLE

Article

Mineral Reactions in Shale Gas Reservoirs: Barite scale formation from reusing produced water as hydraulic fracturing fluid Amelia N. Paukert Vankeuren, J. Alexandra Hakala, Karl Jarvis, and Johnathan E. Moore Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b01979 • Publication Date (Web): 19 Jul 2017 Downloaded from http://pubs.acs.org on July 27, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 35

Environmental Science & Technology

1

Mineral reactions in shale gas reservoirs: barite scale formation from reusing produced water as

2

hydraulic fracturing fluid

3

Amelia N. Paukert Vankeuren* a,b, J. Alexandra Hakala b, Karl Jarvis c,d, Johnathan E. Moore c,d

4 5

a

Geology Department, California State University Sacramento, Sacramento, CA 95819

6

b

National Energy Technology Laboratory, U.S. Department of Energy, Pittsburgh, PA 15236

7

c

National Energy Technology Laboratory, U.S. Department of Energy, Morgantown, WV 26507

8

d

AECOM, Morgantown, WV 26507

9

*Corresponding author contact information: Mailing address: California State University,

10

Sacramento; 6000 J Street, PLR 1016; Sacramento, CA 95819-6043; Phone: (916) 278-7385;

11

Fax: (916) 278-4650; Email: [email protected]

12 13

Abstract

14

Hydraulic fracturing for gas production is now ubiquitous in shale plays, but relatively little is

15

known about shale-hydraulic fracturing fluid (HFF) reactions within the reservoir. To

16

investigate reactions during the shut-in period of hydraulic fracturing, experiments were

17

conducted flowing different HFFs through fractured Marcellus Shale cores at reservoir

18

temperature and pressure (66oC, 20 MPa) for one week. Results indicate HFFs with

19

hydrochloric acid cause substantial dissolution of carbonate minerals, as expected, increasing

1

ACS Paragon Plus Environment

Environmental Science & Technology

20

effective fracture volume (fracture volume + near-fracture matrix porosity) by 56-65%. HFFs

21

with reused produced water composition cause precipitation of secondary minerals, particularly

22

barite, decreasing effective fracture volume by 1-3%. Barite precipitation occurs despite the

23

presence of antiscalants in experiments with and without shale contact, and is driven in part by

24

addition of dissolved sulfate from the decomposition of persulfate breakers in HFF at reservoir

25

conditions. The overall effect of mineral changes on the reservoir has yet to be quantified, but the

26

significant amount of barite scale formed by HFFs with reused produced water composition

27

could reduce effective fracture volume. Further study is required to extrapolate experimental

28

results to reservoir-scale, and to explore the effect that mineral changes from HFF interaction

29

with shale might have on gas production.

30

1. Introduction

31

Over the past decade, technological innovations in horizontal drilling and hydraulic fracturing

32

have spurred a shale gas boom and made the US the largest natural gas producer globally.1 The

33

Marcellus Shale, which underlies parts of Pennsylvania, New York, West Virginia, and Ohio, is

34

a key contributor to US natural gas production.2 In Pennsylvania alone, over 9,500

35

unconventional wells have been drilled in the past 10 years and annual natural gas production

36

has reached over 4.6 trillion cubic feet.3

37

The dramatic increase in gas production has led to a commensurate increase in produced

38

water – wastewater that is produced at the wellhead as a result of oil and gas production, either

39

through flowback of injected hydraulic fracturing fluid (HFF) or formation water from the

40

reservoir. There are now over 1.4 billion gallons of produced water generated in Pennsylvania

41

each year.4 This water contains high total dissolved solids (TDS) and naturally occurring

42

radioactive materials, making it unsuitable for disposal at wastewater treatment plants.5 Other

2

ACS Paragon Plus Environment

Page 2 of 35

Page 3 of 35

Environmental Science & Technology

43

states, e.g., Texas, dispose of the vast majority of their produced water through deep subsurface

44

injection wells, but Pennsylvania has only 8 active injection wells for the disposal of oil and gas

45

produced water.6-8 Some produced water from Pennsylvania is trucked to disposal wells in West

46

Virginia and Ohio, but that carries high costs and additional risks associated with the transport of

47

large volumes of wastewater.9-10 Due to the limited options for disposal of produced water in

48

Pennsylvania, most produced water is now diluted with freshwater and reused as HFF.9, 11-13

49

Reusing produced water as HFF introduces another complication – the high concentration of

50

alkali earth metals (calcium, barium, strontium) in produced water could lead to scale

51

formation.14-17 Precipitation of scale minerals could result in local reductions in reservoir

52

porosity and fracture aperture, preventing gas from reaching the wellbore.16, 18 The potential for

53

scale minerals to damage well productivity is well-recognized by the industry, and antiscalants

54

are usually included in the HFF chemical mixture.19 However, recent experiments have shown

55

antiscalants are ineffective at preventing precipitation of common scaling minerals (e.g., barite,

56

calcite) under supersaturated conditions.20-22 Given the high concentrations of alkali earth metals

57

in produced water,13, 23 mixing produced water with freshwater and HFF chemicals that generate

58

sulfate may cause supersaturation and precipitation of scaling minerals. Interaction between HFF

59

and shale minerals could also change the fluid composition (e.g., increasing dissolved sulfate

60

through pyrite oxidation) and lead to mineral scaling.

61

The most likely timeframe for scale mineral formation related to HFF is during injection–

62

when the fluid temperature rises from ambient to reservoir temperature– and the shut-in

63

(soaking) period– the days to month timeframe after hydraulic fracturing but before well

64

production when HFF sits in contact with the shale with little to no flow. What happens to the

65

fluid during the shut-in period is not well understood, and the relationship between well

3

ACS Paragon Plus Environment

Environmental Science & Technology

66

productivity and length of the shut-in period is debated.24-25 While fluid imbibition into shale

67

during the shut-in period has been examined as a mechanism for reduced well productivity,26

68

geochemical processes such as mineral reactions within the shale reservoir could also impact gas

69

production.

70

Though common mineral reactions have been well characterized in the lab, these

71

experiments are typically done in isolation with an emphasis on a particular reaction (e.g., calcite

72

dissolution or barite precipitation). Interactions between the complex mixture of solutes and

73

chemical additives present in HFF may alter mineral behavior and yield results that do not match

74

those predicted from laboratory studies. Thus it is important to conduct experiments in the lab

75

that use the complex mixtures employed in the field.

76

Prior experiments suggest that mineral reactions due to shale-fluid interaction could be

77

significant. Studies evaluating reactions between water with and without HFF chemicals and

78

various shales at ambient pressure and temperature have shown evidence of dissolution of

79

carbonate minerals and iron and manganese oxides, and precipitation of sulfate minerals.27-29 At

80

elevated temperature (80oC), under oxidizing conditions such as those created by the exposure to

81

the atmosphere and/or inclusion of an oxidant breaker in the HFF chemicals (typically

82

ammonium persulfate)30 these experiments showed evidence of pyrite oxidation, and

83

precipitation of iron oxides and sometimes barite.27, 29, 31 High temperature, high pressure rocking

84

autoclave experiments reacting Marcellus Shale with synthetic shale formation brine mixed with

85

HFF chemicals or synthetic diluted produced water also exhibited carbonate mineral dissolution

86

and minor anhydrite or gypsum precipitation, and predicted precipitation of barite and nontronite

87

clay.32-33 The formation of barite scale in relation to shale gas wells has been investigated in

4

ACS Paragon Plus Environment

Page 4 of 35

Page 5 of 35

Environmental Science & Technology

88

experiments performed at ambient pressure and temperature, including interactions with

89

antiscalants and using abandoned mine drainage to remove barium from HFF.34-35

90

All prior experiments have utilized shale powder and/or rock chips in benchtop or batch

91

reactor settings, precluding evaluation of spatial differences along a flow path and resulting in a

92

fluid:rock ratio much higher than expected in shale reservoirs. Most experiments were conducted

93

at ambient temperature and pressure, far from the conditions found in shale reservoirs. While

94

these experiments provide valuable preliminary data on mineral reactions in shale reservoirs,

95

there is a need for fractured core flow experiments at reservoir temperature and pressure to more

96

closely approximate conditions in the field.

97

To explore mineral reactions within the shale reservoir during the shut-in period and related

98

effects on gas production, fractured core flooding experiments were conducted exposing shale to

99

different compositions of HFF at reservoir temperature and pressure conditions. This study aims

100

to investigate:

101



What mineral reactions result from HFF injection into a shale reservoir?

102



What mineral reactions result from varying HFF composition?

103



What effect do these reactions have on fracture volume, and how might that affect well

104

productivity?

105

2. Experimental method

106

Experiments were conducted exposing Marcellus Shale cores to synthetic HFFs of varying

107

compositions for 7 days. Changes to fluid chemistry and shale fracture surfaces were used to

108

determine dissolution and secondary mineral precipitation reactions resulting from fluid-rock

109

interaction. Experiments were designed to closely approximate conditions in the field during

110

hydraulic fracturing. Pressure and temperature were set to those expected at a depth of 6,600 ft,

5

ACS Paragon Plus Environment

Environmental Science & Technology

111

where the Marcellus Shale occurs in southwestern Pennsylvania (Table 1). Oxygen was limited

112

in fluids by degassing fluids in a vacuum chamber and then flushing the headspace above HFF

113

with nitrogen gas to prevent addition of dissolved oxygen, but oxygen was not entirely removed

114

from the fluid prior to injection because the vast majority of operators do not incorporate oxygen

115

scavengers in their HFF.19 Because these experiments are intended to simulate the shut-in period,

116

the flow rate was set as slow as possible while still generating sufficient effluent for analyses.

117

The residence time of fluid in these experiments is approximately 100 minutes. While this is

118

likely shorter than would be expected during the shut-in period, when neither injection nor

119

production is occurring, it is a reasonable approximation of prolonged contact between shale and

120

HFF. For comparison, during hydraulic fracturing our industry collaborator reported average

121

fluid injection rates of 12,000-17,500 liters/min, and residence time is estimated at 0.5-0.8

122

minutes.36

123

Table 1: Experimental parameters Temperature 65.5oC Pore pressure

20 MPa

Confining pressure

21.4 MPa

Headspace

High purity N2 gas

Fluids tested

1) SW: local spring water (Table S1) 2) SWF: local spring water with HFF chemicals added (Table S2) 2) PW: synthetic reused produced water (Table S3) 3) PWF: synthetic reused produced water with HFF chemicals added 4) PWFNA: synthetic reused produced water with HFF chemicals added, excluding hydrochloric acid (HCl)

6

ACS Paragon Plus Environment

Page 6 of 35

Page 7 of 35

Environmental Science & Technology

Fluid flow rate

0.04 ml/min

Dimensions of material

Length 152.4 mm

in core holder

Diameter 38.1 mm Shale experiments

Control experiments

Duration

7 days

2 days

Fluid sampling interval

2 days, 7 days

2 days

Material in core holder a

Marcellus shale: 46.5% clay (32.5% illite, 7%

316 stainless

illite-smectite, 7% chlorite), 23.3% calcite,

steel cylindrical

18.6% quartz, 4.6% pyrite, 7% organic content

spacers

174 x 103 mm3

N/A

Shale volume

Average fracture volume 3927 mm3

124

Estimated fracture

>210 µm based on minimum proppant grain size, N/A

aperture b

500-1,500 µm based on x-ray CT imaging

Proppant composition c

40/70 mesh northern white sand (>99% SiO2)

a

b

127

Average shale mineralogical composition in percent mineral content from 37, average shale

It was not possible to measure fracture aperture at experimental pressure. Aperture is estimated from x-ray CT images taken at ambient pressure.

128

c

129

2.1 Fluid compositions

130 131

N/A

organic content from 38

125 126

N/A

Unimin Energy Solutions, The Woodlands, TX

All five experimental fluids (Table 1) used local spring water as the base fluid to ensure a common starting point and to match the fresh water used in industrial hydraulic fracturing

7

ACS Paragon Plus Environment

Environmental Science & Technology

132

operations. Synthetic reused produced water was generated by adding salts to spring water to

133

simulate the composition of HFF using diluted reused produced water from a well drilled in the

134

Marcellus Shale in Greene County, PA that was provided by an industry collaborator (Table

135

S3).39 The compositions of HFF from other wells targeting the same reservoir have been reported

136

and discussed in a previous publication.40 HFF chemicals were chosen to represent an average

137

HFF composition used in western Pennsylvania and eastern Ohio, collated from numerous

138

reports found on FracFocus.org.41-43 The selected HFF chemicals included a gelling agent, scale

139

inhibitor, corrosion inhibitor, clay stabilizer, iron control, friction reducer, cross linker, breaker,

140

surfactant, pH adjusters, and biocide. Two fluids also included HCl because HCl is usually

141

injected into a well prior to hydraulic fracturing to clean perforations and allow HFF access to

142

the shale (a step known as the acid pack). Because acid injection precedes injection of HFF,

143

PWFNA may be most representative of conditions within the shale reservoir at the time of

144

hydraulic fracturing. However, the acid neutralizing capacity of the shale will depend on the

145

amount of carbonate minerals accessible to dissolve and buffer the pH, will vary by shale

146

formation and by location within the formation. While it is possible that all acid is spent prior to

147

fracturing, it is also possible that some acid remains. Operators generally do not take samples of

148

fluid from the borehole after acid injection but prior to fracturing, so there are no data to

149

constrain the pH of the fluid in the well prior to fracturing. Due to this uncertainty, experiments

150

were run for both end-member cases: with HCl, as if no acid were spent resulting in pH ~2, and

151

without HCl, as if all acid were spent, resulting in pH ~8. The complicated mixture in HFF

152

contains many compounds that could– and in some cases are intended to– affect mineral

153

reactions, making it exceedingly difficult to predict mineral behavior within the reservoir.

154

2.2 Experimental set up

8

ACS Paragon Plus Environment

Page 8 of 35

Page 9 of 35

Environmental Science & Technology

155

A system was constructed to simulate HFF injection into a shale reservoir and subsequent

156

reaction between fluids and shale minerals (Table 1, Fig S1).

157

For each fluid, a control experiment was conducted in which the fluid flowed through a core

158

holder containing stainless steel spacers rather than shale at experimental temperature and

159

pressure. These control experiments allowed the differentiation between reactions that result

160

from heating and pressurizing the HFF and reactions that result from fluid-shale interaction. For

161

shale experiments, effluent was collected in the syringe pump reservoirs and sampled after 2

162

days (2-day samples), representing an average of effluent from 0-2 days, and again after 7 days

163

(7-day samples), representing an average of effluent from 2-7 days. Detailed experimental

164

protocols are provided in the supporting information.

165

2.3 Shale core preparation:

166

Cores were cut from the interior of large Marcellus Shale samples taken from outcrops away

167

from the weathered surface, artificially fractured using the Brazilian method 44 (Hydrasplit

168

Masonry Stone Splitter CM-10, Park Industries, St. Cloud, MN). Fractures were packed with

169

proppant to maintain a propped fracture at experimental pressure and simulate the nature of the

170

fracture once proppant has been emplaced and fluid pressure is reduced. The mineralogy of the

171

shale exposed along each fracture face varies slightly, but is assumed to be representative of an

172

average Marcellus Shale composition. Scanning electron microscopy with energy dispersive x-

173

ray spectroscopy (SEM/EDS; Quanta 600 FEG, FEI, Hillsboro, OR) supports this assumption

174

(Fig. S1).

175

2.4 Analytical methods:

176

Before and after experiments, SEM/EDS analyses were performed on shale fracture faces, and x-

177

ray computed tomography (CT) with 24 µm pixel resolution (M5000 Industrial Computed

9

ACS Paragon Plus Environment

Environmental Science & Technology

178

Tomography System, North Star Imaging, Rogers, MN) was performed on select core sections:

179

the quarter lengths of core nearest the fluid inlet and outlet (hereafter referred to as the inlet

180

38mm and outlet 38mm, respectively) (Fig S3). Images were processed with ImageJ.45

181

For experiments where CT scans showed significant mineral changes at the conclusion of the

182

experiments, analyses included segmentation of the CT images using ilastik, an interactive

183

segmentation toolkit.46 The pixel classification workflow was used to differentiate between

184

specific components: matrix shale, proppant grains, open fracture, mineral dissolution, and

185

secondary mineral precipitation, where applicable.

186

Fluid chemistry was measured for all fluid samples (see supporting information for detailed

187

analytical methods). The complicated composition of HFFs required special treatment to analyze

188

via instruments using inductively coupled plasma, which is discussed in the supporting

189

information. All but a few samples have charge balances within +4% electroneutrality, and all

190

samples are within +20%.

191

3. Results and discussion

192

The most significant mineral changes observed during experiments were calcite dissolution and

193

barite precipitation. Fluid chemistry data and plots of significant changes are provided in the

194

supporting information (Table S5a and S5b, Fig S2). Fluid chemistry data were speciated and

195

saturation indices calculated (SI = log Q/K, where Q is the ion activity product and K is the

196

equilibrium constant) with PHREEQC v3.2 using the LLNL thermodynamic database at 65.5oC

197

and 0.101 MPa (Table S6).47

198

Volume changes due to mineral reactions were calculated when quantifiable at CT scan

199

resolution (i.e., for the inlet 38 mm of core for experiments with HFF chemicals) (Table 2).

200

Spatial variations in volume changes are depicted in 3-D web enhanced objects highlighting

10

ACS Paragon Plus Environment

Page 10 of 35

Page 11 of 35

Environmental Science & Technology

201

initial fracture volume (yellow), matrix porosity added by dissolution (green), effective fracture

202

volume lost by secondary mineral precipitation (white), and quartz proppant (red) (WEO 1-3).

203

Effective fracture volume (defined for this paper as fracture volume + near-fracture matrix

204

porosity) changes are estimates; cores were depressurized and moved to the CT scanner, so shale

205

cores may have shifted slightly in the core sleeves, preventing the direct comparison of core

206

locations before and after the experiments. Additionally, initial fracture volumes are maximum

207

values because CT scans were performed at 0.101 MPa, while experiments were conducted at 20

208

MPa pore pressure, 21.4 MPa confining pressure. Mineral changes in the outlet 38 mm were

209

below pixel resolution and could not be quantified. All x-ray CT images are provided (Fig S3

210

and S4) but only results from the inlet 38 mm of core are discussed.

211

3.1 Experiments with natural spring water

212

Fluid chemistry

213

Both SW and SWF experiments show an increase in pH, Ca2+, and dissolved inorganic

214

carbon (DIC) in the effluent with respect to the influent (Table S5, Figure S2), though changes

215

are more pronounced in SWF experiments due to the much lower starting pH. In 2-day samples,

216

the increase in Ca2+ in the SWF experiments is over 12 times the increase in the SW experiments.

217

Dissolution of calcite likely is responsible for both pH buffering and the release of Ca2+ and DIC

218

from the shale into the fluid. In SWF, the increase in DIC is not stoichiometric with Ca2+

219

because DIC in the sample is likely an underestimate. Samples were exposed to the atmosphere

220

during sampling, and given the low pH of the solutions, much of the DIC should have been in the

221

form of carbonic acid and may have been lost by degassing of CO2. Nonetheless, calcite

222

dissolution is supported by saturation indices (Table S6): SW-influent and SWF-influent are

223

highly undersaturated with respect to calcite (SI of -6 and -11, respectively) and after 7 days of

11

ACS Paragon Plus Environment

Environmental Science & Technology

224

contact with shale, saturation indices in the effluents are much closer to calcite equilibrium (SI of

225

-1 and 1, respectively).

226

Page 12 of 35

Sulfate also increased in 2 day-samples with respect to the influents for both SW and

227

SWF experiments, with the increase 2.5 times larger in SWF experiments (Table S5b). With SW,

228

control experiments show little increase in SO42- compared to the experiments contacting shale.

229

This implies that the increase in SO42- when shale reacts with fresh water is likely due to

230

oxidation of pyrite, something observed in prior studies on HFF-shale reactions, even – to a

231

limited extent – under anoxic conditions.27, 29, 32 In SWF experiments, SO42- concentration

232

increases even in the control experiment due to thermal activation of ammonium persulfate in the

233

HFF chemicals to produce sulfate ions.48 In both the SW and SWF experiments, SO42- in 7-day

234

samples decreased slightly with respect to the 2-day samples. In SW experiments, though

235

gypsum remains undersaturated in fluid samples (SI -1.7 to -2.3), the increased Ca2+ from calcite

236

dissolution and SO42- from pyrite oxidation caused local oversaturation and precipitation of

237

gypsum, thus lowering dissolved SO42- between days 2 and 7. In SWF experiments, the larger

238

increase in SO42- from HFF chemicals may have allowed the formation of both barite and

239

gypsum and corresponding reductions in SO42-. Though one would expect the concentration of

240

sulfate to reach equilibrium with respect to barite and then remain at that concentration, there

241

may have been temporary oversaturation due to the rate of SO42- release by persulfate

242

decomposition exceeding the rate of removal by barite precipitation. Saturation indices for barite

243

are at equilibrium or moderately oversaturated throughout the experiments (SI of 0 to 0.4). Barite

244

precipitation dynamics are discussed in more detail following the experimental results.

245 246

In both SW and SWF experiments, small increases (up to 0.15 mmol/L) in Mg2+, Na+, SiO2, and Cl- are also observed. Na+ and Cl- increases may be due to dissolution of halite or

12

ACS Paragon Plus Environment

Page 13 of 35

Environmental Science & Technology

247

saline pore water from the shale, and Mg2+ and SiO2 from dissolution of dolomite and clay,

248

respectively.

249

Fracture surface imaging

250

SW experiments: SEM analysis reveals dissolution of calcium carbonate minerals near the fluid

251

inlet, and minor precipitation of iron oxides and gypsum near the outlet (Fig S3). Gypsum seems

252

to have preferentially precipitated on top of primary pyrite (Fig S3). Also near the outlet, new

253

Mg-Al-rich minerals appear, which could be due to precipitation or deflocculation and transport

254

of clays from closer to the inlet. SW experiments were the only ones showing new Mg-Al-rich

255

minerals; the low salinity of the spring water may be responsible for deflocculation of clays in

256

the shale.49

257

X-ray CT images do not show visible alteration to the shale (Fig S3) because changes in

258

mineral volumes are below the 24 µm pixel resolution. This prevents calculation of effective

259

fracture volume change using ilastik for these experiments.

260

SWF experiments: X-ray CT images from SWF experiments show a darkened reaction rim of

261

calcite dissolution extending along the surface of the main fracture from the fluid inlet to a

262

distance at least 30 mm along the core and penetrating a maximum of 0.8 mm into the shale (Fig

263

1). SEM images of the fracture surface confirm that extensive dissolution of calcite has occurred

264

in this area. When a dense grouping of proppant is present along part of the fracture, the reaction

265

rim is often narrower near the proppant than elsewhere along the fracture (Fig S3). Clusters of

266

proppant are thought to slow the flow of fluid, forcing most of the fluid through the open fracture.

267

Dissolution is faster along preferential pathways due to the greater availability of low pH

268

reactive fluid.50-52 Additionally, x-ray CT images reveal the presence of a secondary material less

269

dense than the shale, SEM images indicate that this is amorphous and may be residue of HFF

13

ACS Paragon Plus Environment

Environmental Science & Technology

270

chemicals, such as guar gum (Fig S5). This residue is not seen in experiments with reused

271

produced water and HFF chemicals, possibly due to decreased stability of guar gum with the

272

higher concentration of Na+. Studies testing the performance of guar based gelling agents in

273

reused produced water showed significantly decreased viscosity at levels of 85 mmol/L Na+; the

274

concentration of Na+ in HFF with reused produced water composition is 3.5 times that value.53

275 276

Figure 1: SWF experiments. A) and B) x-ray CT of inlet 38mm of shale parallel to flow, B) with

277

calcite dissolution (dark gray) along fracture. The decrease in fracture aperture in B is likely due

278

to the core halves shifting in the core sleeve during the experiment, not precipitation or

279

geomechanical alteration. C) and D) SEM of main fracture face near the fluid inlet, D) with

280

calcite dissolution (dark holes)

281

X-ray CT analysis indicates a significant increase in effective fracture volume (WEO 1).

282

In the inlet 38 mm, calcite dissolution in shale adjacent to the fracture creates matrix porosity,

283

increasing effective fracture volume by 65%. The amorphous secondary material thought to be

284

HFF chemical residue fills 4% of the fracture volume, yielding a net increase of 61%. For the

14

ACS Paragon Plus Environment

Page 14 of 35

Page 15 of 35

Environmental Science & Technology

285

outlet 38 mm, there was no measurable change in effective fracture volume on the x-ray CT, so

286

the impact of calcite dissolution on effective fracture volume appears to only apply to areas that

287

are exposed to HFF with residual acid.

288

3.2 Experiments with synthetic reused produced water

289

Fluid chemistry

290

All fluids with a reused produced water base have high TDS, so subtle changes in solute

291

concentration relative to the influents may not be perceptible. In PW and PWF experiments, pH

292

and DIC both increase due to dissolution of carbonate minerals. PW experiments have only

293

minor dissolution, causing a slight increase in DIC and no measurable change in Ca2+. PWF

294

experiments have a much more dramatic increase in pH, Ca2+, and DIC due to the lower starting

295

pH. Again, DIC measurements in these samples are likely underestimates due exposure to the

296

atmosphere during sampling. In both sets of experiments, PW and PWF, influents are

297

undersaturated with respect to calcite (SI of -0.03 and -13, respectively) and saturation index

298

increases in the 7-day samples (SI of 1 and -0.3, respectively) due to calcite dissolution.

299

PWFNA experiments show a decrease in pH and DIC due to precipitation of calcite.

300

PWFNA starts off oversaturated with respect to calcite (SI of 1.9) and draws closer to

301

equilibrium in the 7-day samples (SI of 1.3). Precipitation of calcite in the control experiment

302

may have occurred due to the increased temperature in the core holder lowering calcite solubility.

303

PW at 20oC is undersaturated with respect to calcite (SI of -0.65), while the same fluid at 66oC is

304

oversaturated (SI of 1.9).

305

All three experiments, PW, PWF, and PWFNA show decreases in Ba2+ in control

306

experiments, suggesting some barite precipitation occurs from the fluid even without interaction

307

with the shale. In experiments containing HFF chemicals (PWF and PWFNA) with or without

15

ACS Paragon Plus Environment

Environmental Science & Technology

308

shale, the magnitude of Ba2+ decrease in 2-day samples relative to the influents is 2-8 times the

309

decrease in the experiment without HFF chemicals. The greater decrease in Ba2+ is due to barite

310

precipitation driven by the addition of SO42- from activation of ammonium persulfate in the HFF

311

chemicals. In PW experiments, though Ba2+ decreases, SO42- increases in experiments with shale

312

cores so it is likely that SO42- removal from the fluid by barite precipitation is more than

313

balanced by addition of SO42- from pyrite oxidation.

314

Fracture surface imaging

315

PW experiments: Changes in fracture volume in x-ray CT images along the main fracture for the

316

PW experiment are below the 24 µm resolution, preventing quantification of fracture volume

317

change (Fig S4). However, a small side fracture does exhibit mineralization (Fig 2). SEM/EDS

318

of that location reveals the minerals filling the fracture to be a combination of barite and gypsum.

319

SEM images of the shale fracture surface near the inlet show dissolution of carbonate minerals,

320

but farther from the inlet, calcite dissolution ceases. SEM images also show precipitated barite

321

and gypsum at the inlet of the core. In some areas, the gypsum contains significant amounts of

322

strontium. These results are similar to previous experiments exposing samples of Marcellus

323

Shale to synthetic reused produced water at reservoir conditions.33

16

ACS Paragon Plus Environment

Page 16 of 35

Page 17 of 35

Environmental Science & Technology

324 325

Figure 2: PW experiments. A) and B) x-ray CT of shale perpendicular to flow, near fluid outlet,

326

B) with small fracture filled by secondary mineral precipitation. C) and D) SEM images of

327

fracture in B), partially filled with barite and gypsum.

328

PWF experiments: X-ray CT imaging and SEM analysis confirm the extensive dissolution of

329

calcite and precipitation of barite suggested by the fluid chemistry. X-ray CT imaging shows a

330

very bright, dense mineral lining the innermost surface of the main fracture from the inlet to a

331

distance of about 3 mm down the core, and SEM/EDS confirms this is a Ba-S mineral (hereafter,

332

Ba-S minerals are assumed to be barite) (Fig 3). X-ray CT also shows a darker, less dense rim

333

between the barite and unaltered shale extending for the inlet 38 mm; SEM/EDS identifies this as

334

calcite dissolution. In some instances, barite crystals grew into the space vacated by the dissolved

335

calcite (Fig 3). SEM/EDS imaging also shows an amorphous Ca-rich deposit that is interspersed

336

with the barite crystals near the fluid inlet. EDS analysis of the deposit rules out gypsum (no S)

337

and calcite is unlikely given the low pH of the fluid, so the exact composition has yet to be

338

identified. At the inlet, where the fluid first meets the core, x-ray CT and SEM/EDS show barite

17

ACS Paragon Plus Environment

Environmental Science & Technology

339

precipitated in the shape of the fluid distributor (Fig 3), suggesting considerable precipitation of

340

barite upon entry into the core holder. Similar to SWF, x-ray CT images show that when dense

341

groupings of proppant are present, mineral reactions (barite precipitation and calcite dissolution)

342

are more pronounced along open areas of the fracture, the preferred fluid pathway.

343

X-ray CT analysis indicates a net effect fracture volume increase of 55% due to extensive

344

dissolution of calcite in the shale matrix along the fracture (WEO 2). Barite precipitation near the

345

fluid inlet produces a slight decrease in effective fracture volume (1%), but that was more than

346

compensated for by calcite dissolution.

347

PWFNA experiments: There is a visible white coating on the shale fracture surface extending the

348

length of the shale core, though it is concentrated near the fluid inlet (Fig S6). SEM/EDS

349

confirms this is barite. X-ray CT scans show that in addition to precipitating along the main

350

fracture, secondary minerals (presumably barite) filled portions of smaller side fractures.

351

SEM/EDS and x-ray CT show barite precipitated extensively near the fluid inlet, again in the

352

shape of the flow distributor, and filled small fractures at the inlet (Fig 3). Barite also cemented

353

some proppant grains to the fracture face.

354

X-ray CT analysis for these experiments exhibited a net decrease in effective fracture volume

355

(WEO 3). The lack of calcite dissolution meant there was no mechanism for increasing effective

356

fracture volume, while barite precipitation decreased fracture volume by 2%.

357

The barite precipitated in this experiment exhibits many different morphologies (Fig S6). At the

358

inlet of the core, barite forms in both rounded and sharp tabular shapes, sometimes clustering

359

together in masses and sometimes as isolated crystals. In the inlet 38 mm, some of the barite has

360

a more porous look, while other barite crystals are much smaller and more angular. The reasons

361

for this are discussed in the next section.

18

ACS Paragon Plus Environment

Page 18 of 35

Page 19 of 35

362

Environmental Science & Technology

Table 2: Summary of experimental findings for the five different fluids tested. Experimental ID Fluid composition

SW

PW2

SWF2

PWF2

PWFNA2

Spring water

Diluted

Spring water

Diluted

Diluted

produced

with HFF

produced water

produced

water

chemicals

with HFF

water with

chemicals and

HFF

HCl

chemicals

Influent pH

5.1

7.2

1.7

2.0

8.2

Ending pH

7.0

7.9

7.2

6.0

7.4

Starting SI Calcite

-5.9

0.0

-12.8

-10.6

1.9

Ending SI Calcite

-1.0

0.9

1.0

-0.3

1.2

Starting SI Barite

-1.0

0.0

0.4

1.6

1.9

Ending SI Barite

-0.9

0.4

0.4

0.6

1.0

Starting SI Gypsum

-4.1

-3.4

-4.0

-1.8

-1.4

Ending SI Gypsum

-2.3

-2.9

-0.8

-1.9

-1.9

Changes apparent

Calcite

Minor

Calcite

Calcite

Barite and

dissolution,

calcite

dissolution,

dissolution,

minor calcite

gypsum

dissolution,

HFF

barite and

precipitation

precipitation,

barite and

chemical

amorphous Ca-

clay transport

gypsum

residue,

rich

precipitation

minor barite

precipitation

along shale fracture face

precipitation Initial fracture volume (mm3) a

661

1068

1216

Fracture volume lost to secondary mineral

N/A

12

29

19

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 35

precipitation (mm3) Fracture volume lost to secondary amorphous material (HFF chemical residue) (mm3)

24

N/A

N/A

(mm3)

430

601

N/A

Net effective fracture volume change (mm3)

406

589

-29

Net effective fracture volume change (%)

61%

55%

-2%

Matrix porosity added by mineral dissolution

363

a

364

volume changes were quantifiable at x-ray CT resolution (24 µm).

Volume changes were only calculated for the first 38mm of core from the fluid inlet, where

365

20

ACS Paragon Plus Environment

Page 21 of 35

Environmental Science & Technology

366

Figure 3: A-F) PWF and G-

367

L) PWFNA experiments. A-

368

B), and G-H) X-ray CT of

369

inlet 38mm of shale parallel

370

to flow. B) and H) bright

371

spots indicate barite

372

precipitation, and B)

373

darkening indicates calcite

374

dissolution along fracture.

375

C-D) and I-J) X-ray CT of

376

shale perpendicular to flow

377

near fluid inlet. D) and I)

378

with barite densely

379

precipitated in the shape of

380

the flow distributor (M). D) shows pitting from calcite dissolution (3 long dark lines are for core orientation). E) SEM of the fracture

381

parallel to flow with secondary barite in void space created by calcite vein dissolution, and F) close-up of minerals in E). K) and L)

382

SEM of secondary barite filling fractures in J).

21

ACS Paragon Plus Environment

Environmental Science & Technology

383

3.3 Controls on barite precipitation:

384

Barite precipitation occurred in all experiments with synthetic reused produced water. In

385

experiments with produced water and HFF chemicals, dense barite precipitation occurred near

386

the core inlet. At salinities up to those found in these experiments, barite solubility increases with

387

pressure and temperature up to about 150oC, so as the fluid enters the core holder, the solubility

388

should increase.54 There are several potentially complimentary hypotheses for elevated barite

389

precipitation near the fluid inlet: the thermal activation of persulfate in HFF chemicals may add

390

SO42- and increase saturation index with respect to barite, scale inhibitor performance may

391

degrade due to increased temperature and low pH, and the temperature increase may affect

392

nucleation and precipitation kinetics.

393

Effect of temperature

394

Scale inhibitors are less effective at preventing barite precipitation at higher temperature.

395

Previous investigations evaluated barite precipitation from supersaturated solutions with scale

396

inhibitors and found that induction time (time until barite begins to nucleate) in the presence of

397

phosphonate and polycarboxonate scale inhibitors is 10-100 times shorter at temperatures of 50-

398

70oC than at 25oC.55 That temperature swing is similar to the one in this study: when fluid enters

399

the heated core holder the temperature increases from ambient (20oC) to reservoir (65.5oC).

400

Though these experiments utilized a different scale inhibitor (ethylene glycol), it appears the

401

efficacy may have been similarly reduced by increased fluid temperature. Heating of the HFF

402

chemicals also increased dissolved SO42-. Persulfates, which are often included in HFF to

403

decompose gelling agents and decrease fluid viscosity, are activated by heat (50-70oC), acidic

404

pH, and transition metals – all of which may be present during hydraulic fracturing.48, 56 When

405

activated, the persulfate (S2O82-) degrades into sulfate radicals (SO4-) that then act as oxidizers

22

ACS Paragon Plus Environment

Page 22 of 35

Page 23 of 35

Environmental Science & Technology

406

and produce sulfate (SO42-) as a byproduct of the oxidation reaction. The addition of dissolved

407

sulfate from HFF chemicals is supported by the fact that SO42- increases 1.25 mmol/L in SWF

408

control experiments. The degradation of ammonium persulfate in SWF-influent (0.02 wt %)

409

could supply up to 1.8 mmol/L SO42-.

410

Further evidence of SO42- release from persulfate activation is found in the PWF and

411

PWFNA control experiments. In produced water experiments, the presence of Ba2+ and

412

consequent ability to remove SO42- from solution through barite precipitation makes it

413

impossible to directly measure SO42- increase. However, Ba2+ can be used as a proxy for SO42-.

414

In the PWF control experiment, dissolved Ba2+ decreased by 1.4 mmol/L. If this decrease is

415

driven solely by barite precipitation – which is likely since the low pH precludes formation of

416

barium carbonates – it requires a stoichiometric SO42- decrease. However, PWF influent only

417

contains 0.5 mmol/L SO42-. Given that there was no shale core present in the control, the 0.9

418

mmol/L SO42- difference between what was present in the influent and what was necessary for

419

barite precipitation must have come from HFF chemical decomposition, likely the ammonium

420

persulfate. The SO42- addition could be critical for barite precipitation in the presence of

421

antiscalants for PWF and PWFNA experiments. Influents for these experiments are already

422

supersaturated with respect to barite (SI of 1.6 and 1.9, respectively), and the addition of up to

423

1.8 mmol/L SO42- would increase supersaturation (SI of 2.2 and 2.4, respectively) to above the

424

threshold at which barite precipitates regardless of the presence of antiscalants.22, 57

425

Finally, barite precipitation near the inlet may be driven by the increase in temperature and

426

corresponding increase in barite nucleation and precipitation kinetics.58-59 Though the

427

temperature increase almost doubles barite solubility, from 0.04 to 0.075 mmol BaSO4/kg

428

water,54 the lowest concentration of Ba2+ and SO42- in all of the experiments with produced water

23

ACS Paragon Plus Environment

Environmental Science & Technology

429

(0.18 mmol/kg) remains greater than twice the solubility at 65.5oC. Thus, the increased

430

solubility at higher temperature should not impede barite precipitation. Indeed, increased barite

431

formation at the core inlets in PWF and PWFNA experiments is starkly illustrated by Fig 3. X-

432

ray CT images show barite precipitated in the shape of arcs and triangles matching the shape of

433

the flow distributors. The fluid does not even enter the shale core before barite precipitation

434

begins.

435

Effect of pH

436

The efficacy of scale inhibitors at preventing barite precipitation varies not only with

437

temperature, but also pH. In PWF experiments, barite precipitation was strongly concentrated at

438

the inlet of the core. In PWFNA experiments, while there is a concentrated barite deposit at the

439

core inlet, a thinner layer of barite is distributed throughout the length of the core. The

440

concentration of barite formation near the fluid inlet in the lower pH PWF experiments parallels

441

previous studies on the pH dependence of barite inhibition by phosphonate and polycarboxylate

442

antiscalants; in those experiments at pH