Adaptively Evolving Bacterial Communities for ... - ACS Publications

Jul 15, 2015 - These results demonstrated that bacterial communities can adaptively evolve ... Fan , Zhong-Yang Li , Zhen-Jie Du , Chao Hu , Andrew L...
0 downloads 0 Views 1MB Size
Page 1 of 32

Environmental Science & Technology

1

Adaptively evolving bacterial communities for complete and selective reduction of Cr(VI),

2

Cu(II) and Cd(II) in biocathode bioelectrochemical systems

3

Liping Huang1,*, Qiang Wang1, Linjie Jiang1, Peng Zhou2, Xie Quan1,*, Bruce E. Logan3

4 5

1. Key Laboratory of Industrial Ecology and Environmental Engineering, Ministry of Education (MOE),

6

School of Environmental Science and Technology, Dalian University of Technology, Dalian 116024,

7

China

8

2. College of Chemistry, Dalian University of Technology, Dalian 116024, China

9

3. Department of Civil and Environmental Engineering, The Pennsylvania State University, University

10

Park, Pennsylvania, 16802

11 12

Corresponding authors:

13

(Huang L.) [email protected], Tel./Fax.: 86 411 84708546

14

(Quan X.) [email protected], Tel.: 86 411 84706140

15 16 17 18 19

The authors declare no competing financial interest

20 1

ACS Paragon Plus Environment

Environmental Science & Technology

21

ABSTRACT

22

Bioelectrochemical systems (BESs) have been shown to be useful in removing individual metals

23

from solutions, but effective treatment of electroplating and mining wastewaters requires

24

simultaneous removal of several metals in a single system. To develop multiple-reactor BESs for

25

metals removal, biocathodes were first individually acclimated to three different metals using

26

microbial fuel cells with Cr(VI) or Cu(II) as these metals have relatively high redox potentials,

27

and microbial electrolysis cells for reducing Cd(II) as this metal has a more negative redox

28

potential. The BESs were then acclimated to low concentrations of a mixture of metals, followed

29

by more elevated concentrations. This procedure resulted in complete and selective metal

30

reduction at rates of 1.24 ± 0.01 mg/L-h for Cr(VI), 1.07 ± 0.01 mg/L-h for Cu(II), and 0.98 ±

31

0.01 mg/L-h for Cd(II). These reduction rates were larger than the no adaptive controls by

32

factors of 2.5 for Cr(VI), 2.9 for Cu(II) and 3.6 for Cd(II). This adaptive procedure produced less

33

diverse microbial communities and changes in the microbial communities at the phylum and

34

genus levels. These results demonstrated that bacterial communities can adaptively evolve to

35

utilize solutions containing mixtures of metals, providing a strategy for remediating wastewaters

36

containing Cr(VI), Cu(II) and Cd(II).

37 38

2

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

39

Environmental Science & Technology

TOC art

40 41

3

ACS Paragon Plus Environment

Environmental Science & Technology

42

INTRODUCTION

43

Heavy metals such as Cr(VI), Cu(II) and Cd(II) in electroplating and mining industry

44

wastewaters, if not removed, can severely contaminate surface waters, groundwaters, and soils.1-2

45

Electrochemical processes have high energy requirements and can require expensive catalysts to

46

decrease electrode overpotentials.2-3 Conventional biological processes can be used either

47

completely instead of, or coupled with traditional physical-chemical processes,4 but they have

48

disadvantages such as generation of high volumes of sludge, and/or the need for a continuous

49

source of organic carbon.5-6 Development of more environmentally-benign and less energy

50

demanding technologies would therefore be useful for treating these wastewaters.2-4

51

One new promising method for metal removal is the use of bioelectrochemical systems

52

(BESs).7-8 Cr(VI) and Cu(II) can be spontaneously reduced to Cr(III) and Cu(0) on the cathodes

53

of microbial fuel cells (MFCs) due to favorable half cell redox potentials of 1.14 and 0.34 V (vs.

54

standard hydrogen electrode, SHE) relative to that of organic matter (ca. -0.30 V for acetate,

55

standard conditions).9-10 Even under the condition of an approximately neutral pH (eg. 6.0)

56

suitable bacterial survival, the pH-dependent reduction of Cr(VI) has an half cell redox potential

57

of 0.64 V (vs. SHE), and therefore it can spontaneously occur in MFCs. Cd(II) has been shown

58

to be adsorbed to the cathode in single-chamber, air-cathode MFCs11 but it can be chemically

59

reduced in microbial electrolysis cells (MECs) at a more negative potential of -0.52 V (vs. SHE,

60

standard conditions).12-13 While the newly developed MFC-MEC coupled systems show

61

promising for mixed metals under extreme acidic conditions,13-14 the effluents still contain 4

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32

Environmental Science & Technology

62

concentrations of metals that are too high to be discharged into environment. Thus, the process

63

needs to be further improved or additional treatment would still be required. The use of

64

biocathodes in these different BESs to reduce electrode overpotentials avoids the need for

65

expensive inorganic catalysts in purely electrochemical processes, and the microorganisms are

66

self-regenerating and therefore their use is sustainable.7,9,15-16 While biocathodes have been

67

shown to be useful for removing individual metals such as Cr(VI), U(V) and Co(II),17-22 there

68

have been no tests on removing mixtures of different metals. In addition, metal concentrations in

69

BES effluents have been above drinking water limits established by the World Health

70

Organization (WHO) and the US EPA, which are 0.05 mg/L for Cr, 1.0 mg/L for Cu and 0.005

71

mg/L for Cd.20,23-24 The use of microorganisms to reduce a mixture of metals is not trivial as

72

mixed metals can be synergetically more toxic than the additive effect of the individual

73

metals.25-28

74

In this study, biocathodes were first developed in MFCs individually acclimated to Cr(VI)

75

(MFC-Cr) or Cu(II) (MFC-Cu), and in MECs to Cd(II) (MEC-Cd). The microbial communities

76

on these biocathodes were then adaptively evolved to become acclimated to a mixture of these

77

chemicals by initially using low concentrations of these metals, followed by higher

78

concentrations, to achieve nearly complete removal of these metals. The mixed metal solutions

79

were sequentially fed through the first two MFCs for Cr(VI) and Cu(II) removal, and then into

80

the MECs for Cd removal. Species of Cr(VI), Cu(II) and Cd(II) were chosen here mainly due to

81

their frequent observation in electroplating wastewaters, although the specific concentrations of 5

ACS Paragon Plus Environment

Environmental Science & Technology

82

these metals in electroplating wastewaters varies depending on the facility.28-29 The performance

83

of this three-reactor BES system was evaluated in terms of metals and organics removal, biomass

84

production, gas (hydrogen and methane) or power generation, and coulombic efficiencies (CEs).

85

High-throughput 16S rRNA gene sequencing was used to characterize the microbial

86

communities on the biocathodes of the BESs of the original cathodes acclimated to only one

87

metal, compared to biocathodes acclimated to the mixture of metals.

88 89

MATERIALS AND METHODS

90

Reactor Setup. Duplicate tubular two-chamber reactors19 were used in all experiments and for

91

each of the duplicate reactors three replicate experiments were performed. The MFCs were

92

operated at a fixed external resistance of 510 Ω, while a voltage of 0.5 V was added to the circuit

93

of the MECs using a power source (DC Power Supply PS-1502DD, Yihua, Guangzhou, China).

94

The anodes were made of graphite brushes (PANEX33 160 K, ZOLTEK),30 and the cathodes

95

were porous graphite felt (Sanye Co., Beijing, China).19 The anode (25 mL working volume) and

96

cathode (40 mL) chambers were separated by a cation exchange membrane (CEM) (CMI-7000

97

Membranes International, Glen Rock, NJ) with a projected surface area of 47 cm2.

98

Inoculation and Operation. Anodes were inoculated using the effluent of MFCs well

99

acclimated to acetate that were originally inoculated using primary clarifier effluent from the

100

Lingshui Wastewater Treatment Plant (Dalian, China). Anodes were fed with a phosphate buffer

101

(5 mM NaH2PO4) medium containing acetate (1.0 g/L).19 For the initial 3 - 4 fed-batch 6

ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32

Environmental Science & Technology

102

acclimation cycles with each lasting 24 – 36 h, cathodes were fed using the same phosphate

103

buffer medium prepared in domestic wastewater, except acetate was replaced by NaHCO3 (1.0

104

g/L) to provide a carbon source for the autotrophic microorganisms, with metals added to the

105

individual reactors at a concentration of 5 mg/L [Cr(VI) or Cu(II) in MFCs, and Cd(II) in MECs].

106

After the biofilm was fully acclimated, each fed batch cycle time was set at 4 h. Analytical

107

reagents of K2Cr2O7, CuCl2 and CdCl2 were used to prepare these metal solutions. A total of

108

96.7% of Cr2O72- was theoretically present as HCrO4- in this catholyte. Wastewater was then

109

omitted from the catholyte medium in all subsequent experiments.

110

After reproducible cycles of performance in terms of removal of the metals by the

111

biocathodes [acclimation time reaching 25 days for both MFC-Cr and MFC-Cu, and 10 days for

112

MEC-Cd], the reactor operation was switched to sequential flow of catholyte containing a

113

mixture of metals through successive cathode chambers. During this acclimation stage, the

114

solution contained 3.0 mg/L of Cr(VI), 1.0 mg/L of Cu(II), and 1.0 mg/L of Cd(II). This solution

115

was first used in MFC-Cr(VI), and then at the end of the batch cycle it was transferred into the

116

cathode of MFC-Cu(II), followed by transfer into MEC-Cd(II). After multiple cycles using this

117

lower concentration metals mixture, the concentrations of each of the metals were increased to 5

118

mg/L, and the acclimation cycles were continued. These biofilms were considered to be fully

119

acclimated when metal removals were reproducible over 4 - 6 fed batch cycles.

120 121

Two controls were used to evaluate biocathode performance. The first was operation of the reactors under open circuit conditions (OCC) and the second was operation with a closed circuit 7

ACS Paragon Plus Environment

Environmental Science & Technology

122

but without inoculation of the cathodes (CCC).31 Prior to adding the solutions into the electrode

123

chambers, the anolyte and catholyte were sparged with N2 gas for 15 min. All inoculation and

124

solution replacements were performed in an anaerobic glove box (YQX-II, Xinmiao, Shanghai),

125

and all experiments were run at room temperature (20 ± 3 ºC). The initial pH of all catholytes

126

was 5.8, and the solution conductivities in both anolyte and catholyte were adjusted to be 5.8

127

mS/cm using KCl.

128

Measurements and Analyses. The total chemical oxygen demand (COD), metals

129

concentrations (Cr(VI), Cu(II), and Cd(II)), hydrogen and methane concentrations in the

130

headspaces of cathode chambers of the MECs, and soluble COD and biomass in the catholyte

131

were measured as previously described.,19,22 as summarized in the Supplementary Information

132

(SI). Morphologies of the electrodes after Cr(VI), Cu(II) and Cd(II) reduction were examined

133

using a scanning electronic microscopy (SEM) (QUANTA450, FEI company, USA) equipped

134

with an energy dispersive spectrometer (EDS) (X-MAX 20 mm2/50 mm2, Oxford Instruments,

135

UK). Samples for SEM-EDS analysis were prepared as described in SI. Cathode and anode

136

potentials, cyclic voltammetry (CV) and linear sweep voltammetry (LSV) analysis were

137

performed as previously described19,22 and summarized in the SI. Power density is often

138

normalized by the electrode projected surface area. In the MFCs and MECs used here, both the

139

graphite brush anodes and porous graphite felt cathodes had very high and much different

140

surface areas, but the separator (or membrane) is known to often limit power production in

141

MFCs.32 Therefore, power was normalized to the projected surface area of the separator, 8

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32

Environmental Science & Technology

142

allowing comparison to previous studies based on power per area,32 in addition to providing

143

absolute values. All the potentials reported were given in relative to a standard hydrogen

144

electrode (SHE).

145

Bacterial Community Analysis. Samples were collected from the biocathodes of the BESs

146

at the end of experiments acclimated to single metal, and the mixture of metals at an identical 5

147

mg/L. Since values of metal reduction rate, circuit current and power production in the duplicate

148

reactors were highly similar, five pieces of graphite felt (1×1×1 cm3) were removed from

149

multiple locations on the electrodes in each of the duplicate reactors and mixed together for

150

analysis. Bacterial community results from this sample thus reasonably reflected those in the

151

duplicate reactors. These electrodes were fragmented using sterile scissors. Cells attached on the

152

electrodes were removed by rinsing three times with sterile water, and concentrated by

153

centrifugation. DNA was extracted using Qubit2.0 DNA kit (Sangon Biotech (Shanghai) Co.

154

Ltd., China) according to the manufacturer’s instructions. The quality (A260/A280) and quantity

155

(A260) of the extracted genomic DNA were determined with a Nanodrop® 1000

156

spectrophotometer (Thermo Fisher Scientific, Waltham, MA). The extracted DNA from each

157

sample was amplified using PCR reactions and universal 16S rRNA primers (Illumina Miseq).

158

After the PCR amplification, the amplicons were recovered using a gelose extraction kit (Cat:

159

SK8131, Sangon Biotech (Shanghai) Co. Ltd., China), and rinsed with Tris-HCl solution. After

160

amplification and quantification, the amplicons were sequenced using an Illumina Miseq

161

following standard procedures. The data were optimized through removal of low-quality 9

ACS Paragon Plus Environment

Environmental Science & Technology

162

sequences, unrecognized reverse primers, and any ambiguous base calls, with a length < 200 bp.

163

High quality sequences were aligned, grouped into OTUs (97% similarity) using the uclust

164

algorithm, and a representative sequence from each OTU was classified phylogenetically

165

assigned to a taxonomic identity (phylum, class and genus level) using the RDP Naïve Bayesian

166

rRNA classifier at a confidence threshold of 90%.33 Beta diversity metrics using weighted

167

UniFrac distance matrix34 were calculated from the phylogenetic tree and visualized with

168

principal coordinate analysis in quantitative insights into microbial ecology (QIIME software

169

v1.3.0) pipeline.35 Rarefaction curves, Shannon diversity indices and species richness estimators

170

of Chao1 were generated for each sample using QIIME. All analyses described were conducted

171

in triplicates, and the means were reported. The sequence data will be provided upon request.

172 173

Calculations. Cr(VI), Cu(II) and Cd(II) reduction, anodic CEan, cathodic CEca, and yields of metals, biomass, organics, hydrogen and methane were calculated as described in SI.

174 175

RESULTS AND DISCUSSION

176

Acclimation of BESs to Individual Metals. Following 25 d of acclimation (details shown in

177

Table S1), biocathodes developed in the MFCs produced a maximum of 48 µW (369 µA) (Figure

178

S1A) based on a current of 220 µA (510 Ω external resistance) (Figure S1B) with a Cr(VI)

179

catholyte, and a maximum of 27 µW (161 µA) (Figure S1C) and 166 µA at 510 Ω (Figure S1D)

180

with a Cu(II) catholyte. Each of these MFCs demonstrated efficient metals removal, with 1.21 ±

181

0.02 mg/L-h for Cr(VI), and 1.18 ± 0.02 mg/L-h for Cu(II) (Figure S1E). The circuit current in 10

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32

Environmental Science & Technology

182

the MEC used to treat Cd(II) reached 994 µA after 10 days (Figure S1F), with Cd(II) removal at

183

a rate of 1.15 ± 0.01 mg/L-h (Figure S1E). This removal rate for Cd(II) was higher than the 0.78

184

mg/L-h previously obtained at a similar pH of 6.0,12 but lower than the rate of 3.58 mg/L-h at a

185

higher solution conductivity 8.6 mS/cm and lower pH 2.0.13 Both of the reports were at a much

186

higher initial Cd(II) concentration of 50 mg/L, and used only abiotic cathodes. Rates of Cd(II)

187

removal would be expected to increase in proportion to concentration and by using higher

188

solution conductivities. No measurable concentrations of Cr(III), Cu(II) and/or Cd(II) were

189

found in the anolyte after a fed batch cycle, excluding the possibility of metal diffusion from the

190

cathode to the anode. However, this does not preclude metal retention in the ion exchange

191

membrane. CEca reached 92 ± 2% in MFC-Cr and 84 ± 2% in MFC-Cu (Tab.S2), with 60% of

192

this current used for Cr(VI) reduction in the MFC-Cr, and 50% for Cu(II) reduction in the

193

MFC-Cu. The remaining 32% of the electrons for tests with the MFC-Cr, and 34% for the

194

MFC-Cu, were likely lost primarily to oxygen reduction. Despite sparging with ultra pure N2 for

195

15-20 min, there was a residual dissolved oxygen concentration of 0.7 – 1.0 mg/L in the initial

196

catholyte. This dissolved oxygen could have accounted for 20 - 26% of the coulombs in the

197

MFC-Cr, and 28 - 39% in MFC-Cu.36-38

198

CV Analyses on Biocathodes Developed with Single Metals. There were different

199

predominant reductive peak potentials in the CVs of biofilms using the three different metals,

200

with 0.285 V for Cr(VI), 0.075 V for Cu(II), and -0.410 V for Cd(II) (Figures 1A, B and C). In

201

each case, these reductive peak potentials were more positive than those of the abiotic controls 11

ACS Paragon Plus Environment

Environmental Science & Technology

202

despite not excluding the possibility of oxygen evolution from the electrode at high potentials,

203

showing the biofilms catalytic activity toward metals reduction. Reductive peak potential was an

204

effective parameter for assessing bacterial catalytic activities in the biocathodes.37,39-41 Reductive

205

onset potentials (where current values were changed from positive to negative) on the biotic

206

cathodes were also more positive than the abiotic controls for all the three different metals (0.730

207

V vs 0.685 V for Cr(VI), 0.240 V vs 0.215 V for Cu(II), and -0.205 V vs -0.210 V for Cd(II))

208

(inserts in Figure 1). A more positive reductive onset potential on the biotic electrode indicated a

209

decrease in the overall free energy of the electron transfer reaction, mainly due to bacterial

210

interactions with the electrode surface, which appeared to decrease the energy required for

211

Cr(VI), Cu(II) and Cd(II) reduction.42-43 A lower reductive peak current on the biotic cathode,

212

however, suggests some degree of mass transfer inhibition due to both the bacterial attachment to

213

the electrode surface42 and the formation of these metal precipitants on the electrode surface or

214

within the microorganisms. The speciation and location of these metal precipitants may be

215

dependent on the specific electrotrophs on the cathodes. Determining the exact fate of these

216

metals would require more defined conditions, such as pure culture tests. Cr(VI) would be

217

expected to be reduced to insoluble Cr(OH)2+ and Cr(OH)2+, in addition to Cr(OH)3, under the

218

present experimental conditions,44-46 which would explain lower biotic peak currents compared

219

to abiotic controls. The positive effects of both more positive reductive onset potential and more

220

positive peak potentials, together with the negative more mass transfer inhibition on the biotic

221

cathode thus presumably resulted in the more efficient Cr(VI), Cu(II) and Cd(II) reductions than 12

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32

Environmental Science & Technology

222

those in the abiotic controls. Other researchers also observed similar results, where biotic

223

cathodes for hydrogen evolution and Au(III) or Co(II) recovery had lower reductive peak

224

currents than abiotic cathodes despite the biotic cathodes having more positive potentials.42,47

225

Current density based on only the biofilm-covered area (which can only roughly estimated by

226

microscopic examination) was recently suggested as a relatively good method for assessing BES

227

performance, particularly for electrodes of porous graphite felt as used here.48 However, the

228

biofilm area could not be determined here as it was very sparse, and the available area could also

229

be impacted by the area occupied by metal precipitates. Considering the usually sparse coverage

230

of bacteria on the cathodes and the area occupied by the metal precipitates as well as the

231

presence of non-electrochemically active microbes in the mixed cultures, the actual current

232

density based on effective electrochemically active biofilm-covered area was presumably higher

233

than that of the abiotic controls, despite the somewhat higher reduction peak currents (in terms of

234

mA) in the abiotic controls (Figures 1A, B and C).

235

The peak currents were examined at different scan rates (ν) (1 - 10 mV/s) to determine the

236

rate limiting step.49 When a plot of the peak current versus ν is linear, the results indicate

237

“thin-film behavior”, and therefore that electron transfer from the electrode to the bacteria is

238

slower than intracellular reactions. If the peak is current proportional to ν1/2, this indicates a

239

diffusion-controlled regime.37,50-51 Plots of peak current ν were linear for all three metals (Figures

240

1D, E and F), demonstrating that current generation was not limited by diffusion for these

13

ACS Paragon Plus Environment

Environmental Science & Technology

241

biocathodes. The gradients of the scans decreased in the order of Cr>Cu>Cd, indicating different

242

kinetic behavior for microorganisms using these three metals.37,51

243

Morphologies of Cathode Surfaces and Product Analysis with Single Metals. The

244

cathodes of BESs were sparsely covered by bacteria, with precipitates observed in association

245

with groups of cells (Figures S2A, B, and C). Sparse biofilms were consistent with other reports

246

of biocathode biofilms,19,40-41,52 with this morphology in contrast to the relatively thick biofilms

247

that develop on bioanodes. EDS analysis confirmed the presence of Cr, Cu and Cd in these

248

precipitates in the respective BESs (Figures S2D, E, and F). In addition, the presence of carbon,

249

oxygen, phosphorus and sulfur may be attributed to the coating of the metal precipitates on the

250

microorganisms and the microorganisms on the graphite felt, while gold can be likely traced

251

back to sample pretreatment. The observations of carbon and oxygen in the absence of bacteria

252

(Figures S2G, H and I) reflected that metals removal was mostly related with the bacteria. A 48 h

253

abiotic control experiment demonstrated there were no changes in the concentrations of Cr(VI),

254

Cu(II) and Cd(II) in the solution, excluding the formation of precipitates between metals and

255

anions in the catholyte. The absence of carbon and sulfur elements in the metal precipitates-only

256

samples (Figures S2J, K, and L) further precluded the formation of metal sulfide and carbonate

257

metal complexes in the present system. The ratio of elemental metal and oxygen in EDS can be

258

used to deduce the product composition.10,53-55 The Cr:O ratio in the precipitates was 0.30

259

(Figure S2J), approaching that of 0.33 for a precipitate of Cr(OH)3. The Cu:O ratio was 10

260

(Figure S2K), and the Cd:O was 7 (Figure S2L), which were much higher than those of the final 14

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32

Environmental Science & Technology

261

products of 2 or 1 for Cu2O or CuO, and 1 for CdO. These ratios therefore show that Cu(II) and

262

Cd(II) were mostly reduced to the pure metals. Based on the final pHs that ranged from 5.9 – 6.1,

263

the calculated forms of the metals using the Nernst equation would be Cr(III), Cu(0), and Cd(0)

264

(Figure S3). The speciation of chromium at different potentials was extensively reported to be

265

reductively changed into precipitates of Cr(OH)2+, Cr(OH)2+ and Cr(OH)3 at pHs of 5.9 –

266

6.1.44-46 In the case of Cr(VI) conversion in the biocathode/abiotic cathode bioelectrochemical

267

systems, as summarized in Table S3, all Cr(VI) was reductively changed into Cr(III). Similarly,

268

Cu(II) was mainly reduced to Cu(0) with tiny Cu2O whereas Cd(II) was transformed into Cd(0)

269

(Table S3). Our results were therefore consistent with these previous reports on the speciation of

270

these metals at these pHs.

271

Biocathodes Adaptively Evolved to Low and Elevated Mixed Metals. Mixed metal

272

solutions containing Cr(VI), Cu(II) and Cd(II), at each 5 mg/L, was sequentially fed into the

273

MFCs or MECs individually acclimated only to a single metal. Reduction rates [0.49 ± 0.03

274

mg/L-h for Cr(VI); 0.37 ± 0.05 mg/L-h for Cu(II); and 0.27 ± 0.03 mg/L-h for Cd(II)] were all

275

much lower compared to rates when using only the individual metals (Figure S1C). It is well

276

known that combinations of heavy metals are generally more toxic than single heavy metals.25-28

277

The apparent decrease in metal reduction rates in response to the mixed metal influents reflects

278

the detrimental impact of the other metals on the bacterial communities, in good agreement with

279

previous reports.25-28

15

ACS Paragon Plus Environment

Environmental Science & Technology

280

The reactors were next acclimated to low concentrations of these metals (details in SI, and

281

Figures S4 and 5), and then switched to higher concentrations of 5 mg/L of each metal for 11 –

282

15 cycles. Metal reduction rates were greatly improved for Cr(VI), as there was complete

283

removal to below detectable limits (< 0.04 µg/L) at a rate of 1.24 ± 0.01 mg/L-h (Figure 2A).

284

This rate was 2.5 times of that without adaptive evolution (0.49 ± 0.03 mg/L-h) and slightly

285

lower than the rate obtained using the lower concentration mixed metal solution (1.48 ± 0.02

286

mg/L-h) (Figure S4A), but comparable to that obtained with the single metal solution (1.21 ±

287

0.02 mg/L-h) (Figure S1C), demonstrating successful adaptation to the presence of elevated

288

mixed metals. There was little removal of Cu(II) or Cd(II) in this reactor or in the controls

289

(OCC-biotic and CCC-abiotic, 0.12 ± 0.01 - 0.22 ± 0.02 mg/L-h). Metal removals in the

290

CCC-abiotic controls were mainly ascribed to chemical reduction and physical adsorption56

291

whereas observation in the OCC-biotic controls included processes associated with the presence

292

of bacteria, bio-adsorption in addition to physical adsorption, and biological reduction of metals.

293

While adsorption may have contributed to metal removal in this study, it is not expected to be

294

large here due to the sparse biofilms that typically develop on biocathodes, and there was

295

relatively little metal removal in abiotic controls. Although there were decreases in the OCP and

296

maximum power produced by the MFC originally acclimated to only Cr(VI) (Figures 2D and

297

S4D), the high rate of Cr(VI) reduction demonstrated the successful acclimation of the

298

community relative to Cr(VI) removal.

16

ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32

299

Environmental Science & Technology

The biocathodes in the subsequent reactors also showed good acclimation to the elevated

300

mixed metal solutions. In the second MFC [originally acclimated to only Cu(II)], complete Cu(II)

301

removal was observed at a rate of 1.07 ± 0.01 mg/L-h, with very little reduction of Cd(II) (Figure

302

2B). This reduction rate for Cu(II) was 2.9 times of that without adaptive evolution (0.37 ± 0.05

303

mg/L-h), 2.0 times of that with a low concentration of mixed metals (0.53 ± 0.02 mg/L-h)

304

(Figure S4B), and a slightly lower rate than that obtained with the reactors acclimated to only Cu

305

(II) (1.18 ± 0.02 mg/L-h) (Figure S1C). The OCP of 0.34 V and maximum power of 26 µW (188

306

µA) (Figure 2E) were similar to those obtained with the reactor acclimated to only Cu(II) (0.33 V,

307

and 28 µW) (Figure S1B).

308

In the third reactor, Cd(II) was also completely removed at a rate of 0.98 ± 0.01 mg/L-h

309

(Figure 2C), which was 3.6 times of that without any adaptive evolution (0.27 ± 0.03 mg/L-h),

310

but slightly lower than that for the MEC acclimated to only Cd(II) (1.15 ± 0.01 mg/L-h). The

311

Cd(II) reduction rates were also greater than those obtained in either of the controls (CCC-abiotic

312

and OCC-biotic, 0.51 ± 0.06 – 0.89 ± 0.03 mg/L-h). The measured cathode potential in the

313

MECs of -0.61 V (Figure 2F) was identical with that measured using the low concentration

314

mixed metal solution (Figure S4F).

315

Compared to results with a low concentration of mixed metals, an elevated mixed metal

316

concentration was beneficial not only for anodic CEs in all three reactors, but also product yields

317

in the MFC that removed Cu (Table S2). Metal reduction rates in the biocathode reactors could

318

not be directly compared with conventional electrochemical processes due to greatly different 17

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 32

319

operational conditions of much high temperatures of 70 – 90 ºC and more acidic pHs of 1.0 – 2.0,

320

both of which reasonably favored metal reduction.3,57-60 However, apparently higher metal

321

reduction rates in the biocathode reactors were observed than the CCC-abiotic controls,

322

indicating the higher performance of biocathodes than a conventional electrochemical process

323

without microorganisms. The energy produced by the MFCs (0.6 kWh/kg-Cr and 0.3

324

kWh/kg-Cu) was lower than the power required for the Cd removal in the MEC (10.5

325

kWh/kg-Cd). Considering the hydrogen energy of 2.8 kWh/kg-Cd concomitantly produced in

326

MEC for Cd(II) reduction, the overall power required to operate these reactors would only be 6.8

327

kWh/kg-metal. This is substantially lower than the 73 - 113 kWh/kg-metal used for purely

328

electrochemical processes.57-60

329

CV results indicated slight increases in reductive peak currents in all the reactors acclimated

330

to the elevated concentrations of mixed metals (Figures 3A, B and C), relative to those for the

331

low concentrations of mixed metals. This result illustrates the positive link between peak current

332

and substrate concentration, which always affect the electron transfer processes observed in

333

CVs.42,47-48 These changes were consistent with results showing improved metals removals

334

(Figure 2). Current was a linear function of the scan rate for all the three biofilms with these

335

elevated mixed metals (Figures 3D, E and F), demonstrating a lack of diffusional mass transfer

336

limitations in these systems.

337 338

Changes in Bacterial Communities. From six generated libraries, a total of 23254 high-quality 16S rRNA gene sequences were obtained with an average length of 440 nucleotides 18

ACS Paragon Plus Environment

Page 19 of 32

Environmental Science & Technology

339

(Table S4). These sequences were assigned into 2235 OTUs with a distance limit of 0.03.

340

Although the rarefaction curves did not demonstrate a plateau (Figure S6), the Good’s Coverage

341

estimators indicated that the sizes of libraries were sufficient to cover 93 - 97% of the bacterial

342

communities (Table S4).

343

An analysis of these data show that the diversity of the microbial communities decreased,

344

and that different bacterial communities were identified on the biocathodes of the reactors

345

acclimated to the mixed metal solutions, compared to reactors acclimated only to a single metal.

346

The decrease in diversity was shown by a decrease in the Shannon indices from 4.62 to 4.22 for

347

Cr(VI), 4.60 to 4.15 for Cu(II), and 4.08 to 3.86 for Cd (II) (Table S4). These decreases in

348

diversity likely resulted from the greater toxicity of the mixed metals compared to only single

349

metals.26-28

350

Analysis of the bacterial communities based on a Weighted Fast UniFrac Principle

351

Coordinates Analysis, indicated little similarity in the communities acclimated to a single metal

352

with those acclimated to the mixtures of metals (Figure 4A). Only the MECs acclimated to Cd(II)

353

alone, or the Cd(II)-acclimated reactor subsequently acclimated to the mixed metals, grouped in

354

the same quadrant. The reactors acclimated to the mixed metals were also relatively randomly

355

dispersed. A dendrogram constructed on the basis of community phylogenetic lineages further

356

supported differences in these bacterial communities based on an absence of any close

357

alignments of the communities acclimated to single metals with those acclimated to mixtures of

358

metals (Figure 4B). These findings demonstrate that the compositions of the bacterial 19

ACS Paragon Plus Environment

Environmental Science & Technology

359

communities were initially different from each other when acclimated to the individual metals,

360

and that they further diverged after acclimation to the metal mixtures.

361

The bacterial community consisted of 8 phyla, with the majority of sequences (21.3 - 72.0%)

362

belonging to Proteobacteria, which are frequently observed in BESs (Figure 5A).61-62 Compared

363

to the respective single metal acclimated evolved biofilms, the relative abundance of

364

Proteobacteria decreased in reactors acclimated to mixed metals that were originally acclimated

365

to Cu and Cd, whereas Proteobacteria and Actinobacteria increased in reactors originally

366

acclimated to only Cr. The acclimation to the mixed metals stimulated the relative abundance of

367

Firmicutes and Tenericutes in all reactors, and Bacteroidetes in the MFC originally acclimated to

368

Cu (Figure 5A). Conversely, Synergistetes was absent in all communities following acclimation

369

to the mixed metals.

370

Four classes (Alpha-, Beta-, Gamma-, and Delta-) within Proteobacteria were observed from

371

all communities except for Deltaproteobacteria in the MFCs originally acclimated to Cr (Figure

372

5B). Mixed metal acclimation led to a relative increase in Alphaproteobacteria.

373

Betaproteobacteria were the most abundant in all the communities with single metals and the

374

biofilm of MFC with the mixed metal solution that was originally acclimated to Cr.

375

Genus level characterization suggests that bacteria most similar to Geobacter were relatively

376

abundant on biocathodes acclimated only to Cr (10.7%) or Cu (10.1%) (Figure 6). In the abiotic

377

controls, no acetate was observed in the catholyte, suggesting that there was limited or no

378

diffusion of acetate from the anode to cathode at the beginning of the experiment that could have 20

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32

Environmental Science & Technology

379

sustained substantial Geobacter growth. However, the organics formed in the catholyte (Table

380

S2) could have allowed for heterotrophic growth of Geobacter, as this genus is not known to be

381

autotrophic.61-62 The predominant genus in the biocathodes of MFC acclimated to Cr was

382

Dysgonomonas (10.2%), but Azoarcus (16.7%) were relatively abundant in these reactors

383

following acclimation to the mixed metal solution. Similarly, dominant genera were shifted from

384

Geobacter (10.1%) to Myroides (10.8%) and an unclassified bacterium (10.1%) for reactors

385

switched from only Cu to mixed metals. In the case of Cd, Achromobacter (16.8%) and Brucella

386

(15.7%) were predominantly observed in biofilm with treating only Cd, compared to Alcaligenes

387

(16.5%), Tissierella (14.7%), Brevundimonas (10.0%) and Aquamicrobium (9.4%) with mixed

388

metal solutions. While ecological roles of these bacteria cannot be ascertained based on their

389

presence, these changes do show that the switch from a single metal to the metals mixtures

390

altered the composition of the microbial communities.

391

The use of biocathodes in BESs shows promise as a low energy and sustainable method for

392

remediation of toxic metals in wastewaters. These studies have shown that biocathode BESs can

393

be used to remediate Cr(VI), Cu(II) and Cd(II) individually, or together in mixtures, using

394

cathodes of MFCs and MECs. Following acclimation to a single metal, the microbial populations

395

in reactors exhibited adaptive evolution whereby the original target metals could be completely

396

removed at rates similar to that originally obtained with the reactors acclimated to a single metal.

397

In all cases, the reactors did not show acclimation for removal of the other target metals,

398

providing an effective strategy for sequential and selective recovery of chromium, copper and 21

ACS Paragon Plus Environment

Environmental Science & Technology

399

cadmium from mixed wastes. Community analysis showed that the bacterial compositions were

400

significantly different after the elevated mixed metal influents, and generally they had less

401

diverse communities following exposure to the mixed metal solutions. Although there was no

402

clear identification of the specific genera responsible for the metals reduction, as the

403

communities remained relatively diverse, future identification of specific bacterial species could

404

help to improve our understanding of the operation of these systems. Such information could

405

make it possible to acclimate biocathodes in a single reactor to remove and recover the multiple

406

metals, when those metals have similar half cell potentials for reduction.

407

Compared to the biocathodes with only Cr(VI) reduction and abiotic cathodes under extreme

408

acidic conditions for Cr(VI), Cu(II) and/or Cd(II) reduction (Table S3), the present study

409

suggests it should be possible to treat neutral pH and low-strength wastewaters from

410

electroplating and mining industry processes. Practical implementation will depend on the

411

long-term operation of the biocathodes and other characteristics of the wastewaters, as well as

412

the cost of the system relative to conventional treatment processes. The costs of the materials

413

used in MFCs and MECs are decreasing,63-65 but at this point in time it is likely that this

414

technology is not yet ready for commercial applications as larger-scale studies would be required

415

for evaluating the process prior to commercial and full scale applications.

416

ACKNOWLEDGEMENTS

417

The authors gratefully acknowledge financial support from the National Natural Science

418

Foundation of China (No. 21377019), the National Basic Research Program of China (No. 22

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32

Environmental Science & Technology

419

2011CB936002), Specialized Research Fund for the Doctoral Program of Higher Education

420

“SRFDP” (No. 20120041110026), the Program for Changjiang Scholars and Innovative

421

Research Team in University (IRT_13R05) and the Programme of Introducing Talents of

422

Discipline to Universities (B13012).

423

Supporting Information Available

424

Information as mentioned in the text. This information is available free of charge via the Internet

425

at http://pubs.acs.org.

426 427

REFERENCES

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

(1) Soares, E. V.; Soares, H. M. V. M. Bioremediation of industrial effluents containing heavy metals using brewing cells of Saccharomyces cerevisiae as a green technology: a review. Environ. Sci. Pollution Res. 2012, 19, 1066-1083. (2) García, V.; Häyrynen, P.; Landaburu-Aguirre, J.; Pirilä, M.; Keiski, R. L.; Urtiaga, A. Purification techniques for the recovery of valuable compounds from acid mine drainage and cyanide tailings: application of green engineering principles. J. Chem. Technol. Biotechnol. 2014, 89, 803-813. (3) Freitas, M. B. J. G.; Garcia, E. M. Electrochemical recycling of cobalt from cathodes of spent lithium-ion batteries. J. Power Sources 2007, 171, 953-959. (4) Johnson, D. B. Recent developments in microbiological approaches for securing mine wastes and for recovering metals from mine waters. Minerals 2014, 4, 279-292. (5) Heinzel, E.; Hedrich, S.; Janneck., E.; Glombitza, F.; Seifert, J.; Schlömann, M. Bacterial diversity in a mine water treatment plant. Appl. Environ. Microbiol. 2009, 75, 858-861. (6) Ňancucheo, I.; Hedrich, S.; Johnson, D.B. New (micro-) biological strategies that enable the selective recovery and recycling of metals from acid mine drainage and mine process waters. Mineral. Mag. 2012, 76, 2683-2692. (7) Rosenbaum, M. A.; Franks, A. E. Microbial catalysis in bioelectrochemical technologies: status quo, challenges and perspectives. Appl. Microbiol. Biotechnol. 2014, 98, 509-518. (8) Li, W., Yu, H., He, Z. Towards sustainable wastewater treatment by using microbial fuel cells-centered technologies. Energy Environ. Sci. 2014, 7, 911-924. (9) Huang, L.; Cheng, S.; Chen, G. Bioelectrochemical systems for efficient recalcitrant wastes treatment. J. Chem. Technol. Biotechnol. 2011, 86, 481-491. (10) Cheng, S.; Wang, B.; Wang, Y. Increasing efficiencies of microbial fuel cells for collaborative treatment of copper and organic wastewater by designing reactor and selecting operating parameters. Bioresour. Technol. 2013, 147, 332-337. (11) Abourached, C.; Catal, T.; Liu, H. Efficacy of single-chamber microbial fuel cells for removal of cadmium and zinc with simultaneous electricity production. Water Res. 2014, 51, 228-233. (12) Choi, C.; Hu, N.; Lim, B. Cadmium recovery by coupling double microbial fuel cells. 23

ACS Paragon Plus Environment

Environmental Science & Technology

457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510

Page 24 of 32

Bioresour. Technol. 2014, 170, 361-369. (13) Zhang, Y.; Yu, L.; Wu, D.; Huang, L.; Zhou, P.; Quan, X.; Chen, G. Dependency of simultaneous Cr(VI), Cu(II) and Cd(II) reduction on the cathodes of microbial electrolysis cells self-driven by microbial fuel cells. J. Power Sources 2015, 273, 1103-1113. (14) Luo, H.; Qin, B.; Liu, G.; Zhang, R.; Tang, Y.; Hou, Y. Selective recovery of Cu2+ and Ni2+ from wastewater using bioelectrochemical system. Front. Environ. Sci. Eng. 2015, DOI: 10.1007/s11783-014-0633-5 (15) Pisciotta, J. M.; Zaybak, Z.; Call, D. F.; Nam, J. Y.; Logan, B. E. Enrichment of microbial electrolysis cell biocathodes from sediment microbial fuel cell bioanodes. Appl. Environ. Microbiol. 2012, 78, 5212-5219. (16) Huang, L.; Chai, X.; Quan, X.; Logan, B. E.; Chen, G. Reductive dechlorination and mineralization of pentachlorophenol in biocathode microbial fuel cells. Bioresour. Technol. 2012, 111, 167-174. (17) Gregory, K. B.; Lovley, D. R. Remediation and recovery of uranium from contaminated subsurface environments with electrodes. Environ. Sci. Technol. 2005, 39, 8943-8947. (18) Tandukar, M.; Huber, S. J.; Onodera, T.; Pavlostathis, S. G. Biological chromium(VI) reduction in the cathode of a microbial fuel cell. Environ. Sci. Technol. 2009, 43, 8159-8165. (19) Huang, L.; Chai, X.; Chen, G.; Logan, B. E. Effect of set potential on hexavalent chromium reduction and electricity generation from biocathode microbial fuel cells. Environ. Sci. Technol. 2011, 45, 5025-5031. (20) Hsu, L.; Masuda, S. A.; Nealson, K. H.; Pirbazari, M. Evaluation of microbial fuel cell Shewanella biocathodes for treatment of chromate contamination. RSC Adv. 2012, 2, 5844-5855. (21) Xafenias, N.; Zhang, Y.; Banks, C. J. Enhanced performance of hexavalent chromium reducing cathodes in the presence of Shewanella oneidensis MR-1 and lactate. Environ. Sci. Technol. 2013, 47, 4512-4520. (22) Huang, L.; Jiang, L.; Wang, Q.; Quan, X.; Yang, J.; Chen, L. Cobalt recovery with simultaneous methane and acetate production in biocathode microbial electrolysis cells. Chem. Eng. J. 2014, 253, 281-290. (23) Cazón, J. P.; Bernardelli, C.; Viera, M.; Donati, E.; Guibal, E. Zinc and cadmium biosorption by untreated and calciumtreated Macrocystis pyrifera in a batch system. Bioresour. Technol. 2012, 116, 195-203. (24) Gonzalez, S.; Lopez-Roldan, R.; Cortina, J. L. Presence of metals in drinking water distribution networks due to pipe material leaching: a review. Toxicol. Environ. Chem. 2013, 95, 870-889. (25) Gikas, P. Single and combined effects of nickel (Ni(II)) and cobalt (Co(II)) ions on activated sludge and on other aerobic microorganisms: A review. J. Hazard Mater. 2008, 159, 187-203. (26) Kamika, I.; Momba, M. N. B. Synergistic effects of vanadium and nickel on heavy metal-tolerant microbial species in wastewater systems. Desalin. Water Treat. 2013, 51, 40-42. (27) Gauthier, P. T.; Norwood, W. P.; Prepas, E. E.; Pyle, G. G. Metal-PAH mixtures in the aquatic environment: A review of co-toxic mechanisms leading to more-than-additive outcomes. Aquat. Toxicol. 2014, 154, 253-269. (28) Horvat, T.; Vidaković-Cifrek, Ž.; Oreščanin, V.; Tkalec, M.; Pevalek-Kozlina, B. Toxicity assessment of heavy metal mixtures by Lemna minor L. Sci. Total Environ. 2007, 384, 229-238. (29) Sciban, M.; Radetic, B.; Kevresan, D.; Klasnja, M. Adsorption of heavy metals from electroplating wastewater by wood sawdust. Bioresour. Technol. 2007, 98, 402-409. (30) Logan, B.E.; Cheng, S.; Watson, V.; Estadt, G. Graphite fiber brush anodes for increased power production in air-cathode microbial fuel cells. Environ. Sci. Technol. 2007, 41, 3341-3346. (31) Logan, B. E. Essential data and techniques for conducting microbial fuel cell and other 24

ACS Paragon Plus Environment

Page 25 of 32

511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564

Environmental Science & Technology

types of bioelectrochemical system experiments. ChemSusChem 2012, 15, 988-994. (32) Li, W.; Sheng, G.; Liu, X.; Yu, H. Recent advances in the separators for microbial fuel cells. Bioresour. Technol. 2011, 102, 244-252. (33) Wang, Q.; Garrity, G. M.; Tiedje, J. M.; Cole, J. R. Naïve Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl. Environ. Microbiol. 2007, 73, 5261-5267. (34) Caporaso, J. G.; Kuczynski, J.; Stombaugh, J.; Bittinger, K.; Bushman, F. D.; Costello, E. K.; Fierer, N.; Pena, A. G.; Goodrich, J. K.; Gordon, J. I.; Huttley, G. A.; Kelley, S. T.; Knights, D.; Koenig, J. E.; Ley, R. E.; Lozupone, C. A.; McDonald, D.; Muegge, B. D.; Pirrung, M.; Reeder, J.; Sevinsky, J. R.; Turnbaugh, P. J.; Walters, W. A.; Widmann, J.; Yatsunenko, T.; Zaneveld, J.; Knight, R. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 2010, 7, 335-336. (35) Lozupone, C.; Knight, R. UniFrac: a new phylogenetic method for comparing microbial communities. Appl. Environ. Microbiol. 2005, 71, 8228-8235. (36) Ter Heijne, A.; Liu, F.; Van Der Weijden, R.; Weijma, J.; Buisman, C. J. N.; Hamelers, H. V. M. Copper recovery combined with electricity production in a microbial fuel cell. Environ. Sci. Technol. 2010, 44, 4376-4381. (37) Huang, L.; Liu, Y.; Yu, L.; Quan, X.; Chen, G. A new clean approach for production of cobalt dihydroxide from aqueous Co(II) using oxygen-reducing biocathode microbial fuel cells. J. Clean. Prod. 2015, 86, 441-446. (38) Wang, Q.; Huang, L.; Yu, H.; Quan, X.; Li, Y.; Fan, G.; Li, L. Assessment of five different cathode materials for Co(II) reduction with simultaneous hydrogen evolution in microbial electrolysis cells. Int. J. Hydrogen Energy 2015, 40, 184-196. (39) Zhuang, L.; Yuan, Y.; Yang, G.; Zhou, S. In situ formation of graphere/biofilm composites for enhanced oxygen reduction in biocathode microbial fuel cells. Electrochem. Commun. 2012, 21, 69-72. (40) Xia, X.; Sun, Y.; Liang, P.; Huang, X. Long-term effect of set potential on biocathodes in microbial fuel cells: electrochemical and phylogenetic characterization. Bioresour. Technol. 2012, 120, 26-33. (41) Huang, L.; Wang, Q.; Quan, X.; Liu, Y.; Chen, G. Bioanodes/biocathodes formed at optimal potentials enhance subsequent pentachlorophenol degradation and power generation from microbial fuel cells. Bioelectrochemistry 2013, 94, 13-22. (42) Varia, J.; Martínez, S. S.; Orta, S. V.; Bull, S.; Roy, S. Bioelectrochemical metal remediation and recovery of Au3+, Co2+ and Fe3+ metal ions. Electrochimica Acta 2013, 95, 125-131. (43) Liang, B.; Cheng, H. Y.; Kong, D. Y.; Gao, S. H.; Sun, F.; Cui, D.; Kong, F. Y.; Zhou, A. J.; Liu, W. Z.; Ren, N. Q.; Wu, W. M.; Wang, A. J.; Lee, D. J. Accelerated reduction of chlorinated nitroaromatic antibiotic chloramphenicol by biocathode. Environ. Sci. Technol. 2013, 47, 5353-5361. (44) Sun, X. F.; Ma, Y.; Liu, X. W.; Wang, S. G.; Gao, B. Y.; Li, X. M. Sorption and detoxification of chromium(VI) by aerobic granules functionalized with polyethylenimine. Water Res. 2010, 44, 2517-2524. (45) Barrera-Díaz, C. E.; Lugo-Lugo, V.; Bilyeu, B. A review of chemical, electrochemical and biological methods for aqueous Cr(VI) reduction. J. Hazard. Mater. 2012, 223-224, 1-12. (46) Kotaś, J.; Stasicka, Z. Chromium occurrence in the environment and methods of its speciation. Environ. Pollut. 2000, 107, 263-283. (47) Batlle-Vilanova, P.; Puig, S.; Gonzalez-Olmos, R.; Vilajeliu-Pons, A.; Bañeras, L.; Balaguer, M. D.; Colprim, J. Assessment of biotic and abiotic graphite cathodes for hydrogen production in microbial electrolysis cells. Inter. J. Hydrogen Energy 2014, 39, 1297-1305. (48) Sharma, M.; Bajracharya, S.; Gildemyn, S.; Patil, S. A.; Alvarez-Gallego, Y.; Pant, D.; Rabaey, K.; Dominguez-Benetton, X. A critical revisit of the key parameters used to describe microbial electrochemical systems. Electrochim. Acta 2014, 140, 191-208. (49) Harnisch, F.; Freguia, S. A basic tutorial on cyclic voltammetry for the investigation of 25

ACS Paragon Plus Environment

Environmental Science & Technology

565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605

Page 26 of 32

electroactive microbial biofilms. Chem. Asian J. 2012, 7, 466-475. (50) Fricke, K.; Harnisch, F.; Schrӧder, U. On the use of cyclic voltammetry for the study of anodic electron transfer in microbial fuel cells. Energy Environ. Sci. 2008, 1, 144-147. (51) Fu, Q.; Kobayashi, H.; Kawaguchi, H.; Wakayama, T.; Maeda, H.; Sato, K. A thermophilic gram-negative nitrate-reducing bacterium, Calditerrivibrio nitroreducens, exhibiting electricity generation capability. Environ. Sci. Technol. 2013, 47, 12583-12590. (52) Pons, L.; Delia, M. L.; Bergel, A. Effect of surface roughness, biofilm coverage and biofilm structure on the electrochemical efficiency of microbial cathodes. Bioresour. Technol. 2011, 102, 2678-2683. (53) Tao, H.; Liang, M.; Li, W.; Zhang, L.; Ni, J.; Wu, W. Removal of copper from aqueous solution by electrodeposition in cathode chamber of microbial fuel cell. J. Hazard. Mater. 2011, 189, 186-192. (54) An, Z.; Zhang, H.; Wen, Q.; Chen, Z.; Du, M. Desalination combined with copper(II) removal in a novel microbial desalination cell. Desalination 2014, 346, 115-121. (55) Liu, L.; Yuan, Y.; Li, F.; Feng, C. In-situ Cr(VI) reduction with electrogenerated hydrogen peroxide driven by iron-reducing bacteria. Bioresour. Technol. 2011, 102, 2468-2473. (56) Wang, G.; Huang, L.; Zhang, Y. Cathodic reduction of hexavalent chromium [Cr(VI)] coupled with electricity generation in microbial fuel cells. Biotechnol. Lett. 2008, 30, 1959-1966. (57) Szpyrkowicz, L.; Zilio-Grandi, F.; Kaul, S. N.; Rigoni-Stern, S. Electrochemical treatment of copper cyanide wastewaters using stainless steel electrodes. Wat. Sci. Tech. 1998, 38, 261-268. (58) Montiel, T.; Solorza, O.; Sánchez, H. Study of cadmium electrochemical deposition in sulfate medium. J. Electrochem. Soc. 2000, 147, 1031-1037. (59) Elsayed, E. M.; Saba, A. E. The electrochemical treatment of toxic hexavalent chromium from industrial effluents using rotating cylinder electrode cell. Int. J. Electrochem. Sci. 2009, 4, 627-639. (60) Grau J. M.; Bisang J. M. Electrochemical removal of cadmium from dilute aqueous solutions using a rotating cylinder electrode of wedge wire screens. J. Appl. Electrochem. 2007, 37, 275-282. (61) Logan, B. E. Exoelectrogenic bacteria that power microbial fuel cells. Nat. Rev. Microbiol. 2009, 7, 375-381. (62) Lesnik, K. L.; Liu, H. Establishing a core microbiome in acetate-fed microbial fuel cells. Appl. Microbiol. Biotechnol. 2014, 98, 4187-4196. (63) Wei, J.; Liang, P.; Huang, X. Recent progress in electrodes for microbial fuel cells. Bioresour. Technol. 2011, 102, 9335-9344. (64) Li, W.; Yu, H.; He, Z. Towards sustainable wastewater treatment by using microbial fuel cells-centered technologies. Energy Environ. Sci. 2014, 7, 911-924. (65) Liu, X.; Li, W.; Yu, H. Cathodic catalysts in bioelectrochemical systems for energy recovery from wastewater. Chem. Soc. Rev. 2014, 43, 7718-7745.

26

ACS Paragon Plus Environment

Page 27 of 32

Environmental Science & Technology

-4

10 bio-5mg/L abiotic-5mg/L bio-0mg/L abiotic-0mg/L

-10

0.65 0.70 0.75 0.80 p

0.02 0.01 0.00 -0.01 -0.02

abiotic-v1 abiotic-v1/2

y = -0.4762x - 0.468 R² = 0.9521

-2.0

y = -0.0863x - 2.7389 R² = 0.9831 y = -0.3627x - 2.3984 R² = 0.9516

0.200.220.240.26

-4.0 0.0

bio-5mg/L abiotic-5mg/L bio-0mg/L abiotic-0mg/L

bio-v1 abiotic-v1

C -1.0

0 0.04 0.02 0.00 -0.02 -0.04 -0.06 -0.22

-2

E

y = -0.1137x - 0.9128 R² = 0.9906

-3.0

-1

y = -0.2853x - 8.9798 R² = 0.9812

-1.0

Current (mA)

Current (mA)

y = -1.1977x - 7.8579 R² = 0.9469

bio-v1 bio-v1/2

1

-0.20

bio-v1/2 abiotic-v1/2

F

y = -0.3995x - 1.125 R² = 0.9787 y = -0.0941x - 1.5051 R² = 0.9915

-2.0

-3.0

y = -0.3721x - 1.5789 R² = 0.9799

y = -0.0876x - 1.9331 R² = 0.9917

-0.18

-3

-4.0 -1.5

606 607 608 609 610 611 612 613

y = -1.3104x - 5.4266 R² = 0.9592

B

0

2

y = -0.3102x - 6.6646 R² = 0.9818

-14 0.0

bio-5mg/L abiotic-5mg/L bio-0mg/L abiotic-0mg/L

-4 3

-8

-12

2

-2

D

-10

0.2 0.1 0.0 -0.1 -0.2

-15 4

Current (mA)

Current (mA)

0

-5

bio-v1/2 abiotic-v1/2

-6

Current (mA)

Current (mA)

5

bio-v1 abiotic-v1

A

-1.0

-0.5

0.0 0.5 1.0 Potential (vs. SHE, V)

1.5

-4

0

4 Scan rate (mV/s)

8

12

Figure 1 (A, B and C) CV tests and (D, E and F) peak current – scan rate on the biocathodes of (A and D) MFC-Cr at an initial Cr(VI) of 5 mg/L, (B and E) MFC-Cu at a Cu(II) of 5 mg/L, and (C and F) MEC-Cd at a Cd(II) of 5 mg/L. Inserts in subfigures (A, B and C): enlarging for comparisons of biocathodes and abiotic cathodes.

27

ACS Paragon Plus Environment

Environmental Science & Technology

0.5

Metal concentrations (mg/L)

5

OCCs-Cr OCCs-Cu OCCs-Cd

abiotic-Cr abiotic-Cu abiotic-Cd

0.4

Voltage output (V)

4 3 2

30

20 0.2 10

0.1 0

0 0

1

2 Time (h)

3

0 0

4

6

0.5 bio-Cu bio-Cd

Metal concentrations (mg/L)

5

OCCs-Cu OCCs-Cd

0.4

Voltage output (V)

3 2

B

1

100

200 Current density (µA)

300

400 40

mixed-elevated-voltage mixed-low-voltage mixed-elevated-power mixed-low-power

abiotic-Cu abiotic-Cd

4

E 30

0.3 20 0.2 10

0.1 0.0

0 0

1

2 Time (h)

bio-Cd

OCCs-Cd

3

0 0

4

5

100

200 Current density (µA)

300

400

-0.600 abiotic-Cd

F

4 -0.605

Cathode potential (V)

Metal concentrations (mg/L)

D

0.3

A

1

40 mixed-elevated-voltage mixed-low-voltage mixed-elevated-power mixed-low-power

Power production (µW)

bio-Cr bio-Cu bio-Cd

Power production (µW)

6

3

-0.610

2

-0.615

1

C

0

-0.620 0

614 615 616 617 618

Page 28 of 32

1

2 Time (h)

3

4

10

12

14 16 Number of cycle

18

Figure 2 Change of Cr(VI), Cu(II) and Cd(II) in (A) MFC-Cr, (B) MFC-Cu, and (C) MEC-Cd, polarization curves in (D) MFC-Cr and (E) MFC-Cu, and (F) cathode potential of MEC-Cd with elevated mixed metals of identical Cr(VI), Cu(II) and Cd(II) at 5 mg/L.

28

ACS Paragon Plus Environment

Page 29 of 32

Environmental Science & Technology

0

5

mixed-low mixed-elevated

mixed-low mixed-elevated -2

Current (mA)

Current (mA)

A 2

-1

y = -0.3835x - 2.6135 R² = 0.9819

-4

y = -0.375x - 3.0114 R² = 0.9935

-6

p

-4 1.5

-8 0 mixed-low mixed-elevated

B

-1

0.5

-3 0

-1.5 2 C

mixed-low mixed-elevated

mixed-low mixed-elevated

Current (mA)

-1

0

y = -0.1246 x - 0.8378 R² = 0.9803 y = -0.1425 x - 0.8570 R² = 0.9827

-2

-1

-3

-2 -1.0

620 621 622 623 624

F

C

1

Current (mA)

y = -0.1138x - 0.8915 R² = 0.9833

-2

-0.5

E

y = -0.1221x - 0.763 R² = 0.9779

Current (mA)

Current (mA)

mixed-low mixed-elevated

619

D

-0.5

0.0 0.5 1.0 Potential (vs. SHE, V)

1.5

0

2

4

6 8 Scan rate (mV/s)

10

12

Figure 3 Comparison of (A, B and C) CV tests and (D, E and F) peak current – scan rate on the biocathodes of (A and D) MFC-Cr, (B and E) MFC-Cu, and (C and F) MEC-Cd under mixed metal influents of either Cr(VI), Cu(II) and Cd(II) at an identical 5 mg/L, or 3 mg/L Cr(VI), 1 mg/L Cu(II) and 1 mg/L Cd(II).

29

ACS Paragon Plus Environment

Environmental Science & Technology

625 626 627 628 629 630

Figure 4 Weighted Fast UniFrac (A) principle coordinates analysis and (B) cluster of the bacterial communities on the basis of phylogenetic lineages that samples contain.

30

ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32

Environmental Science & Technology

120 others Synergistetes

unclassified Bacteroidetes

Actinobacteria Tenericutes

Firmicutes Proteobacteria

Deferribacteres

Relative abundance (%)

100

80

60

40

20

0 120 others Betaproteobacteria

Deltaproteobacteria Alphaproteobacteria

Gammaproteobacteria B

100

Relative abundance (%)

80

60

40

20

0 single

631 632 633 634 635 636 637 638 639 640 641

mixed MFC-Cr

single

mixed MFC-Cu

single

mixed MEC-Cd

Figure 5 Comparison of relative abundance of bacterial reads retrieved from the biocathodes adaptively evolving with single and mixed metals, and classified at (A) the phylum level and (B) the class level distribution of the most dominant phylum of Proteobacteria. Phyla and classes that represent less than 1.0% of the total bacterial community composition were classified as “others”.

31

ACS Paragon Plus Environment

Environmental Science & Technology

642 643 644 645

Figure 6 Comparison of relative abundance of bacterial reads retrieved from the biocathodes adaptively evolving with single and mixed metals at the genus level. Genera that represent less than 1.0% of the total bacterial community composition were classified as “others”.

646

32

ACS Paragon Plus Environment

Page 32 of 32