Accepted Manuscript Electrochemical sulfide removal and caustic recovery from spent caustic streams Eleni Vaiopoulou, Thomas Provijn, Antonin Prévoteau, Ilje Pikaar, Korneel Rabaey PII:

S0043-1354(16)30038-0

DOI:

10.1016/j.watres.2016.01.039

Reference:

WR 11793

To appear in:

Water Research

Received Date: 6 October 2015 Revised Date:

30 December 2015

Accepted Date: 18 January 2016

Please cite this article as: Vaiopoulou, E., Provijn, T., Prévoteau, A., Pikaar, I., Rabaey, K., Electrochemical sulfide removal and caustic recovery from spent caustic streams, Water Research (2016), doi: 10.1016/j.watres.2016.01.039. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

1 2

4 5

SC

RI PT

3

Electrochemical sulfide removal and caustic recovery from spent caustic

7

streams

M AN U

6

Eleni Vaiopoulou1, Thomas Provijn1, Antonin Prévoteau1, Ilje Pikaar2, Korneel Rabaey1*

8 1

Laboratory of Microbial Ecology & Technology, Faculty of Bioscience Engineering, University

10

of Ghent; Coupure Links 653, 9000 Ghent, Belgium 2

11

School of Civil Engineering, The University of Queensland, Brisbane QLD 4072, Australia *Corresponding author. Tel.: +32 9 264 5985; fax: +32 9 264 6248;

EP

12

TE D

9

E-mail address: [email protected]

AC C

13 14 15

1

ACCEPTED MANUSCRIPT

ABSTRACT

17

Spent caustic streams (SCS) are produced during alkaline scrubbing of sulfide containing sour

18

gases. Conventional methods mainly involve considerable chemical dosing or energy

19

expenditures entailing high cost but limited benefits. Here we propose an electrochemical

20

treatment approach involving anodic sulfide oxidation preferentially to sulfur coupled to

21

cathodic caustic recovery using a two-compartment electrochemical system. Batch experiments

22

showed sulfide removal efficiencies of 84 ± 4% with concomitant 57 ± 4% efficient caustic

23

production in the catholyte at a final concentration of 6.4 ± 0.1 wt% NaOH (1.6 M) at an applied

24

current density of 100 A m-2. Subsequent long-term continuous experiments showed that stable

25

cell voltages (i.e. 2.7 ± 0.1 V) as well as constant sulfide removal efficiencies of 67 ± 5 % at a

26

loading rate of 47 g(S) L-1 h-1 were achieved over a period of 77 days. Caustic was produced at

27

industrially relevant strengths for scrubbing (i.e. 5.1 ± 0.9 wt% NaOH) at current efficiencies of

28

96 ± 2 %. Current density between 0-200 A m-2 and sulfide loading rates of 50-200 g(S) L-1 d-1

29

were tested. The higher the current density the more oxidized the sulfur species produced and the

30

higher the sulfide oxidation. On the contrary, high loading rate resulted in a reduction of sulfide

31

oxidation efficiency. The results obtained in this study together with engineering calculations

32

show that that the proposed process could represent a cost-effective approach for sodium and

33

sulfur recovery from SCS.

AC C

EP

TE D

M AN U

SC

RI PT

16

34 35

Keywords: electrochemical treatment; spent caustic; sulfide; sodium hydroxide; recovery

36 37 2

ACCEPTED MANUSCRIPT

1. Introduction

39

Hydrogen sulfide is a toxic, malodorous and corrosive compound. The removal of sulfide

40

dissolved in wastewater and off-gases from chemical and petrochemical industrial activities

41

represents a considerable cost (Maugans et al., 2010; Paulino and Alfonso, 2012; Veerabhadraiah

42

et al., 2011). The resulting wastewater is known as spent caustic stream (SCS), named after the

43

wasted or used caustic soda. A typical SCS contains 5-12 wt% NaOH and 0.1-4 wt% S2- and can

44

be characterized as sulfidic, cresylic or naphthenic depending on their origin and composition

45

(Alnaizy, R., 2008; Veerabhadraiah et al., 2011). The high pH and sulfide toxicity of SCS limit

46

direct biological treatment, whereas neutralization and dilution may release H2S(g). SCS is a

47

strong reducing agent and has a high oxygen demand (2 mol O2 per mol HS-) (Henshaw and Zhu,

48

2001), resulting in dissolved oxygen depletion.

49

The most commonly used methods to treat SCS involve physico-chemical processes including

50

wet air oxidation and incineration (Alnaizy, R., 2008; Veerabhadraiah et al., 2011), oxidation

51

with oxidant agents addition, precipitation and neutralization/acidification (Tanaka and

52

Takenaka, 1995; Sheu and Weng, 2001), electrochemical (Hariz et al., 2013; Nuñez et al., 2009;

53

Paulino and Alfonso, 2012), biological (De Graaf et al., 2012) or bio-electrochemical processes

54

(Zhang et al., 2013).

55

Despite the variety of available methods to treat SCS, the key limitations that restrict their

56

application are cost, complexity, high consumption of chemicals, safety / handling issues

57

(Alnaizy, R., 2008; Veerabhadraiah et al., 2011) and most importantly the lack of recovered

58

product. Biological processes can alleviate some of these issues, as well as delivering

59

hydrophilic sulfur as recovery product, but require SCS pre-treatment, biomass acclimation and

AC C

EP

TE D

M AN U

SC

RI PT

38

3

ACCEPTED MANUSCRIPT

sludge handling to overcome limitations imposed by high toxicity, pH and COD load.

61

Therefore, there is a general interest in more cost-effective, energy efficient and chemical free

62

methods such as (bio)electrochemical treatment. Several studies showed its feasibility via in situ

63

production of e.g iron (Hariz et al., 2013), hypochlorous acid (Martinie et al., 2006), oxygen or

64

other alike oxidizing agents, and possible coupling of sulfide removal to energy recovery (Kim

65

and Han, 2014; Wei et al., 2012, 2013; Zhang et al., 2013). While these above mentioned studies

66

revealed the potential and can be considered a step forward, they come with some disadvantages

67

including sacrificial anodes, high-energy input to generate oxidizing agents and short life

68

expectancy of materials.

69

Here we propose a novel method that could avoid these concerns. The method relies on the

70

simultaneous anodic oxidization of sulfide coupled to cathodic caustic generation in a two-

71

compartment electrochemical cell. In the anode, sulfide is oxidized to elemental sulfur and other

72

sulfur oxyanions, while in the cathode water is reduced to hydroxide anions. In order to maintain

73

electroneutrality, sodium from the anode migrates through a cation exchange membrane (CEM)

74

that separates the two chambers and allows the selective migration of sodium from the anode to

75

the cathode chamber. Key advantages of this approach would be 1) elimination of chemical

76

dosing for sulfide oxidation and thus, less operational, transport, handling and storage cost of

77

potentially hazardous chemicals, which reduces occupational health and safety concerns, 2)

78

recovery of sodium and oxidized sulfur species that can be re-used in situ or be sold, 3)

79

straightforward process design, 4) potentially low energy demand that can be sourced from

80

renewable supply and 5) a neutralized stream with lower salinity and sulfide is generated towards

81

discharge. Therefore, the overall objective of this study is to investigate the feasibility of this

82

approach and identify the key operational aspects.

AC C

EP

TE D

M AN U

SC

RI PT

60

4

ACCEPTED MANUSCRIPT

2. Materials and Methods

84

2.1 Reactor setup and operation

85

2.1.1 Batch-fed reactor

86

The electrochemical cell consisted of two parallel Perspex frames with internal dimensions of 20

87

× 5 × 2 cm separated by a CEM (Fumasep FKB-PK-130, Fumatech GmbH, Germany) according

88

to (Pikaar et al., 2011). A reference electrode (Ag/AgCl (3M KCl), ALS, Japan, + 0.210 V vs.

89

SHE at 25 °C) was placed in the anode compartment. A flattened mesh shaped tantalum-iridium

90

mixed metal oxide (TaO2/IrO2 : 0.65/0.35) coated titanium electrode (Magneto Anodes BV, The

91

Netherlands) with a projected surface area of 100 cm2 was used as anode material. Stainless steel

92

fine mesh (projected surface area of 100 cm2) was used as cathode (mesh width 44 mµ, wire

93

thickness: 33 mµ, Solana, Belgium) and a stainless steel frame was serving as current collector.

94

A spacer (ElectroCell Europe A/S, Tarm, Denmark) was placed between the electrodes and the

95

CEM to prevent membrane contact with the electrodes. The batch electrochemical cell was

96

galvanostatically controlled using a power source (type PL-3003D, Protek) at a current density of

97

100 A m-2.

98

The anolyte consisted of 4 wt% NaOH and 1 wt% Na2S-S simulating a typical SCS. The

99

catholyte was 4 wt% NaOH at the onset of the experiment which ensured sufficient initial

AC C

EP

TE D

M AN U

SC

RI PT

83

100

conductivity and avoided putative non-electrically driven diffusion of sodium across the CEM. A

101

recirculation flow of 6 L h-1 was applied to obtain sufficient mixing by a peristaltic pump

102

(Watson-Marlow Inc., Massachusetts, US). Masterflex Norprene tubing with an internal diameter

103

of 6 mm was used for both anolyte and catholyte and recirculation lines. H2 produced in the

104

cathode was collected in the cathode effluent bottle. 5

ACCEPTED MANUSCRIPT

105

Figure 1 2.1.2 Continuous reactor operation

107

The continuous-mode electrochemical cell (Fig. 1 red lines) was set up as a low-volume cell

108

(internal dimensions of the cell compartments were 7 × 2 × 1 cm with 5 × 2 cm effective

109

membrane area). SCS of the same composition as for batch mode was used as anolyte. The

110

catholyte was initially 4 wt% sodium hydroxide and then distilled water was fed continuously at

111

a flow rate of 82 ± 17 mL d-1 (HRT 4 h). Recirculation flow was set at 2 L h-1 to provide

112

sufficient mixing in both compartments. Peristaltic pumps (Watson-Marlow Inc., Massachusetts,

113

US) and flows were verified daily to assure accuracy and calculate standard deviations.

114

Three different sets of experiments were performed. In the first one, the reactor was run in a

115

continuous mode to determine a long-term operation performance at a fixed current density of

116

100 A m-2. The operation time was run in four periods: 1) day 0-25 and 35-49; the anolyte flow

117

rate was 128 ± 5 mL d-1 and sulfide loading rate (SLR) of 40 ± 3 g(S) L-1 d-1, 2) day 25-35; anode

118

flow rate was decreased to 73 ± 6 mL d-1 and SLR was 26 ± 2 g(S) L-1d-1, 3) day 49-104; batch

119

mode operation at 0.5 A m-2 (reactor remains assembled) and 4) day 104-132; anolyte flow rate

120

was increased to 134 ± 6 mL d-1 and SLR was 47 ± 2 g(S) L-1d-1. The catholyte flow rate was kept

121

constant at 82 ± 17 mL d-1. Along with flow rates, cell voltage, sulfur species, sodium and

122

hydroxide concentrations were monitored on a daily basis. The second experiments assessed the

123

impact of current density on reactor performance and sulfur speciation. Experiments were run at

124

50, 100, 150 and 200 A m-2 and the values presented herein are the ones recorded in triplicates

125

once a new steady state was reached (typically following the 5 times the hydraulic residence time

126

thumb rule and as long as concentrations remained constant). Anolyte and catholyte flow rates

AC C

EP

TE D

M AN U

SC

RI PT

106

6

ACCEPTED MANUSCRIPT

were 121 ± 10 mL d-1 at SLR of 42 ± 4 g(S) L-1d-1 and 73 ± 3 mL d-1 respectively. The third

128

experiments aimed to investigate the impact of SLR - by applying different flow rates ranging

129

from 135 to 530 mL d-1. Experiments were run at 50, 100, 150 and 200 g(S) L-1d-1 at 100 A m-2

130

and the values presented herein are the ones recorded in triplicates once a new steady state was

131

reached. The ratio between the smallest SLR and smallest flow rate is slightly different than the

132

ratio of the highest SLR and highest flow rate due to slight differences in HS- concentration

133

when preparing feeding. An open circuit replicate was run to confirm that sulfide removal and

134

NaOH recovery are only driven by the applied current. To assess whether sulfur species were

135

crossing the membrane, catholyte samples were taken periodically.

136

2.3 Chemical analysis

137

Samples from the reactor were immediately preserved in previously prepared Sulfide

138

Antioxidant Buffer solution prior to analysis as suggested by Keller-Lehmann et al. (2006).

139

Sulfide, sulfite (SO32-) and thiosulfate (S2O32-) concentrations were measured by ion

140

chromatography (IC), using an IC930 compact Metrohm IC system (Metrohm, Switzerland),

141

according to Keller-Lehmann et al. (2006). The eluent consists of 3.5 mM Na2CO3 and 3mM

142

NaHCO3 at a flow rate of 0.8 mL min-1. A 0.1 M NaOH solution is used to produce a pH

143

gradient needed for thiosulfate detection in the IC system. Sulfate (SO42-) was determined on an

144

IC761 compact Metrohm IC system (Metrohm, Switzerland) equipped with a Metrosep A Supp

145

5-150 anion exchange column and a conductivity detector, according to Standard Methods

146

(APHA, 1992). The eluent, consisting of 3.2 mM Na2CO3 and 1 mM NaHCO3, had a flow rate of

147

0.7 mL min-1. To measure the polysulfide and elemental sulfur concentrations, all sulfur species

148

were oxidized to sulfate with excess H2O2 as described elsewhere (Dutta et al., 2010). The

149

difference in sulfur equivalent between the sulfate after H2O2 oxidation and other species

AC C

EP

TE D

M AN U

SC

RI PT

127

7

ACCEPTED MANUSCRIPT

measured before H2O2 oxidation (i.e. sulfide, sulfate, thiosulfate and sulfite) was regarded as the

151

sum of polysulfides and elemental sulfur.

152

Alkalinity, as indicator of NaOH concentration, is measured by titrating the cathode effluent with

153

a 1 M HCl solution. Sodium was determined following procedures outlined in Standard Methods

154

(APHA, 1992), using an IC761 compact Metrohm IC system (Metrohm, Switzerland) equipped

155

with Metrosep C6-250/4.0 cation-exchange column and a conductivity meter. The eluent,

156

consisting of 1.7 mM HNO3 and 1.7 mM dipicolinic acid, ran at a flow rate of 0.9 mL min-1.

157

All samples were run in triplicates. Experimental values are provided as the mean +/- standard

158

deviation.

159

2.4 Electrochemical measurements and calculations

160

During the continuous mode operation, cell voltage was monitored every 3 min with a VSP

161

multichannel potentiostat (Princeton Applied Research, France). The electrical resistance of the

162

cells was monitored by the current interrupt technique as described in the Supplementary

163

Material. Current densities are reported with respect to the projected surface area of the anode.

164

Depending on sulfide oxidation product, the oxidation process can involve different amount of

165

electrons per sulfide consumed. Anodic reactions of sulfide oxidation are described elsewhere

166

(Dutta et al., 2010). The coulombic efficiency (CE) for sulfide conversion was calculated

167

assuming only the 2-electron conversion of sulfide to elemental sulfur as described elsewhere

168

(Dutta et al., 2010). CE for sodium recovery is only restricted by membrane transport and was

169

calculated as the ratio of sodium theoretically transferred due to current applied to hydroxyl

170

anions produced (based on the assumption sodium cations and hydroxyl anions are equal, since

171

no other cations are in solution).

AC C

EP

TE D

M AN U

SC

RI PT

150

8

ACCEPTED MANUSCRIPT

3. Results and Discussion

173

3.1 Batch mode reactor

174

At a fixed current density of 100 A m-2, 84 ± 4% of the sulfide was converted (Fig.2) at a

175

coulombic efficiency (CE) of 75 ± 4%. A NaOH solution was recovered at an efficiency over the

176

batch of 57 ± 4% at a CE of 91 ± 5%. The final concentration was 6.4 ± 0.1 wt% NaOH (1.6 M)

177

after 8 h experiments. The pH of the anolyte was initially 13.7 but decreased to 13.2 ± 0.1 at the

178

end of the experiment due to proton production from water splitting in the anode compartment

179

(Fig.2). The remaining high pH cannot be directly discharged, but ensures that residual sulfide

180

remains into solution rather than stripping off (pKa (H2S/HS-) = 6.9). In terms of energy input,

181

cell voltage evolution shows a moderate increase over time from 1.79 to 2.47 V (Fig. S1). This

182

increase is mainly because of sulfide depletion in the anolyte (low conductivity measured);

183

eventually leading to more energy demanding O2 evolution once HS- mass transfer cannot

184

sustain the current applied. This assumption is further based on the low internal resistance of the

185

cell, which was 0.10 ± 0.03 Ω (SI). At this current (1 A), this implies an ohmic drop of 0.10 ±

186

0.03 V accounting for about 4% of the operating voltage. The relatively low ohmic drop is

187

attributed to the high conductivity of both electrolytes.

EP

TE D

M AN U

SC

RI PT

172

AC C

188

Figure 2

189

The effluent NaOH concentration of 1.6 M (at a fixed current density of 100 A m-2) is considered

190

high, when compared with an electrodialysis SCS treatment that recovered up to 0.15 M NaOH

191

at higher current densities (800 A m-2) (Wei et al., 2012). In the same study, experiments at fixed

192

current density of 300 A m2 resulted in 0.05 M NaOH in 2 h batch experiments at CE of 100%.

9

ACCEPTED MANUSCRIPT

The higher CE in this case can be explained by the smaller electrode (7 cm2) and bigger

194

catholyte volume leading to lower cathode concentrations ( > 500 cm2).

195

3.2 Continuous mode reactor

196

The electrochemical cell was operated in a continuous mode for 77 d exhibiting stable

197

performance and reproducibility as confirmed by stable cell voltage, sulfide removal efficiency

198

and restored values when disturbed (Fig.3). Cell voltage was constant at 2.74 ± 0.10 V for a

199

sulfide loading rate (SLR) of 47 ± 2.2 g(S) L-1 h-1 and flow rate of 0.13 L d-1 in the anodic

200

chamber. A SLR decrease from 40 ± 3.2 (day 0-25) to 26 ± 2.2 g(S) L-1 h-1 (day 25-34) resulted in

201

higher cell voltage of 3.08 ± 0.26 V. The cell voltage was restored after the SLR increased again

202

to 47 g(S) L-1 h-1 (day 35 - 45) (Fig.3). The same reproducible behavior was recorded after the

203

batch mode operation period (day 49 - 104) and unexpected single day lab implications (day 47

204

and day 124). Sulfide conversion efficiency for the overall 77 d of continuous operation was 67

205

± 5 % at a CE of 54 ± 7 % for both SLR of 47 and 26 g(S) L-1 h-1. For the different SLR applied,

206

sulfide conversion efficiency was 68 ± 5 % at a CE of 69 ± 3 % for a SLR of 47 g(S) L-1 h-1.

207

When the SLR was decreased to 26 g(S) L-1 h-1, these efficiencies were 71 ± 8% and 45 ± 6%

208

respectively,

AC C

209

EP

TE D

M AN U

SC

RI PT

193

Figure 3

210

Influent sulfide of 10.8 ± 0.3 g(S) L-1 was oxidized for 30% to thiosulfate (3.2 ± 0.3 g(S) L-1), then

211

to sulfate (1.4 ± 0.1 g(S) L-1) for 13 %, to polysulfide and elemental sulfur (≈ 0.6 ± 0.7 g(S) L-1) for

212

5.5 %, while 3.7 ± 0.1 g(S) L-1 of sulfide remained unconverted. These values come from analysis

213

data and do not include the elemental sulfur that has been deposited on the electrode or other

214

parts of the reactor. 10

ACCEPTED MANUSCRIPT

Clean NaOH solution was recovered in the catholyte via hydroxide electrogeneration and Na+

216

migration across the CEM. As Na+ was the only cation present, alkalinity analysis results can be

217

linked to NaOH concentrations. NaOH was produced at high CE (96 ± 2%) with a catholyte

218

effluent concentration of 1.3 ± 0.1 M (5.1 ± 0.4 wt%).

219

The iridium-tantalum oxide coated titanium anode showed stable performance through the whole

220

experimental period of about 5 months, despite the harsh conditions of high sulfide

221

concentrations and high pH. A study applying higher current densities than our range of 0-200 A

222

m-2 reported corrosion and limited lifetime of a mixed iridium-tantalum oxide coated titanium

223

electrode used as anode (Behm & Simonsson, 1997b). Corrosion in this case was attributed to

224

the hydrodynamics, as it occurred mainly on points of turbulence, and at high applied potentials.

225

Electrochemical sulfide oxidation at carbon electrodes can result in an elemental sulfur layer on

226

the anode, causing electrode passivation (Dutta et al., 2008). Here the anode remained fully

227

functional through the whole experimental period due to the different electrode material used

228

(higher oxidizing power) and higher pH (neutral vs highly alkaline). In our study, possible

229

mechanisms of continuous reactivation of the anode could be a de-flaking process by

230

concomitant oxygen generation on the anode surface, sulfur dissolution by polysulfide anions in

231

alkaline condition or further oxidation of sulfur in contact with the electrode to dissolved

232

thiosulfate or sulfate. A joint mechanism could be possible if the inner layer of the elemental

233

sulfur is oxidized to sulfur oxyanions while the remaining sulfur in the outer layer reacts with

234

sulfides to form polysulphide (Behm & Simonsson, 1997a). This appears the most plausible

235

mechanism as there was no observation of flakes or sulfur particles in the reactor or the effluent.

236

3.3.Impact of current density and sulfide loading rate on sulfide oxidation products

AC C

EP

TE D

M AN U

SC

RI PT

215

11

ACCEPTED MANUSCRIPT

Current density affects anode potential and thus, further affects anode sulfide oxidation and the

238

formation of more oxidized sulfur species. The higher the current density, the higher the

239

proportion of sulfur oxyanions such as thiosulfate and sulfate became (Fig.4A). Controls in

240

absence of current (OCV) showed no sulfide oxidation. At 50 A m-2 no sulfate but elemental

241

sulfur and polysulfide were produced, at 100 and 150 A m-2 sulfide oxidation was gradually

242

enhanced and concentrations of thiosulfate and sulfate increased, whereas at 200 A m-2 no steady

243

state was achieved due to constantly increasing cell voltage (Fig.S2). Cell voltage attained the

244

same values on the current density steps backwards (Fig.S2). The same trends were also

245

confirmed for sulfur species production when repeating the experiments applying the reverse

246

scheme of current densities, i.e. from 200 to 0 by steps of 50 A m-2 (Fig.4A, S2). The production

247

of more oxidized sulfur species in higher current density is in agreement with the previously

248

described sulfide oxidation mechanism. The same behaviour in high alkaline conditions has been

249

observed elsewhere (Behm & Simonsson, 1997a; Kim and Han, 2014).

TE D

M AN U

SC

RI PT

237

250

Figure 4

Sulfide removal efficiency reached a maximum of 86 ± 3 % at 150 A m-2. The lowest recorded

252

sulfide removal was 73 ± 1 % at 50 A m-2. The CE of sulfide oxidation decreased with increasing

253

current density, as expected. A CE higher than 100 % for sulfide oxidation at 50 A m-2 is likely

254

related to polysulfide formation occurring at high pH at the electrode interface (Fig.4A) via the

255

chemical reaction between previously formed S0 and dissolved sulfide (Behm & Simonsson,

256

1997a). This reaction is thermodynamically and kinetically favored at high pH (Steudel and

257

Eckert, 2003), which in our study is above 13. High sulfide loading rates (SLR) due to high

258

influent flow rates resulted in lower overall sulfide removal and less oxidized sulfur species,

259

which further induced elemental sulfur and polysulfide formation at high pH (Fig.4B). Higher

AC C

EP

251

12

ACCEPTED MANUSCRIPT

SLR implied lower steady state voltage values and higher CE (Fig.S3). This is an apparent effect

261

as it is mostly linked to the lower anode potential and low mass transfer limiting current for

262

sulfide oxidation. Although it was not possible to differentiate analytically between elemental

263

sulfur and polysulfide, at lower current densities the anode effluent had a characteristic intense

264

yellow color that is indicative of polysulfide (Steudel and Eckert, 2003). Results regarding sulfur

265

speciation, as far as polysulfide and elemental sulfur are concerned, were mostly qualitative

266

based on the color and texture of the anode effluent.

267

The NaOH production rate increased linearly with the current density increase at high CE (98 ± 3

268

%). Effluent concentration of NaOH increased from 2.7 ± 0.1 to 9.8 ± 1.4 wt% at current

269

densities from 50 to 200 A m-2. Theoretically expected NaOH concentration values were

270

calculated according to the current applied to drive the crossing of Na+ from anode to cathode

271

compartment and catholyte flow rate, which was recorded daily (73 ± 3 mL d-1). Production

272

efficiency of NaOH was 76 ± 2 % for current densities of 50 to 150 A m-2, based on the ratio of

273

expected concentrations to actual sodium measurements. These findings imply, as expected, that

274

the current density has within the range tested had no significant effect on CE, whereas caustic

275

strength can have a greater impact as at higher NaOH concentrations, sodium and hydroxyl ion

276

back diffusion might happen. Since the SLR experiments are based on flow rate changes in the

277

anolyte, they have also no impact on the NaOH recovery. Effluent concentration of NaOH ranges

278

from 5.2 ± 0.2 to 6.4 ± 0.8 wt% when SLR ranged from 50 to 150 g(S) L-1 h-1 and current density

279

remained fixed at 100 A m-2. The fluctuations are explained by influent catholyte flow rate

280

fluctuations ranging between 54 and 81 mL d-1. The measured values correspond well with the

281

theoretically predicted values and they are well reproducible on the backwards step from 200 to

282

50 g(S) L-1 h-1. CE is recorded close to 100 ±7 % for all experiments.

AC C

EP

TE D

M AN U

SC

RI PT

260

13

ACCEPTED MANUSCRIPT

3.5 Implications for practice

284

Electrochemical SCS treatment was technologically feasible and resulted in recovery of sodium

285

as a clean sodium hydroxide solution and different sulfide oxidation products, such as elemental

286

sulfur, polysulfide, thiosulfate and sulfate. To implement such process into practice, there are a

287

number of considerations to be addressed as a first step. These include design issues, economic

288

viability, and optimization in terms of anode potential, higher caustic strengths, different

289

membranes to increase stability of the membrane and electrode over longer period of time etc.

290

Further, the presence of other compounds such as organic pollutants need to be considered.

291

Economics of the process would rely on the cost of investment, including materials and

292

engineering cost, the operational cost and possible savings from recovery of chemicals and

293

omission of other costly methods to remove sulfide. An estimation of this cost has been

294

calculated as an indication only (Table S1). Further process optimization and testing with real

295

wastewater is required before up-scaling this technology and calculating in more detail the

296

process cost. Another preliminary techno-economical approach estimated electrochemical NaOH

297

recovery cost from SCS at less than 1 USD per kg NaOH (Wei et al., 2012).

298

4. Conclusions

300

Electrochemical spent caustic steams treatment has been shown herein to potentially be

AC C



299

EP

TE D

M AN U

SC

RI PT

283

an economically feasible and environmental friendly process that can recover valuable

301

products from waste.



302 303

Sulfide conversion can be driven towards elemental sulfur, polysulfide, thiosulfate and sulfate depending on the current density and sulfide loading rate.

14

ACCEPTED MANUSCRIPT



304

alleviate investment and operational costs. •

306 307

Electrochemical sulfide oxidation and sodium recovery processes were robust and reactor materials remained unaffected during operation.



308

RI PT

305

Besides sulfur species, sodium can be recovered at high coulombic efficiencies and

Before this concept can be applied for more industrial applications and thus, promote implementation of sustainable technology and resource recovery, rigorous economic

310

assessment, process optimization and testing in field conditions are required.

SC

309

M AN U

311

AC C

EP

TE D

312

15

ACCEPTED MANUSCRIPT

References

314

Alnaizy, R., 2008. Economic Analysis for Wet Oxidation Processes for the Treatment of Mixed

315

Refinery Spent Caustic. Environmental Progress 27 (3), 295–301.

316

American Public Health Association (APHA), Greenberg, A.E., Clesceri, L.S. and Eaton, A.D

317

1992. Standard Methods for the Examination of Water and Wastewater, 18th ed. Washington,

318

DC, USA.

319

Behm, M. & Simonsson, D. 1997a. Electrochemical production of polysulfides and sodium

320

hydroxide from white liquor Part I: Experiments with rotating disc and ring-disc electrodes

321

Journal of applied electrochemistry 27, 507-518.

322

Behm, M. & Simonsson, D. 1997b. Electrochemical production of polysulfides and sodium

323

hydroxide from white liquor Part II: Electrolysis in a laboratory scale flow cell. Journal of

324

applied electrochemistry 27, 519-528.

325

Buisman, C.J.N, Vellinga, S.H.J., Janssen, G.H.R. and Dijkman, H. 1999. Biological sulfide

326

production for metal recovery. In TMS Congress 1999, Fundamentals of lead and zinc extraction

327

for metal recovery 1–7.

328

Hariz, I. Ben, Halleb, A., Adhoum, N., Monser, L. 2013. Treatment of petroleum refinery

329

sulfidic spent caustic wastes by electrocoagulation. Separation and Purification Technology 107,

330

150–157.

331

Dutta, P.K., Rabaey, K., Yuan, Z., Keller, J. 2008. Spontaneous electrochemical removal of

332

aqueous sulfide. Water Research 42, 4965 – 4975.

AC C

EP

TE D

M AN U

SC

RI PT

313

16

ACCEPTED MANUSCRIPT

Dutta, P.K., Rabaey, K., Yuan, Z., Rozendal, R.A., Keller, J. 2010. Electrochemical sulfide

334

removal and recovery from paper mill anaerobic treatment effluent Water Research 44, 2563 –

335

2571.

336

Henshaw, P. F., Zhu, W. 2001. Biological conversion of hydrogen sulphide to elemental sulphur

337

in a fixed-film continuous flow photo-reactor Water Research 35 (15), 3605-3610.

338

Keller-Lehmann, B., Corrie, S., Ravn, R., Yuan, Z., Keller, J., 2006. Preservation and

339

simultaneous analysis of relevant soluble sulfur species in sewage samples. In: 2nd International

340

IWA Conference on Sewer Operation and Maintenance, Vienna, Austria.

341

Kim, K. and Han, J.-I., 2014. Performance of direct alkaline sulfide fuel cell without sulfur

342

deposition on anode International Journal of Hydrogen Energy 39, 7142 – 7146.

343

Paulino, J.F. and Alfonso, J. 2012. New Strategies For Treatment and Reuse of Spent Sulfidic

344

Caustic Streams from Petroleum Industry Quim. Nova 35 (7), 1447–1452

345

Pikaar, I., Rozendal, R.A., Yuan, Z., Keller, J., Rabaey, K. 2011. Electrochemical sulfide

346

removal from synthetic and real domestic wastewater at high current densities. Water Research

347

45, 2281 – 2289.

348

Martinie, G.D., Shabrani, F.M.A., Dabrousi, B.O. 2006. Process for treating a Sulfur-containing

349

spent caustic refinery stream using a membrane electrolyzer powered by a fuel cell. Pub. No.: US

350

2006/0254930 A.

351

Maugans, C., Howdeshell, M., De Haan, S. Apr 2010. Update: Spent caustic treatment.

352

Hydrocarbon Processing, 89 (4), 61.

AC C

EP

TE D

M AN U

SC

RI PT

333

17

ACCEPTED MANUSCRIPT

Nuñez P., Hansen, H.K., Rodriguez, N., Guzman, J., Gutierrez C. 2009. Electrochemical

354

Generation of Fenton's Reagent to Treat Spent Caustic Wastewater. Separation Science and

355

Technology 44, 2223-2233.

356

Sheu, S.H. and Weng H.S. 2001. Treatment of olefin plant spent caustic by combination of

357

neutralization and Fenton reaction. Water Research 35 (8), 2017–2021.

358

Steudel, R. and Eckert, B. 2003. Solid Sulfur Allotropes. Topics in Current Chemistry, 230, 1–

359

79.

360

Tanaka, N. and Takenaka, K., 1995. Control of hydrogen sulfide and degradation of organic

361

matter by air injection into a wastewater force main. Water Science and Technology 31 (7), 273-

362

282.

363

Veerabhadraiah, G., Mallika, N., Jindal, S. 2011. Spent caustic management: Remediation

364

review. Hydrocarbon Processing

365

(http://www.hydrocarbonprocessing.com/Article/2925664/Spent-caustic-management-

366

Remediation-review.html).

367

Wei, Y. Li, C., Wang, Y., Zhang, X., Li, Q., Xu, T. 2012. Regenerating sodium hydroxide from

368

the spent caustic by bipolar membrane electrodialysis (BMED) Separation and Purification

369

Technology 86, 49–54.

370

Wei, Y. Wang, Y., Zhang, X., Xu., T. 2013. Comparative study on regenerating sodium

371

hydroxide from the spent caustic by bipolar membrane electrodialysis (BMED) and electro-

372

electrodialysis (EED) Separation and Purification Technology 118, 1–5.

AC C

EP

TE D

M AN U

SC

RI PT

353

18

ACCEPTED MANUSCRIPT

Zhang, J. Zhang, B., Tian, C., Ye, Z., Liu, Y., Lei, Z., Huang, W., Feng, C. 2013. Simultaneous

374

sulfide removal and electricity generation with corn stover biomass as co-substrate in microbial

375

fuel cells. Bioresource technology 138, 198–203

RI PT

373

376 377

SC

378

M AN U

379 380 381

TE D

382 383

EP

384 385

AC C

386 387 388 389 390 19

ACCEPTED MANUSCRIPT

List of figure legends

392

Fig.1 Schematic diagram of the electrochemical cell. The cell was run in batch (proof-of-concept

393

experiments) and continuous (long term experiments) mode. Anode influent consisted of 4 wt%

394

NaOH and 1 wt% Na2S-S, and cathode influent was 4 wt% NaOH for the batch experiments and

395

distilled water for the continuous experiments.

396

Fig. 2 Sulfide removal in the anode at high pH in 8h batch experiments and CE of 75± 4%.

397

Fig. 3 Cell voltage evolution during the long term experiment. Without considering the

398

interruption period (day 49 - 104), the decrease in loading rate (day 25 - 34), and other single day

399

lab implications (day 47 and day 124), a stable cell voltage of 2.74 ± 0.10 V was maintained for

400

77 days of operation at 100 A m-2.

401

Fig. 4 (A) Impact of current density on sulfur speciation in the anolyte effluent. Influent SLR: 42

402

± 4 g(S) L-1d-1, flow rate: 121 ± 10 mL d-1. Higher current densities result in more oxidized

403

species and higher sulfide oxidation. (B) Effect of different sulfide loading rates 50 - 200 g(S) L-1

404

h-1 by a step of 50 g(S) L-1 h-1 on sulfur speciation at 100 A m-2. Higher loading rates result in less

405

sulfide removal, less oxidized sulfur species and induce polysulfide formation.

AC C

EP

TE D

M AN U

SC

RI PT

391

20

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Electrochemical spent caustic treatment allows economical product recovery



Current density and sulfide loading rate determine sulfide oxidation products



Sodium is recovered at high coulombic efficiencies and alleviates costs



Reactor materials remained unaffected during long term electrochemical operation

AC C

EP

TE D

M AN U

SC

RI PT



Electrochemical sulfide removal and caustic recovery from spent caustic streams.

Spent caustic streams (SCS) are produced during alkaline scrubbing of sulfide containing sour gases. Conventional methods mainly involve considerable ...
3MB Sizes 0 Downloads 6 Views