High-Pressure Reverse Osmosis for Energy-Efficient Hypersaline

Jun 29, 2018 - *E-mail: [email protected]. Phone: ... is also discussed, emphasizing several process design considerations unique to HPRO...
0 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Review

High Pressure Reverse Osmosis for Energy-Efficient Hypersaline Brine Desalination: Current Status, Design Considerations, and Research Needs Douglas M Davenport, Akshay Deshmukh, Jay Ryan Werber, and Menachem Elimelech Environ. Sci. Technol. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.estlett.8b00274 • Publication Date (Web): 29 Jun 2018 Downloaded from http://pubs.acs.org on July 4, 2018

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 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 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.

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 28

Environmental Science & Technology Letters

1 2 3 4 5 6

High Pressure Reverse Osmosis for Energy-Efficient Hypersaline Brine Desalination: Current Status, Design Considerations, and Research Needs

7 8 9 10 11 12 13 14 15

Douglas M. Davenport, Akshay Deshmukh, Jay R. Werber, and Menachem Elimelech*

16 17 18 19

Department of Chemical and Environmental Engineering, Yale University, New Haven, Connecticut 06520-8286, United States

20 21 22 23 24

* Corresponding author: Menachem Elimelech, Email: [email protected], Phone: (203) 432-2789

25

1 ACS Paragon Plus Environment

Environmental Science & Technology Letters

26

ABSTRACT

27

Water scarcity, expected to become more widespread in the coming years, demands renewed

28

attention to freshwater protection and management. Critical to this effort is the minimization of

29

freshwater withdrawals and elimination of wastewater discharge, both which can be achieved via

30

zero liquid discharge (ZLD), an aggressive wastewater management approach. Because of the

31

high energetic cost of thermal desalination, ZLD is particularly challenging for high-salinity

32

wastewaters. In this review, we discuss the potential of high pressure reverse osmosis (HPRO)

33

(i.e., reverse osmosis operated at hydraulic pressure greater than ~100 bar) to efficiently

34

desalinate hypersaline brines. We first discuss the inherent energy-efficiency of membrane

35

processes as compared to conventional thermal processes for brine desalination. We then

36

highlight the opportunity of HPRO to reduce energy requirements for desalination of key high-

37

salinity industrial wastewaters. The current state of membrane materials and processes for

38

hypersaline brine desalination is also discussed, emphasizing several process-design

39

considerations unique to HPRO. Lastly, we discuss the most pressing research needs for the

40

development of HPRO, notably the development of membranes and modules suitable for high

41

pressures as well as fundamental studies of compaction and transport at HPRO conditions.

42

2 ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

Environmental Science & Technology Letters

43

INTRODUCTION

44

Recent estimates suggest 1.2 billion people live in areas of physical water scarcity.1 Effective

45

water resource management is critical to ensure the widespread availability of freshwater in the

46

coming decades.2 Two specific concerns are the treatment and disposal of wastewater and the

47

utilization of previously untapped water sources. Wastewater management from industrial

48

sources is particularly challenging as it can be highly saline, vary temporally in volume and

49

composition, and contain complex mixtures of contaminant species.3 The beneficial reuse of

50

industrial wastewaters mitigates unsafe disposal by eliminating wastewater discharge, while

51

simultaneously decreasing freshwater withdrawals.4, 5 A major challenge for the reuse of high-

52

salinity wastewaters, however, is the high energetic cost of brine desalination.6 In the case of

53

inland desalination, brine-disposal costs often render the desalination of brackish groundwater

54

economically unfeasible.7 Low-cost brine management is critically needed to enable the

55

utilization of inland brackish groundwater and prevent the unsafe discharge of saline industrial

56

wastewaters.

57

Due to the energy requirements for high-salinity wastewater desalination, the preferred brine

58

management option  when it is both safe and effective to do so  is to dispose or reuse the

59

waste stream without extensive treatment. For example, brine solutions (also called the retentate

60

or reject) produced from seawater reverse osmosis (SWRO) processes can be safely discharged

61

to the ocean without extensive treatment.8 In the oil and gas industry, produced water undergoes

62

relatively moderate treatment such as dissolved air flotation, chemical precipitation, or advanced

63

oxidation to remove oil and grease and reduce organic content.9 Produced water treated in this

64

fashion is often discharged to the ocean at offshore locations, albeit with some environmental

65

concerns.10 At inland sites, treated produced water can be reused in hydraulic fracturing fluid.9, 11

66

When safe discharge or reuse are not possible, brines are commonly disposed of via deep-

67

well injection or treated using evaporation ponds prior to solids disposal in landfills.9, 11-13 These

68

processes, however, have serious environmental limitations. For instance, some geologic

69

formations, such as those surrounding the Marcellus shale fields in the US, are unsuitable for

70

wastewater disposal via deep-well injection due to a high potential to contaminate

71

groundwater.14 Furthermore, deep-well injection has been linked to groundwater contamination

72

in other areas as well as increased seismic activity.13 Volume minimization using evaporation

3 ACS Paragon Plus Environment

Environmental Science & Technology Letters

73

ponds is similarly problematic as it poses a threat to groundwater and birdlife, requires large

74

areas of land, and is only effective in warm, arid climates.15, 16 The challenges associated with

75

brine management often result in exorbitant disposal costs, which can even lead to otherwise

76

promising brackish water resources remaining untapped as sources for potable water.7 In order to

77

avoid expensive brine disposal, increasing attention is being given to brine concentration and

78

zero liquid discharge (ZLD) practices, in which waste is disposed of in solid form.17

79

A critical step in the ZLD process chain is desalination, wherein water is recovered from a

80

saline waste stream. Desalination can be achieved by membrane-based processes, such as reverse

81

osmosis (RO) and electrodialysis, or phase-change-based (designated here as “thermal”)

82

processes, including multi-effect distillation (MED), multi-stage flash (MSF), and mechanical

83

vapor compression (MVC).18 RO utilizes hydraulic pressure, in excess of solution osmotic

84

pressure, to drive the transport of water across a semi-permeable membrane while retaining most

85

solutes.19 A combination of membrane and thermal processes are often used to achieve ZLD for

86

saline wastewaters.20 First, RO concentrates wastewater to approximately 70,000 mg L-1 total

87

dissolved solids (TDS), which has an osmotic pressure, π, of ~59 bar. At higher salinities,

88

thermal technologies are used to concentrate brine streams to approximately 250,000 mg L-1 (π ≈

89

290 bar), the typical inlet concentration for crystallizers in ZLD.17 Thermal-based brine

90

crystallizers then concentrate the waste stream above its solubility limit (e.g., 357,000 mg L-1 for

91

NaCl) to extract solid salts for disposal.17

92

Because of its superior energy-efficiency, RO has displaced thermal processes in recent

93

decades for the major applications of seawater and brackish water desalination for drinking water

94

production.21 As such, membrane materials and processes have been optimized to treat

95

feedwaters with salinities equal to or less than seawater (typically ~35,000 mg L-1, π ≈ 28 bar).

96

Consequently, the maximum operating pressure for RO is typically around 80 bar,22 suitable to

97

overcome the retentate osmotic pressure of seawater treated to 50% recovery (~70,000 mg L-1, π

98

≈ 59 bar). Due to hydraulic pressure limitations of RO, hypersaline brines, here defined as

99

solutions with a TDS concentration above 70,000 mg L-1, are primarily desalinated via thermal

100

processes.17 These processes, however, are energy- and cost-intensive, which limits the

101

implementation of brine wastewater desalination prior to disposal.

4 ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

Environmental Science & Technology Letters

102

In this article, we critically discuss the application of high-pressure RO (HPRO) — defined

103

here as RO operating above ~100 bar — to the treatment of hypersaline brines. We start by

104

considering the large volumes of industrial brines that could require desalination and the energy

105

requirements of HPRO versus other processes. We largely consider two pressure limits for

106

HPRO: (i) 150 bar as roughly double the current operating limit and (ii) 300 bar to enable

107

retentate concentrations of 250,000 mg L-1, the inlet concentration for brine crystallizers. We

108

then discuss the current state of HPRO and the process design requirements for its effective

109

application. We conclude with discussion of the research needs that must be addressed to enable

110

viable HPRO processes.

111

ENERGY-EFFICIENCY OF MEMBRANE DESALINATION

112

Desalinating highly saline waste streams is inherently energy intensive. The minimum specific

113

energy consumption (SECmin) of desalination (i.e., the energy required by a thermodynamically

114

reversible process per unit volume of product water) depends on the salinity of the feed stream,

115

cF, and the water recovery ratio, R, the proportion of water recovered from the feed stream:23, 24

 =

  + 1 −    −   

(1)

116

where is the specific Gibbs free energy as a function of composition at fixed temperature and

117

pressure, and cP and cR are the solute concentrations in the product and retentate streams,

118

respectively. Assuming complete solute rejection, cP = 0 and  =  / 1 − . Because of its

119

salinity dependence, SECmin is high for hypersaline brines (5.3 kWh m-3 for 125,000 mg L-1 at

120

50% recovery) as compared to more dilute solutions (1.1 kWh m-3 for 35,000 mg L-1 at 50%

121

recovery). Further, the desalination energy efficiency, η, is defined as the useful specific

122

desalination work performed by the system (i.e., SECmin) divided by its specific energy

123

consumption (SEC):  =  /.

124

Membrane-based desalination processes such as RO and electrodialysis, which do not

125

require a phase change to separate water from dissolved solutes, are able to achieve energy

126

efficiencies greater than 40%, particularly with multi-stage systems.24-26 In contrast, phase-

127

change-based or thermal desalination processes, such as MED, MSF, and MVC, are inherently

128

less efficient, with typical η values less than 20%.24,

129

processes is strongly dependent on the recovery of the latent heat of vaporization, Δ . Given

27

The energetic performance of thermal

5 ACS Paragon Plus Environment

Environmental Science & Technology Letters

130

that Δ , greater than 630 kWh m-3 for saline waters, is two orders of magnitude larger than

131

typical SECmin, imperfect heat recovery severely limits the energy efficiency of all thermal

132

desalination processes.27

133

To directly compare membrane- and thermal-based desalination (Figure 1A), we can

134

quantitatively analyze the energetics of relevant processes. Typical SEC values and feed

135

concentrations are shown in Figure 1B for conventional (largely thermal) desalination

136

technologies,17,

137

assuming a terminal hydraulic pressure (∆Pt) of 5 bar above the osmotic pressure of the retentate

138

(πR).26, 28, 29 More detailed insights can be gained by modeling different processes for the same

139

inlet and outlet streams. In particular, a feed solution of 70,000 mg L-1 NaCl is concentrated to

140

250,000 mg L-1 via HPRO-based, MVC-based, and hybrid HPRO-MVC desalination systems

141

(Figure 1A). MVC is chosen as a representative thermal process due to its highly effective heat

142

recovery. Its SEC is calculated assuming isentropic (i.e., adiabatic and reversible) vapor

143

compression with a terminal temperature difference (∆Tt) of 5 oC above the boiling point

144

elevation of the retentate (δR).24, 30, 31 In the two-stage HPRO and MVC models, water recovery

145

in each stage is optimized to minimize SEC. In the hybrid HPRO-MVC model, the HPRO stage

146

is limited to a maximum hydraulic pressure of 150 bar. Further details on SEC calculations are

147

given in the Supporting Information.

27

in addition to SEC calculated for HPRO. The SEC of HPRO is modelled

148

FIGURE 1

149

Figures 1C and 1D show the SEC and energy efficiency of each process, which illustrates

150

the drastic three-fold energy savings made possible using HPRO rather than a thermal

151

desalination process such as MVC. For example, concentration of 70,000 mg L-1 hypersaline

152

feed to 250,000 mg L-1 would require 24 kWh m-3 with two-stage MVC (15% energy efficiency)

153

and only 7.3 kWh m-3 (47% energy efficiency) with two-stage HPRO. For context, the RO stage

154

in a typical SWRO process can operate as low as 1.8 kWh m-3 (59% energy efficiency).25

155

Achieving a retentate concentration of 250,000 mg L-1 with HPRO would require approximately

156

300 bar hydraulic pressure, a challenging target from membrane materials and module design

157

perspectives. For this reason, a hybrid process is modeled where HPRO is limited to a maximum

158

pressure of 150 bar, after which the retentate is further concentrated by MVC. At a retentate

159

concentration of 250,000 mg L-1, the hybrid process SEC is 12 kWh m-3, representing a two-fold 6 ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

Environmental Science & Technology Letters

160

reduction compared to two-stage MVC. Below a retentate osmotic pressure of 150 bar, the SEC

161

of the hybrid HPRO-MVC process is identical to the SEC of one-stage HPRO, as the optimal

162

water recovery ratio of its first stage (HPRO) is almost equal to its final recovery. As the

163

retentate osmotic pressure increases beyond 150 bar, the water recovery of the second stage

164

(MVC) begins to increase, leading to an increase in SEC. It is important to note that the HPRO-

165

MVC hybrid is a two-stage process and thus the brine flowrate entering the less-efficient MVC

166

stage is significantly lower than the initial flow rate (approximately 50% lower for the conditions

167

modeled). Consequently, the energetic performance of the hybrid HPRO-MVC process should

168

be assessed relative to two-stage HPRO and two-stage MVC, rather than one-stage HPRO.

169

POTENTIAL IMPACT OF HPRO

170

A variety of industrial sources generate hypersaline brine wastewaters. Table 1 shows several

171

brines of particular concern due to their large volume, concentration, or strict disposal

172

regulations. In many circumstances, regulatory, financial, and environmental restrictions

173

motivate brine concentration and volume reduction.17 The coal-to-chemicals industry in China,

174

for example, requires ZLD for all new facilities.32 Additionally, the financial cost of brine

175

disposal from inland desalination facilities motivates brine volume minimization.7 Further, strict

176

disposal regulations for landfill leachate in the US and effluent from the textile industry in India

177

necessitate brine volume reduction to minimize environmental impacts and health risks, despite

178

the financial and energetic cost of brine desalination.33, 34

179

TABLE 1

180

Figure 2A shows the global desalination capacity which is currently installed or under

181

construction for brine feedwaters with salinity ≥ 50,000 mg L-1 TDS.35 At present, these brine

182

desalination processes utilize conventional thermal separation processes (e.g. MVC, MED, MSF)

183

or conventional RO.35 Of the nearly 16,000 desalination plants online or under construction in

184

the GWI/IDA Desalting Inventory, only 110 treat brine feedwaters.35 In practice, the high cost of

185

brine desalination limits hypersaline wastewater desalination before disposal. Therefore, the total

186

volume of hypersaline wastewater present in each industry is surely much greater. In the US, the

187

Environmental Protection Agency cites the high financial cost of brine desalination as the reason

188

it does not require evaporative brine concentration of flue gas desulfurization (FGD)

7 ACS Paragon Plus Environment

Environmental Science & Technology Letters

189

wastewater.36 If cost-effective technologies were available, it is possible regulations would

190

require brine desalination prior to disposal.

191

FIGURE 2

192

At present, thermal technologies account for 54% of the global brine desalination capacity.35

193

The remaining 46% is treated using RO, but information is unavailable on water recoveries and

194

operating pressures. To understand the energetic cost of treating these waste streams, we

195

estimated the energy (Figure 2B) needed to concentrate the total quantity of brine wastewaters

196

currently desalinated in each industry to 250,000 mg L-1 (initial volumes and concentrations are

197

provided in Table S1 of Supporting Information). In these calculations, two-stage RO

198

concentrates wastewater up to 70,000 mg L-1 followed by two-stage HPRO or two-stage MVC.

199

The energy required to desalinate each stream is shown alongside the proportion this represents

200

of the average daily US national electricity generation (NEG).37 In each scenario, MVC would

201

require approximately two to three times more energy than HPRO. In the power industry, for

202

example, approximately one million kWh day-1 (the equivalent of ~34,000 US homes)37 would

203

be saved using HPRO as compared to MVC.

204

Because of the lack of widespread brine desalination, a hypothetical scenario (Figure 2B) is

205

imagined to estimate the potential energy savings of HPRO compared to MVC for each

206

wastewater in Table 1. In this scenario, regulations require all saline wastewaters to be

207

concentrated to 250,000 mg L-1 prior to disposal (i.e., approaching ZLD). The potential energy

208

savings are most notable for large-volume applications like produced water, where MVC would

209

require 130 million kWh day-1, representing 1.2% of the daily US NEG.37 In contrast, HPRO

210

would utilize 85 million kWh day-1 less than MVC, albeit still requiring 0.4% of the daily US

211

NEG.37 Consequently, despite being two- to three-fold more efficient than MVC, the high

212

thermodynamic minimum energy of desalination will mean considerable quantities of energy

213

would be required for HPRO. As such, work should be done to model the financial costs and

214

human health and environmental impacts of hypersaline wastewater management38 to identify

215

applications best suited for HPRO.

8 ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

Environmental Science & Technology Letters

216

CURRENT STATE OF MEMBRANE BRINE DESALINATION

217

A few commercially available membranes rated above 80 bar currently exist. Most notable is the

218

Pall Disc TubeTM (DT) Module System, which has been used for landfill leachate treatment.39

219

DT modules are plate-and-frame configuration with large feed channels designed to reduce

220

fouling from highly contaminated leachate streams. As a result, they have a low active

221

membrane area (9 m2 per element)40 compared to high-pressure membranes with a spiral wound

222

module design (27 m2 per element).41 Several DT module configurations are available ranging in

223

maximum pressure during filtration from 70 bar to 150 bar.40 Average water output, however, is

224

fairly low at 3 m3 day-1 per module for the treatment of landfill leachate or other wastewaters.40

225

Membranes were also designed by Toray in the early 2000s for operation up to 100 bar to

226

achieve 60% recovery SWRO.42 Nonetheless, at present, the maximum operating pressure for

227

currently available Toray RO membranes is 83 bar.43 Additionally, Dow recently started

228

manufacturing a specialty line of RO elements rated at operating pressures up to 120 bar.41, 44

229

The largest of these spiral wound elements (8-inch diameter) is rated to generate permeate flow

230

rates of 24.2 m3 day-1, albeit when treating 32,000 mg L-1 NaCl to 8% recovery at 55 bar

231

(standard SWRO test conditions).41 The achievable permeate flowrate during high-pressure

232

operation is unclear.

233

To circumvent the need for high hydraulic pressures, alternative “osmotically-assisted” RO

234

process configurations have recently been proposed that theoretically could use conventional

235

(~70 bar) pressures.45,

236

membrane permeate-side to reduce the osmotic pressure difference across the membrane and

237

thus the required applied hydraulic pressure.45, 46 Several stages would be used to sequentially

238

decrease the feed and permeate osmotic pressures, until finally conventional RO can be used to

239

produce pure water. While these osmotically-assisted RO processes could theoretically avoid

240

high pressure operation, their performance is inherently limited by internal concentration

241

polarization (ICP) in the membrane support layer. As previous studies on forward osmosis have

242

shown, ICP reduces the solute concentration at the permeate-side membrane-solution interface,

243

drastically reducing the transmembrane driving force for water permeation and thus limiting

244

water flux.47, 48

46

These hypothetical processes would use a saline solution at the

9 ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 10 of 28

245

Designing membranes for osmotically-assisted RO to mitigate ICP will be very challenging.

246

An intrinsic tradeoff exists between ICP reduction and mechanical integrity: the membrane

247

properties that decrease ICP (low support layer thickness and high porosity) will decrease the

248

pressure tolerance of the membrane. The use of pressure retarded osmosis (PRO) membranes has

249

been proposed in order to balance these contradictory needs;45 however, recent work has shown

250

dramatic deformation of PRO membranes to take place at 55 bar applied pressure, the maximum

251

pressure achieved prior to failure.49 In addition to ICP, increased capital costs from staging and

252

high frictional pressure drops in the permeate channels49 would also hinder the proposed

253

processes.

254

HPRO PROCESS DESIGN CONSIDERATIONS

255

The design and operation of conventional RO processes is well-established for brackish water

256

RO and SWRO. However, at high pressure and salinity, HPRO will likely require unique design

257

considerations. In this section, we discuss several aspects of HPRO process design which must

258

be addressed for the development of membrane-based hypersaline brine desalination.

259

Concentration polarization (CP) in the feed channel during RO increases the osmotic

260

pressure at the membrane surface, πm, thus increasing the pressure needed to effect flux. Using

261

film theory and the van’t Hoff approximation, CP modulus can be defined as 19

262

   

%

= exp $ & ( '

(2)

263

where πb and πp are the osmotic pressures in the bulk feed and permeate solutions, respectively,

264

Jw is the water flux, and k is the mass transfer coefficient. Neglecting the permeate osmotic

265

pressure (πp = 0 assuming complete solute rejection) and changes in mass transfer coefficient,

266

which will be affected to some extent by salinity (Figure S1), the ratio πm/πb is constant for a

267

given water flux, meaning that the relative increase in osmotic pressure should be similar for

268

different feed solutions. Conversely, the absolute increase in osmotic pressure from CP is

269

correspondingly greater for high-salinity feeds. For example, a modest CP modulus of 1.15

270

increases the feed osmotic pressure of flowback water (modeled as 157,000 mg L-1 TDS)50 by 26

271

bar, approximately the osmotic pressure of seawater.

10 ACS Paragon Plus Environment

Page 11 of 28

Environmental Science & Technology Letters

272

We performed module-scale modeling to assess the impact of CP for one- and two-stage

273

HPRO processes (Figure 3A). Comparing 35,000 mg L-1, 70,000 mg L-1, and 125,000 mg L-1

274

feeds (each at 50% recovery), water fluxes are surprisingly similar. This result is partially a

275

function of conventional direct-pass RO operation where the hydraulic pressure driving force is

276

roughly the same throughout the module: the increased driving force at the head of the module

277

for hypersaline feeds roughly balances the impact of increased CP. To improve process

278

efficiency, inter-stage design (i.e., placing higher-rejection, low-flux membrane elements at the

279

head of the module) may be useful to balance water flux along the module length (Figure S2).51

280

While this analysis only considers a few cases, it appears that SWRO-like water fluxes (10−15 L

281

m-2 h-1) should theoretically be attainable for HPRO, despite the increased CP stemming from the

282

high feed osmotic pressures.

283

FIGURE 3

284

Although CP is expected to impact water flux relatively little in HPRO, it may critically

285

influence inorganic scaling. In RO, sparingly soluble salts near their solubility limit can

286

precipitate due to CP or increased retentate concentrations as water recovery increases.52 Scaling

287

could be a performance-limiting phenomenon in HPRO processes, particularly at high water

288

recovery (e.g., ZLD). Common scale forming species, including Ca2+, Mg2+, CO32-, SO42-, Ba2+,

289

Sr2+, and silica,53,

290

example, has average concentrations of gypsum precursors Ca2+ and SO42- of 2,000 mg L-1 and

291

13,000 mg L-1, respectively.36 If FGD effluent (33,000 mg L-1 TDS)36 was concentrated to

292

250,000 mg L-1 TDS (87% recovery), Ca2+ and SO42- concentrations would be 15,000 mg L-1 and

293

100,000 mg L-1, respectively, far above the 2,400 mg L-1 solubility limit of gypsum.55 Further,

294

antiscalants which are commonly used to prevent scaling in conventional RO53 have been

295

ineffective at high ionic strength.56 Pretreatment processes are commonly employed to remove

296

foulants in conventional RO desalination processes.22 Similarly, pretreatment to remove scale-

297

forming species may be necessary to achieve high water recoveries in HPRO.

54

are present in many hypersaline brine wastewaters. FGD effluent, for

298

An important consideration for HPRO is the large volumes of hypersaline brines which exist

299

globally. To meet this need, HPRO technologies should have a high specific membrane area (i.e.,

300

membrane surface area per unit volume of membrane module) in order to desalinate large-

301

volume brines in compact facilities. Figure 3B shows the specific membrane surface area for 11 ACS Paragon Plus Environment

Environmental Science & Technology Letters

302

several common module design configurations. While hollow fibers would provide the greatest

303

specific areas, their employment in RO is affected by high bore-side pressure drops and poorly-

304

defined shell-side fluid flow.19 In contrast, spiral wound modules are commonly used for

305

conventional RO and, because SWRO-like water fluxes are expected for HPRO (Figure 3A),

306

they are potentially an optimal HPRO module design. The design of modules and pressure

307

vessels suitable for HPRO should be relatively simple using high-strength materials or more

308

robust designs. Costs may increase for auxiliary system components and piping which must

309

withstand high pressure and resist corrosion when exposed to hypersaline brines. The most

310

challenging module design consideration is likely gluing membrane sheets together to form an

311

envelope which can withstand high pressure, although this can likely be achieved using high-

312

strength adhesives. Permeate spacers should also be designed for HPRO to minimize membrane

313

deformation into spacer voids without large increases in frictional pressure losses.

314

From a process design standpoint, high salinities in HPRO will incentivize unconventional

315

configurations to decrease energy consumption. Conventional RO process designs (e.g., one-

316

stage RO) are likely the best options to concentrate feed waters up to 70,000 mg L-1 due to

317

capital cost concerns. At higher concentrations (i.e., HPRO), energy costs would likely

318

incentivize two-stage processes, which would substantially reduce energy requirements

319

compared to a one-stage process, as shown in Figure 1C. Additional configurations are also

320

possible such as batch and semi-batch RO (also called closed-circuit RO), which use retentate

321

recycle streams.26 Energetic modeling has shown these processes could achieve similar energy

322

consumption as two- or three-stage RO, without physical staging.26

323

RESEARCH NEEDS FOR HPRO

324

Conventional RO is a mature technology with high-performing materials, well-understood

325

transport phenomena, and robust process designs.25 However, much is unknown about the

326

behavior of membrane materials and processes at HPRO conditions. In this section, we highlight

327

the primary research needs for the development and large-scale implementation of HPRO.

328

One of the greatest unknown aspects of HPRO is the fundamental effect of membrane

329

compaction (i.e., physical deformation) on membrane performance at high pressure. At present,

330

only a few studies have assessed RO above 100 bar, with maximum tested pressures of 200

331

bar.57, 58 These studies observed decreased water permeability at high pressure and attributed it to 12 ACS Paragon Plus Environment

Page 12 of 28

Page 13 of 28

Environmental Science & Technology Letters

332

compaction, although compaction was not directly observed. Indeed, even at conventional RO

333

pressures, much remains unknown about the nature of compaction. Laboratory-scale studies

334

commonly attribute declining water permeability to compaction; however, recent work indicates

335

fouling may be the primary cause.59 Compaction is best studied in ultrafiltration (UF) processes,

336

where it has a strong dependence on membrane composition and morphology.60-63 Decreased

337

water permeability has been correlated to changes in overall thickness63 and some results

338

speculate it is primarily a result of compaction in the uppermost skin layer of UF membranes.62

339

The compaction behavior of UF membranes may contribute to developing an understanding of

340

RO compaction as the support layer of state-of-the-art thin-film composite (TFC) membranes is

341

typically a polysulfone (PSf) UF membrane. However, the polyamide selective layer dominates

342

resistance in TFC membranes, which complicates direct translation from study of UF

343

membranes. In TFC membranes, compaction is typically considered to occur in the support

344

layer;64 however, a recent study speculated that compaction can occur in the dense selective layer

345

as well.65 A much greater understanding of the location and impact of compaction is critically

346

needed for the development of deformation-resistant HPRO membranes.

347

Once the mechanisms of compaction are understood, high-strength membranes must be

348

fabricated with high water permeability and salt rejection at HPRO conditions. At high pressures,

349

membranes may follow two possible modes of failure: severe compaction, resulting in no

350

permeability, or rupture, resulting in no salt rejection. In the case of severe compaction, polymer

351

networks may deform to the extent they form a dense, impermeable film. Above a certain

352

pressure limit, however, membranes are likely to rupture, causing a catastrophic loss of solute

353

rejection. In spiral wound membranes, rupture is likely to occur within the voids of the permeate-

354

side spacer. Alternatively, in hollow fibers with a shell-side feed, membranes are likely to

355

collapse under pressure.

356

To gain an understanding of membrane mechanical strength, Figure 3C estimates the

357

compression pressure for non-porous hollow fibers fabricated from different polymers, with steel

358

and Kevlar shown for reference. This analysis focuses on hollow fiber membranes because their

359

cylindrical shape is more well-defined than flat sheet membranes supported by permeate spacers,

360

as in spiral wound modules. The von Mises criterion estimates the first point of fiber

361

compression as a function of applied pressure, material compressive yield strength, and fiber 13 ACS Paragon Plus Environment

Environmental Science & Technology Letters

362

radii ratio (details in Supporting Information).66, 67 The results in Figure 3C, calculated for non-

363

porous polymer materials, suggest high pressures may be achievable; however, material strength

364

will decrease considerably for porous materials. A critical research need is to identify porous

365

membrane morphologies that exhibit high compressive yield strength without compromising

366

membrane performance. It can be seen that polysulfone (PSf), the typical TFC RO support layer

367

material, has a relatively large pressure tolerance in comparison to the other polymers shown. As

368

such, it may be necessary to develop novel high-strength support layer materials to provide the

369

pressure resistance needed for HPRO. Figure 3C also shows a plateau in compression pressure

370

may be reached at a certain fiber radii ratio, indicating that increased support layer thickness may

371

not significantly increase pressure resistance and high-strength materials may be needed.

372

In addition to the unknown behavior of compaction, transport mechanisms in HPRO are

373

poorly understood. Water and solute transport in RO is traditionally characterized by the

374

solution-diffusion model.19 While this model is widely accepted and adequate for conventional

375

RO, it has not been tested at salinities and pressures relevant to HPRO. In the solution-diffusion

376

model, salt flux, Js, is typically defined as )* = +∆, where ∆C is the difference in salt

377

concentration between the feed- and permeate-side interfaces of the selective layer and B is the

378

salt permeability coefficient, typically assumed constant for a given membrane.19 However,

379

recent work has found B to increase substantially with salt concentration, possibly due to

380

decreased impacts of electrostatic repulsion to reject ions at higher salinities.49 To understand

381

this behavior, the mechanisms and assumptions of the solution-diffusion model must be studied

382

at HPRO conditions.

383

OUTLOOK FOR HPRO

384

The development of HPRO has the potential to enable energy-efficient and cost-effective

385

hypersaline brine desalination, although, due to the high thermodynamic minimum energy of

386

desalination, the quantity of energy required by HPRO remains relatively large. Therefore,

387

HPRO may be best-suited for applications where strict regulations or operational circumstances

388

preclude direct reuse or safe disposal of saline brines. Nevertheless, saline wastewater discharge

389

to the environment is undesirable. If the development of HPRO enables energy-efficient brine

390

desalination, future regulations may further restrict brine discharge practices, motivating greater

391

need for ZLD. 14 ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

Environmental Science & Technology Letters

392

Decades of work improving conventional RO will provide a foundational knowledge from

393

which HPRO can be developed. It is possible many aspects of conventional RO membrane and

394

process design will directly translate to HPRO. Even so, significant work will be needed where

395

knowledge gaps exist, notably the fundamental nature of compaction. Similarly, inorganic

396

scaling, a performance-limiting phenomenon in conventional RO, may limit the achievable

397

recoveries of HPRO and necessitates the development of effective and efficient pretreatment to

398

remove scale-forming species.

399

Current brine management practices are expensive and environmentally unsustainable. At

400

present, saline waste streams are discharged to the environment, evaporated in large ponds, or

401

injected into the ground. Where ZLD is required, inefficient thermal separation processes must

402

be used. These costly and environmentally unsound disposal practices prevent the use of inland

403

brackish groundwater, often abundant in water scarce regions, and inhibit beneficial reuse of

404

saline industrial wastewater. In order to help increase the utilization of these waters, HPRO

405

offers great promise as an energy-efficient and cost-effective brine desalination technology.

406

ASSOCIATED CONTENT

407

Supporting Information

408

Derivation of energy-consumption equations, methodology of module-scale water flux analysis,

409

methodology for specific area calculations of various module designs, calculation of nonporous

410

hollow fiber compression pressure, industrial wastewater data (Table S1), CP modulus

411

dependence on Re and Sc (Figure S1), water flux as a function of module position (Figure S2),

412

and pressure drop for various feed concentrations (Figure S3). This information is available free

413

of charge on the ACS Publications website.

414

ACKNOWLEDGMENTS

415

We acknowledge the support received from the National Science Foundation under Grant CBET-

416

1701658 and Graduate Research Fellowship DGE-1122492 awarded to J.R.W.

417 418

15 ACS Paragon Plus Environment

Environmental Science & Technology Letters

419

REFERENCES

420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465

1. Water for a Sustainable World; United Nations Educational, Scientific and Cultural Organization: New York, NY, 2015. 2. A Post-2015 Global Goal for Water: Synthesis of key findings and recommendations from UNWater; United Nations Water: New York, NY, 2014. 3. Lefebvre, O.; Moletta, R., Treatment of organic pollution in industrial saline wastewater: A literature review. Water Res. 2006, 40, 3671-3682. 4. Grant, S. B.; Saphores, J.-D.; Feldman, D. L.; Hamilton, A. J.; Fletcher, T. D.; Cook, P. L. M.; Stewardson, M.; Sanders, B. F.; Levin, L. A.; Ambrose, R. F.; Deletic, A.; Brown, R.; Jiang, S. C.; Rosso, D.; Cooper, W. J.; Marusic, I., Taking the “Waste” Out of “Wastewater” for Human Water Security and Ecosystem Sustainability. Science 2012, 337, 681-686. 5. Adewumi, J. R.; Ilemobade, A. A.; Van Zyl, J. E., Treated wastewater reuse in South Africa: Overview, potential and challenges. Resour., Conserv. and Recycl. 2010, 55, 221-231. 6. Zander, A. K.; Elimelech, M.; Furukawa, D. H.; Gleick, P.; Herd, K. R.; Jones, K. L.; Rolchigo, P.; Sethi, S.; Tonner, J.; Vaux, H. J.; Weis, J. S.; Wood, W. W. Desalination: A National Perspective; 978-0309-11923-8; National Research Council: Washington DC, 2008; p 312. 7. Brady, P. V.; Kottenstette, R. J.; Mayer, T. M.; Hightower, M. M., Inland Desalination: Challenges and Research Needs. Journal of Contemporary Water Research & Education 2005, 132, 46-51. 8. Lattemann, S.; Höpner, T., Environmental impact and impact assessment of seawater desalination. Desalination 2008, 220, 1-15. 9. Technical Development Document for the Effluent Limitations Guidelines and Standards for the Oil and Gas Extraction Point Source Category; United States Environmental Protection Agency: Washington DC, 2016. 10. Bakke, T.; Klungsøyr, J.; Sanni, S., Environmental impacts of produced water and drilling waste discharges from the Norwegian offshore petroleum industry. Mar. Environ. Res. 2013, 92, 154-169. 11. Clark, C. E.; Veil, J. A., Produced water volumes and management practices in the United States; Argonne National Laboratory: Argonne, IL, 2009. 12. Shaffer, D. L.; Arias Chavez, L. H.; Ben-Sasson, M.; Romero-Vargas Castrillón, S.; Yip, N. Y.; Elimelech, M., Desalination and Reuse of High-Salinity Shale Gas Produced Water: Drivers, Technologies, and Future Directions. Environ. Sci. Technol. 2013, 47, 9569-9583. 13. Vidic, R. D.; Brantley, S. L.; Vandenbossche, J. M.; Yoxtheimer, D.; Abad, J. D., Impact of Shale Gas Development on Regional Water Quality. Science 2013, 340, 1235009. 14. Lutz, B. D.; Lewis, A. N.; Doyle, M. W., Generation, transport, and disposal of wastewater associated with Marcellus Shale gas development. Water Resour. Res. 2013, 49, 647-656. 15. Ahmed, M.; Shayya, W. H.; Hoey, D.; Mahendran, A.; Morris, R.; Al-Handaly, J., Use of evaporation ponds for brine disposal in desalination plants. Desalination 2000, 130, 155-168. 16. Ramirez, P., Bird Mortality in Oil Field Wastewater Disposal Facilities. Environ. Manage. 2010, 46, 820-826. 17. Tong, T.; Elimelech, M., The Global Rise of Zero Liquid Discharge for Wastewater Management: Drivers, Technologies, and Future Directions. Environ. Sci. Technol. 2016, 50, 6846-6855. 18. Bahar, R.; Hawlader, M. N. A.; Woei, L. S., Performance evaluation of a mechanical vapor compression desalination system. Desalination 2004, 166, 123-127. 19. Baker, R. W., Membrane Technology and Applications. McGraw-Hill: New York, NY, 2000. 20. Stanford, B. D.; Leising, J. F.; Bond, R. G.; Snyder, S. A., Inland Desalination: Current Practices, Environmental Implications, and Case Studies in Las Vegas, NV. In Sustainability Science and Engineering, Escobar, I. C.; Schäfer, A. I., Eds. Elsevier: Amsterdam, The Netherlands, 2010; Vol. 2, pp 327-350. 21. Desalination Markets 2016; Global Water Intelligence: Oxford, UK, 2016. 16 ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

Environmental Science & Technology Letters

466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498

22. Fritzmann, C.; Löwenberg, J.; Wintgens, T.; Melin, T., State-of-the-art of reverse osmosis desalination. Desalination 2007, 216, 1-76. 23. Mistry, K.; Lienhard, J., Generalized Least Energy of Separation for Desalination and Other Chemical Separation Processes. Entropy 2013, 15, 2046. 24. Mistry, K. H.; McGovern, R. K.; Thiel, G. P.; Summers, E. K.; Zubair, S. M.; Lienhard, J. H., Entropy Generation Analysis of Desalination Technologies. Entropy 2011, 13, 1829. 25. Elimelech, M.; Phillip, W. A., The Future of Seawater Desalination: Energy, Technology, and the Environment. Science 2011, 333, 712-717. 26. Werber, J. R.; Deshmukh, A.; Elimelech, M., Can batch or semi-batch processes save energy in reverse-osmosis desalination? Desalination 2017, 402, 109-122. 27. Deshmukh, A.; Boo, C.; Karanikola, V.; Lin, S.; Straub, A. P.; Tong, T.; Warsinger, D. M.; Elimelech, M., Membrane distillation at the water-energy nexus: limits, opportunities, and challenges. Energy Environ. Sci. 2018, 11, 1177-1196. 28. Sharqawy, M. H.; Lienhard, J. H.; Zubair, S. M., Thermophysical properties of seawater: a review of existing correlations and data. Desalin. Water Treat. 2010, 16, 354-380. 29. Nayar, K. G.; Sharqawy, M. H.; Banchik, L. D.; Lienhard V, J. H., Thermophysical properties of seawater: A review and new correlations that include pressure dependence. Desalination 2016, 390, 124. 30. Thiel, G. P.; Tow, E. W.; Banchik, L. D.; Chung, H. W.; Lienhard V, J. H., Energy consumption in desalinating produced water from shale oil and gas extraction. Desalination 2015, 366, 94-112. 31. Stoughton, R. W.; Lietzke, M. H., Thermodynamic properties of sea salt solutions. J. Chem. Eng. Data 1967, 12, 101-104. 32. Xiong, R.; Wei, C., Current status and technology trends of zero liquid discharge at coal chemical industry in China. Journal of Water Process Engineering 2017, 19, 346-351. 33. Effluent Limitations Guidelines, Pretreatment Standards, and New Source Performance Standards for the Landfills Point Source Category; Final Rule; United States Environmental Protection Agency: Washington DC, 2000. 34. Environment (Protection) Fifth Amendment Rules, 2016. In India Ministry of Environment, F. a. C. C., Ed. The Gazette of India: New Delhi, India, 2016; Vol. G.S.R. 978(E). 35. GWI/IDA Desalting Inventory. http://www.desaldata.com (February 26, 2018). 36. Technical Development Document for the Effluent Limitations Guidelines and Standards for the Steam Electric Power Generating Point Source Category; United States Environmental Protection Agency Washington DC, 2015.

499 500 501 502 503 504 505 506 507 508 509 510 511

37. Electric Power Annual 2016; United States Energy Information Administration: Washington DC, 2017. 38. Bartholomew, T. V.; Mauter, M. S., Multiobjective Optimization Model for Minimizing Cost and Environmental Impact in Shale Gas Water and Wastewater Management. ACS Sustainable Chem. Eng. 2016, 4, 3728-3735. 39. Renou, S.; Givaudan, J. G.; Poulain, S.; Dirassouyan, F.; Moulin, P., Landfill leachate treatment: Review and opportunity. J. Hazard. Mater. 2008, 150, 468-493. 40. Disc Tube Module System Datasheet; Pall Corporation: Port Washington, NY, 2010. 41. DOW™ XUS180808 Reverse Osmosis Element Product Data Sheet; The Dow Chemical Company: Midland, MI, 2016. 42. Taniguchi, M.; Kurihara, M.; Kimura, S., Behavior of a reverse osmosis plant adopting a brine conversion two-stage process and its computer simulation. J. Membr. Sci. 2001, 183, 249-257. 43. Standard SWRO: TM800M; Toray Industries, Inc.: Tokyo, Japan, 2014.

17 ACS Paragon Plus Environment

Environmental Science & Technology Letters

512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558

44. DOW™ Specialty Membrane XUS180804 and XUS180802 Reverse Osmosis Elements Product Data Sheet; The Dow Chemical Company: Midland, MI, 2017. 45. Bartholomew, T. V.; Mey, L.; Arena, J. T.; Siefert, N. S.; Mauter, M. S., Osmotically assisted reverse osmosis for high salinity brine treatment. Desalination 2017, 421, 3-11. 46. Chen, X.; Yip, N. Y., Unlocking High-Salinity Desalination with Cascading Osmotically Mediated Reverse Osmosis: Energy and Operating Pressure Analysis. Environ. Sci. Technol. 2018, 52, 2242-2250. 47. McCutcheon, J. R.; Elimelech, M., Influence of concentrative and dilutive internal concentration polarization on flux behavior in forward osmosis. J. Membr. Sci. 2006, 284, 237-247. 48. Deshmukh, A.; Yip, N. Y.; Lin, S.; Elimelech, M., Desalination by forward osmosis: Identifying performance limiting parameters through module-scale modeling. J. of Membr. Sci. 2015, 491, 159-167. 49. Straub, A. P.; Osuji, C. O.; Cath, T. Y.; Elimelech, M., Selectivity and Mass Transfer Limitations in Pressure-Retarded Osmosis at High Concentrations and Increased Operating Pressures. Environ. Sci. Technol. 2015, 49, 12551-12559. 50. Haluszczak, L. O.; Rose, A. W.; Kump, L. R., Geochemical evaluation of flowback brine from Marcellus gas wells in Pennsylvania, USA. Appl. Geochem. 2013, 28, 55-61. 51. Peñate, B.; García-Rodríguez, L., Reverse osmosis hybrid membrane inter-stage design: A comparative performance assessment. Desalination 2011, 281, 354-363. 52. Shirazi, S.; Lin, C.-J.; Chen, D., Inorganic fouling of pressure-driven membrane processes — A critical review. Desalination 2010, 250, 236-248. 53. Antony, A.; Low, J. H.; Gray, S.; Childress, A. E.; Le-Clech, P.; Leslie, G., Scale formation and control in high pressure membrane water treatment systems: A review. J. Membr. Sci. 2011, 383, 1-16. 54. Fakhru’l-Razi, A.; Pendashteh, A.; Abdullah, L. C.; Biak, D. R. A.; Madaeni, S. S.; Abidin, Z. Z., Review of technologies for oil and gas produced water treatment. J. Hazard. Mater. 2009, 170, 530-551. 55. Deng, M.; Liu, Q.; Xu, Z., Impact of gypsum supersaturated water on the uptake of copper and xanthate on sphalerite. Miner. Eng. 2013, 49, 165-171. 56. Greenlee, L. F.; Lawler, D. F.; Freeman, B. D.; Marrot, B.; Moulin, P., Reverse osmosis desalination: Water sources, technology, and today's challenges. Water Res. 2009, 43, 2317-2348. 57. Rautenbach, R.; Linn, T., High-pressure reverse osmosis and nanofiltration, a “zero discharge” process combination for the treatment of waste water with severe fouling/scaling potential. Desalination 1996, 105, 63-70. 58. Rautenbach, R.; Linn, T.; Eilers, L., Treatment of severely contaminated waste water by a combination of RO, high-pressure RO and NF — potential and limits of the process. J. Membr. Sci. 2000, 174, 231-241. 59. Van Wagner, E. M.; Sagle, A. C.; Sharma, M. M.; Freeman, B. D., Effect of crossflow testing conditions, including feed pH and continuous feed filtration, on commercial reverse osmosis membrane performance. J. Membr. Sci. 2009, 345, 97-109. 60. Brinkert, L.; Abidine, N.; Aptel, P., On the relation between compaction and mechanical properties for ultrafiltration hollow fibers. J. Membr. Sci. 1993, 77, 123-131. 61. Persson, K. M.; Gekas, V.; Trägårdh, G., Study of membrane compaction and its influence on ultrafiltration water permeability. J. Membr. Sci. 1995, 100, 155-162. 62. Stade, S.; Kallioinen, M.; Mikkola, A.; Tuuva, T.; Mänttäri, M., Reversible and irreversible compaction of ultrafiltration membranes. Sep. Purif. Technol. 2013, 118, 127-134. 63. Aghajani, M.; Maruf, S. H.; Wang, M.; Yoshimura, J.; Pichorim, G.; Greenberg, A.; Ding, Y., Relationship between permeation and deformation for porous membranes. J. Membr. Sci. 2017, 526, 293-300. 64. Jonsson, G., Methods for determining the selectivity of reverse osmosis membranes. Desalination 1977, 24, 19-37.

18 ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598

Environmental Science & Technology Letters

65. Pendergast, M. T. M.; Nygaard, J. M.; Ghosh, A. K.; Hoek, E. M. V., Using nanocomposite materials technology to understand and control reverse osmosis membrane compaction. Desalination 2010, 261, 255-263. 66. Ekiner, O. M.; Vassilatos, G., Polyaramide hollow fibers for hydrogen/methane separation — spinning and properties. J. Membr. Sci. 1990, 53, 259-273. 67. Koh, D.-Y.; McCool, B. A.; Deckman, H. W.; Lively, R. P., Reverse osmosis molecular differentiation of organic liquids using carbon molecular sieve membranes. Science 2016, 353, 804-807. 68. Schock, G.; Miquel, A., Mass transfer and pressure loss in spiral wound modules. Desalination 1987, 64, 339-352. 69. Balster, J.; Pünt, I.; Stamatialis, D. F.; Wessling, M., Multi-layer spacer geometries with improved mass transport. J. Membr. Sci. 2006, 282, 351-361. 70. Inoue, N., Polymers. In Hydrostatic Extrusion: Theory and Applications, Inoue, N.; Nishihara, M., Eds. Springer Netherlands: Dordrecht, 1985; pp 333-362. 71. Raddon, B. J. Biaxial restraint of axially loaded steel cores. M.S., The University of Utah, Ann Arbor, 2010. 72. Deteresa, S. J.; Allen, S. R.; Farris, R. J.; Porter, R. S., Compressive and torsional behaviour of Kevlar 49 fibre. J. Mater. Sci. 1984, 19, 57-72. 73. Renner, R., Pennsylvania to regulate salt discharges. Environ. Sci. Technol. 2009, 43, 6120-6120. 74. Mickley, M. C. Membrane Concentrate Disposal: Practices and Regulation; U.S. Department of the Interior Bureau of Reclamation: Washington DC, 2006. 75. Huang, Y. H.; Peddi, P. K.; Tang, C.; Zeng, H.; Teng, X., Hybrid zero-valent iron process for removing heavy metals and nitrate from flue-gas-desulfurization wastewater. Sep. Purif. Technol. 2013, 118, 690-698. 76. Gingerich, D. B.; Grol, E.; Mauter, M. S., Fundamental challenges and engineering opportunities in flue gas desulfurization wastewater treatment at coal fired power plants. Environ. Sci.: Water Res. Technol. 2018, 4, 909-925. 77. Effluent Limitations Guidelines and Standards for the Steam Electric Power Generating Point Source Category; Final Rule United States Environmental Protection Agency: Washington DC, 2015. 78. Postponement of Certain Compliance Dates for the Effluent Limitations Guidelines and Standards for the Steam Electric Power Generating Point Source Category; United States Environmental Protection Agency: Washington DC, 2017. 79. Browner, C. M.; Fox, J. C.; Grubbs, G. H.; Frace, S. E.; Forsht, E. H.; Ebner, M. C. Development Document for Final Effluent Limitations Guidelines and Standards for the Landfills Point Source Category; United States Environmental Protection Agency: Washington DC, 2000. 80. Dasgupta, J.; Sikder, J.; Chakraborty, S.; Curcio, S.; Drioli, E., Remediation of textile effluents by membrane based treatment techniques: A state of the art review. J. Environ. Manage. 2015, 147, 55-72. 81. Vishnu, G.; Palanisamy, S.; Joseph, K., Assessment of fieldscale zero liquid discharge treatment systems for recovery of water and salt from textile effluents. J. Cleaner Prod. 2008, 16, 1081-1089. 82. Verma, A. K.; Dash, R. R.; Bhunia, P., A review on chemical coagulation/flocculation technologies for removal of colour from textile wastewaters. J. Environ. Manage. 2012, 93, 154-168.

599

19 ACS Paragon Plus Environment

Environmental Science & Technology Letters

600

TOC ART

601

602 603

20 ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

Environmental Science & Technology Letters

604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620

FIGURE 1. (A) Brine desalination process schematics. Three different process configurations are shown for the concentration of a 70 g L-1 NaCl solution to 250 g L-1. (B) Specific energy consumption (SEC) of various desalination processes. The typical feed concentration range and energy consumption are shown for conventional RO and thermal brine concentrators.17, 27 The expected SEC calculated for high pressure reverse osmosis (HPRO), at pressures up to 150 bar and 300 bar, is also shown. (C) Specific energy consumption and (D) energy efficiency for HPRO and mechanical vapor compression (MVC) brine desalination processes. The energy consumption requirements of five membrane- and thermal-based process configurations are analyzed for the desalination of 70 g L-1 NaCl at increasing retentate concentration (i.e., water recovery). Energy efficiency, η, is the ratio of the minimum energy of desalination for a given brine salinity and recovery, SECmin, to the SEC for each process. Shown here are one- and twostage processes for both HPRO and MVC. A hybrid HPRO-MVC process is also shown where an HPRO stage is employed up to a maximum applied pressure of 150 bar followed by an MVC stage. The water recovery ratio of each stage is optimized to reduce specific energy consumption for each two-stage process. Further details about the calculations are provided in the Supporting Information.

621 21 ACS Paragon Plus Environment

Environmental Science & Technology Letters

622 623 624 625 626 627 628 629 630 631 632 633 634 635 636

FIGURE 2. (A) Global installed desalination capacity for brine feedwaters with TDS ≥ 50,000 mg L-1.35 For each industry, hollow bars represent the capacity desalinated by membrane-based processes and solid bars represent brine capacity desalinated by thermal processes. (B) Energy consumption of brine desalination. The left-most portion of the figure shows the daily energy consumption required to desalinate the total quantity of brine wastewaters currently treated in each industry. The right-most portion of the figure shows a hypothetical scenario where the total volume of each saline wastewater presented in Table 1 must be concentrated to 250,000 mg L-1 prior to disposal. In each case, the energy consumption shown is that to concentrate the saline wastewaters to 70,000 mg L-1 using a two-stage RO system followed by either a two-stage HPRO or two-stage MVC process to concentrate the brine to 250,000 mg L-1. Hollow bars represent the HPRO process scheme and solid bars represent MVC. The corresponding percentage of the daily US national electricity generation (NEG) is also shown.37 Additional calculation details are provided in the Supporting Information.

637

22 ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28

Environmental Science & Technology Letters

638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659

FIGURE 3. (A) Average water flux calculated from module-scale membrane process modeling. The average water flux for one- and two-stage HPRO processes at 50% water recovery is shown for different terminal hydraulic pressures, ∆Pt (i.e., applied pressure in excess of the retentate osmotic pressure). Osmotic pressures of the exiting retentate solution are also shown. Mass transfer coefficients are calculated using Sherwood-Reynolds-Schmidt correlations and a water permeability coefficient, A, of 1 L m-2 h-1 bar-1 is assumed. (B) Specific membrane area for common module designs. The specific membrane area is shown as a function of characteristic length along with the permeate production rate for a typical 8-inch diameter, 40-inch long module operating at 10 L m-2 h-1. Characteristic length shown here is the feed channel height for typical spiral wound68 and plate-and-frame modules.69 For hollow fiber modules, the characteristic length is the average distance between fibers for typical module packing densities.19 Schematics are shown where dark blue regions are the feed stream, light blue is the permeate, black is the membrane active layer, gray is the membrane support layer, and the characteristic length, a, is shown in white. (C) Compression pressure for nonporous hollow fibers. The compression pressure is calculated using the von Mises criterion as a function of the fiber radius ratio.66 Shown here are dense, nonporous materials which will have a greater compressive yield strength than porous hollow fiber membranes. The materials shown are polytetrafluoroethylene (PTFE), cellulose acetate (CA), polyvinyl chloride (PVC), polysulfone (PSf), polyimide (PI), A36 Steel, and Kevlar 49. The compressive yield strength, σ, is listed for each material.70-72

660

23 ACS Paragon Plus Environment

Environmental Science & Technology Letters

661 662

Page 24 of 28

Table 1: Characteristics of Significant Brine Sources, Current Practices, and Disposal Regulations brine wastewater oil & gas produced water

representative flowrate United States11 8.7 million m3 day-1

typical TDS (mg L-1) 13,000 - 210,0009

current disposal practice offshore11 direct ocean discharge onshore9, 11 deep well injection, where possible

regulations regarding disposal United States9, 11 regulations vary regionally based on permits Pennsylvania:73 discharge monthly average TDS < 500 mg L-1

reuse for hydraulic fracturing evaporation ponds brackish groundwater desalination retentate

United States35, 56 1.6 million m3 day-1

flue gas desulfurization (FGD) wastewater

United States36 1.2 million m3 day-1

5,000 - 55,00056

surface or sewer discharge74 deep well injection, where possible74

16,000 - 50,00036, 75

settling ponds36, 76 chemical precipitation and surface discharge36, 76

United States74 regulations vary regionally based on permits United States77 monthly average TDS < 24 mg L-1 (effective starting November 2020)78

zero liquid discharge36, 76 landfill leachate

United States79 230,000 m3 day-1

0 - 50,00079

zero liquid discharge via land application or recirculation to the landfill79

United States33 strict discharge regulations on 9 to 14 parameters not including TDS

coal-tochemicals wastewater

China32 320,000 m3 day-1

2,000 - 16,00017

zero liquid discharge 32

China17, 32 zero liquid discharge required for new plants

textile industry wastewater

estimates not available

1,500 -30,00080

zero liquid discharge81

India34 TDS < 2,100 mg L-1 for inland discharge

chemical and biological treatment prior to surface discharge82

663

24 ACS Paragon Plus Environment

APage 25 of 28

-1

70 g L TDS

250 g L TDS water for reuse

Conventional RO ∆Pmax = 80 bar

Hybrid HPRO-MVC ∆Pmax = 150 bar

All-HPRO ∆Pmax = 300 bar

Specific Energy Consumption (kWh m-3)

C

30

40%

Recovery 60%

70% VC ge M 1-Sta VC 2-Stage M

20

-MVC HPRO id r b y H RO ge HP 1-Sta RO 2-Stage HP

10

0 100

150

D Energy Efficiency

-1

Specific Energy Consumption (kWh m -3)

All-Thermal Brine Concentration Environmental Science &BTechnology Letters

80 60 40

20 Conventional RO 0

60%

Retentate Concentration (g L )

HPRO

∆Pmax= 80 bar HPRO ∆Pmax= 300 bar ∆Pmax= 150 bar

0

50

100

150

200

250

-1

Feed Concentration (g L ) 40%

Recovery 60%

70%

2-Stage HPRO

40%

1-Sta ge H PRO Hybrid H PRO-MV C

20%

2-Stage MVC 1-Stage MVC

0% ACS Paragon Plus Environment 100 200 250 -1

Thermal Brine Concentrator

150

200

250

-1

Retentate Concentration (g L )

1 0

2

*

*

*

1 0

5

1 0

4

1 0

B F D L

P o w e r

R e fin in g & C h e m ic a ls

1 0

6

O il & G a s

1 0

7

F o o d & B e v e ra g e

8

M in in g

U n s p e c ifie d

P o w e r

R e fin in g & C h e m ic a ls

1 0

2 -S ta g e H P R O 2 -S ta g e M V C

E le c tr o n ic s

1 0

3

M e ta ls

E le c tr o n ic s

1 0

4

9

1 0

E n e r g y C o n s u m p tio n -1 (k W h d a y )

M in in g

1 0

5

O il & G a s

1 0

F o o d & B e v e ra g e

M e m b ra n e T h e rm a l 6

H y p o th e tic a l S c e n a r io

C

C

L

P ro d u c e d W a te r r a c k is h W a t e r D e s a l lu e G a s e s u lf u r iz a t io n a n d f ill e a c h a te o a l to h e m ic a ls

Page 26 of 28

1 0 % 1 % 0 .1 % 0 .0 1 % 0 .0 0 1 % 0 .0 0 0 1 %

ACS Paragon Plus Environment

3

0 .0 0 0 0 1 %

P e r c e n t D a ily U S E le c tr ic ity G e n e r a tio n

G lo b a l In s ta lle d B r in e 3 e s a lin a tio n C a p a c ity ( m d a y

-1

)

B

M e ta ls

A

C u r r e n t In s ta lle d C a p a c ity

Environmental Science & Technology Letters

= 1 b a r

∆P t

= 5 b a r t

= 1 0 b a r

H o llo w F ib e r

-3

)

t

a

3

1 0

2 -S ta g e

1 -S ta g e

2 -S ta g e

1 -S ta g e

2 -S ta g e

F e e d C o n c e n tr a tio n ( g L

-1

1 2

)

1 0

-4

1 0

-3

1 0

-2

C h a r a c te r is tic L e n g th , a ( m ) ACS Paragon Plus Environment

-1

1 2 5

a

-1

7 0

a

5 0 0

S t e e l σ= 3 1 0 M P P I K e v la r 7 0 0 M P a P S a

a P M 6 0 1 = σ P a M 0 1 1 = f σ

a P M 6 = 7 σ C P V

2 5 0

P a M 8 4 = σ C A

0

P T F E σ= 1 4 M P a

1 .0

2 .0

3 .0

)

3 5

1 0

S p ir a l W o u n d

1 0

d a y

1 -S ta g e

1 0

P la te a n d F ra m e

3

S p e c ific A r e a ( m

2

πr e t = 1 3 3 b a r πr e t = 5 9 b a r πr e t = 2 8 8 b a r

0

7 0

m

2 0

1 0

∆P ∆P

4

C

C o m p r e s s io n P r e s s u r e o f N o n p o r o u s F ib e r s ( b a r )

B

m o d u le

A v e r a g e W a te r F lu x a t 5 0 % -2 -1 W a te r R e c o v e ry (L m h )

A

Environmental Science & Technology Letters

P r o d u c tio n R a te ( m

Page 27 of 28

R a d ii R a tio , r

o u te r

/r

in n e r

TDS = 70,000 mg L-1

Energy Consumption

∆P