Accepted Manuscript Impact of Electrode Configurations on Retention Time and Domestic Wastewater Treatment Efficiency Using Microbial Fuel Cells Kyoung-Yeol Kim, Wulin Yang, Bruce E. Logan PII:

S0043-1354(15)00300-0

DOI:

10.1016/j.watres.2015.05.021

Reference:

WR 11296

To appear in:

Water Research

Received Date: 20 March 2015 Revised Date:

5 May 2015

Accepted Date: 7 May 2015

Please cite this article as: Kim, K.-Y., Yang, W., Logan, B.E., Impact of Electrode Configurations on Retention Time and Domestic Wastewater Treatment Efficiency Using Microbial Fuel Cells, Water Research (2015), doi: 10.1016/j.watres.2015.05.021. 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.

RI PT

ACCEPTED MANUSCRIPT

SC

1

AC C

EP

TE D

M AN U

2

1

ACCEPTED MANUSCRIPT

1 2 3

Date: March 17, 2015 Submitted to: Water Research

Impact of Electrode Configurations on Retention Time and Domestic

5

Wastewater Treatment Efficiency Using Microbial Fuel Cells

RI PT

4

6

Kyoung-Yeol Kim, Wulin Yang, and Bruce E. Logan*

8

Department of Civil and Environmental Engineering, The Pennsylvania State University, 231Q

9

Sackett Building, University Park, PA 16802, USA

10

*Corresponding author: e-mail: [email protected]; phone: +1-814-863-7908; fax: +1-814-863-7304

M AN U

SC

7

11 Abstract

13

Efficient treatment of domestic wastewater under continuous flow conditions using microbial

14

fuel cells (MFCs) requires hydraulic retention times (HRTs) that are similar to or less than those

15

of conventional methods such as activated sludge. Two MFCs in series were compared at

16

theoretical HRTs of 8.8, 4.4 and 2.2 h using two different brush-electrode MFC configurations: a

17

full brush evenly spaced between two cathodes (S2C); and trimmed brush anodes near a single

18

cathode (N1C). The MFCs with two cathodes produced more power than the MFCs with a single

19

cathode, with 1.72 mW for the S2C, compared to and 1.12 mW for the N1C at a set HRT = 4.4 h.

20

The single cathode MFCs with less cathode area removed slightly more COD (54.2 ± 2.3%, N1C)

21

than the two-cathode MFCs (48.3 ± 1.0%, S2C). However, the higher COD removal was due to

22

the longer HRTs measured for the MFCs with the N1C configuration (10.7, 5.3 and 3.1 h) than

23

with the S2C configuration (7.2, 3.7 and 2.2 h), despite the same theoretical HRT. The longer

AC C

EP

TE D

12

1

ACCEPTED MANUSCRIPT

HRTs of the N1C MFCs also resulted in slightly higher coulombic efficiencies (≤ 37%) than

25

those of the S2C MFCs (≤ 29%). While the S2C MFC configuration would be more

26

advantageous based on electrical power production, the N1C MFC might be more useful on the

27

basis of capital costs relative to COD removal efficiency due to the use of less cathode surface

28

area per volume of reactor.

RI PT

24

29

Keywords: Microbial fuel cell; Continuous flow mode; Brush anode electrode; Flow pattern;

31

Hydraulic retention time

SC

30

33

M AN U

32 1. Introduction

Microbial fuel cells (MFCs) have been studied as an energy-neutral process which can

35

produce electrical energy from organic compounds while accomplishing wastewater treatment

36

(Li et al. 2014, Rabaey and Verstraete 2005, Zhang et al. 2013). However, recent studies have

37

shown that both high power densities and low effluent concentrations of chemical oxygen

38

demand (COD) cannot be simultaneously achieved in MFCs under continuous flow conditions

39

(Akman et al. 2013, Nam et al. 2010, Zhuang et al. 2012). When the COD concentration is

40

reduced to a range of ~100 mg/L, there is minimal current generation relative to higher COD

41

concentrations (Ren et al. 2014a, Zhang et al. 2015). Thus, a secondary treatment process such as

42

an anaerobic fluidized bed membrane bioreactor (AFMBR) is needed to reduce the COD to

43

levels suitable for discharge (Ren et al. 2014b). The goal of MFC treatment is therefore not COD

44

removal to a level suitable for discharge, but COD reduction to levels that maintain good power

45

production. The low COD levels needed for discharge can be achieved using AFMBRs or other

AC C

EP

TE D

34

2

ACCEPTED MANUSCRIPT

46

types of processes such as biofilters or solids contact systems (Chernicharo 2006, Kim et al.

47

2011). The materials used in larger-scale MFCs for wastewater treatment need to be inexpensive,

49

effective for treatment of actual wastewaters that contain particulate organic matter, and produce

50

stable power over time. Flat anodes made of carbon cloth or mesh can allow close electrode

51

spacing and high power densities with acetate solutions (Liu et al. 2005), but they have been

52

found to result in unstable power production when used in MFCs treating domestic wastewater

53

(Hays et al. 2011) and small electrode spacing could result in reactor clogging. Graphite fiber

54

brush anodes can be used in MFCs to achieve good rates of COD removal, they have high

55

surface areas and porosities for growth of exoelectrogenic microorganisms, and they produce

56

stable power over time using actual domestic wastewaters (Ahn and Logan 2013, Logan et al.

57

2007). MFCs have been designed and tested with multiple brush anodes configured primarily in

58

two ways: with space between the anodes and the cathodes in the presence and absence of a

59

separator; or with brushes placed directly on top of the separator with little space between the

60

electrodes (Ahn et al. 2014, Zhang et al. 2014a). The two configurations have different

61

advantages relative to power production, but separators are needed for both configurations to

62

reduce cathode biofouling (Ren et al. 2014b). Cathode materials and performance are also

63

important factors in MFC designs. Activated carbon cathodes have been shown to produce power

64

densities similar to, or even larger than, cathodes made with a Pt catalyst (Cheng and Wu 2013,

65

Zhang et al. 2009, Zhang et al. 2014b). The performance of activated carbon is maintained much

66

better over time compared to Pt, and power production can be nearly restored to levels

67

comparable to new cathodes using an acid treatment. Cathode performance typically limits

AC C

EP

TE D

M AN U

SC

RI PT

48

3

ACCEPTED MANUSCRIPT

68

power generation in MFCs at higher COD concentrations. Doubling the cathode surface area, for

69

example, increased maximum power densities from domestic wastewater by 62% in fed-batch

70

tests (Cheng and Logan 2011). In order for large-scale wastewater treatment using MFCs to be feasible using these low-cost

72

brush anodes and activated carbon cathodes, the hydraulic retention times (HRTs) in the reactors

73

has to be reduced to be more similar to that used for conventional activated sludge processes.

74

However, there have been relatively few studies focused on achieving wastewater treatment with

75

HRTs in the range of ~10 h or less using domestic wastewater (Min and Logan 2004, Puig et al.

76

2011). Short HRTs can be accomplished by improving current densities, and therefore achieving

77

higher rates of COD removal. Two different graphite fiber brush anode reactor configurations

78

were examined here for wastewater treatment at HRTs of less than 10 h: anodes (trimmed flat on

79

one side) placed directly on top of a separator covering a single cathode so that they are very near,

80

in order to reduce solution ohmic resistance (N1C), resulting in no flow between the electrodes;

81

and anodes (full brush) positioned in the middle of the chamber with space for flow between the

82

brush and the two cathodes each covered by a separator (S2C). The two types MFCs were

83

operated under continuous flow conditions at three different theoretical HRTs (8.8, 4.4, and 2.2 h)

84

using two MFCs in series to minimize the change in COD in a single reactor. Power generation

85

and COD removal were monitored over time, with maximum power densities evaluated using

86

polarization data. The different anode configurations can affect flow patterns in these reactors,

87

and thus the actual HRTs were measured using a salt tracer.

89

SC

M AN U

TE D

EP

AC C

88

RI PT

71

2. Materials and methods 4

ACCEPTED MANUSCRIPT

90

2.1. MFC construction

91

Anodes were graphite fiber brushes (Mill-Rose, Mentor, OH) with a titanium core (25 mm

92

diameter by 35 mm length) heat treated at 450 °C for 30 min before use. Cathodes (40 cm2,

93

projected

94

poly(vinylidenefluoride, PVDF) binder and a mixture of activated carbon (AC, 8.8 mg/cm2) and

95

carbon black (CB, Vulcan XC-72, Cabot Corporation, USA) as previously described (Yang et al.

96

2014).

were

made

by

a

phase

inversion

technique

using

RI PT

area)

a

SC

surface

Four single-chamber, air-cathode MFCs were constructed using three graphite fiber brush

98

anodes per reactor. Two of these reactors had the anodes placed directly next to a single cathode

99

(N1C) similar to previous designs (Ahn et al. 2014) (Fig. 1). The other two MFCs had the anodes

100

placed in the middle of the anode chamber with a space between the edge of the anodes and two

101

cathodes on either side of the brushes (S2C). The spaced electrode reactor design is different

102

from previous spaced electrode designs as two cathodes each containing separators were used

103

per anode array, compared to MFCs tested with a single cathode without a separator (Zhang et al.

104

2014a). Brushes used in the N1C reactor (100 mL empty bed volume) were trimmed along their

105

length to form a half cylinder, with the flat side placed on the separator that was positioned

106

against the air-cathode (0.5 cm distance from the cathode). Full brushes were used in the S2C

107

reactors (140 mL empty bed volume) with an edge-cathode distance of 0.8 cm (Fig. 1). These

108

different configurations resulted in a higher cathode surface area (57 m2/m3) for the S2C reactor

109

than the N1C reactors (40 m2/m3). Each reactor had an entrance and exit chamber (no electrodes)

110

of 10 mL each to help distribute flow evenly into the reactors (Fig. S1). Two layers of a textile

111

separator (46% cellulose, 54% polyester; 0.3 mm thick; Amplitude Prozorb, Contec Inc.) were

AC C

EP

TE D

M AN U

97

5

ACCEPTED MANUSCRIPT

used as separators on each cathode surface (liquid side) to reduce cathode biofouling and oxygen

113

intrusion into the anode chamber, and in the case of the N1C reactors, to prevent electrode short-

114

circuiting. A reference electrode (Ag/AgCl; + 200 mV vs. standard hydrogen electrode, SHE)

115

was installed in the upper and middle part of the MFCs to record electrode potentials (all

116

potentials reported versus SHE).

RI PT

112

117 2.2. Operational methods

SC

118

The four MFC reactors (2 N1Cs, 2 S2Cs) were arranged with two flow paths, with the two

120

similar types of MFCs hydraulically connected in series (Fig. 1). The first MFC was the labeled

121

the upstream (U) reactor, and the second one the downstream (D) reactor. The MFCs were

122

inoculated and fed domestic wastewater from the primary clarifier of the Pennsylvania State

123

University Wastewater Treatment Plant. Wastewater was collected once a week and stored in a

124

refrigerator (4 °C). When used, the wastewater was placed next to the reactor in a container filled

125

with ice, which preserved the wastewater composition and allowed the solution to warm as it was

126

pumped into the upstream MFCs using peristaltic pumps (model no.7523-90, Masterflex, Vernon

127

Hills, IL).

EP

TE D

M AN U

119

The theoretical HRT was calculated from the set flow rate (Q, mL/h) and reactor volume (V,

129

mL) as HRT = V/Q. Flow rates were chosen to obtain the same theoretical HRTs for the two

130

reactors of: 8.8 h (31.8 mL/h, S2C; 22.8 mL/h, N1C); 4.4 h (63.6 mL/h, S2C; 45.6 mL/h, N1C);

131

2.2 h (127.2 mL/h, S2C; 91.2 mL/h, N1C). A salt tracer test was conducted to measure the actual

132

residence time of solutes in the two pairs of MFCs during continuous mode operation. A NaCl

133

solution (1 M) was injected into the flow line of the upstream flow line of the MFC for 36 min

AC C

128

6

ACCEPTED MANUSCRIPT

(8.8 h HRT), 18 min (4.4 h HRT), or 9 min (2.2 h HRT), with the effluent conductivity monitored

135

using a meter (SB90M5, VWR International, USA). Measured HRTs were defined as the time to

136

the peak of the tracer input as previously described (Min et al. 2004, Zhang et al. 2006). The

137

HRT measured using this approach includes the entrance and exit volumes (not working area) of

138

the reactors. Thus, the measured HRTs were adjusted by multiplying a ratio of working volume

139

to total reactor volume. A dispersion number (d) was calculated from the normalized variance of

140

the salt tracer curve. A Peclet number (Pe) was calculated as Pe = 1/d, where Pe = uL/E is a ratio

141

of mass transfer by advection to that of dispersion, where u is the average cross-sectional

142

velocity, L is the distance of flow, and E is the dispersion coefficient.

143 144

2.3. Analysis and calculations

M AN U

SC

RI PT

134

Cell voltages (external resistor of 200 Ω) of each MFC were monitored at every 10 min using

146

a multimeter (Model 2700, Keithley Instruments, USA) and used to calculate the current based

147

on the voltage drop across the resistor. Polarization and power curves were obtained by varying

148

the external resist from open circuit to 50 Ω. Reactors were first operated under open circuit

149

conditions for at least 4 h prior to each polarization test (Kim et al. 2014).

EP

150

TE D

145

Coulombic efficiency (CE) was calculated as CE (%) = Ct/Cth

100, where Ct was the total

coulombs calculated by integrating the current over time (Ct = ∑I ∆t, where ∆t is the time

152

interval of one HRT), and Cth was the theoretical amount of coulombs available based on the

153

COD removed in the MFC over the same amount of time, calculated as Cth = [Fb (CODin –

154

CODout) Q ∆t]/M, where F is Faraday’s constant, b = 4 is the number of electrons exchanged per

155

mole of oxygen, CODin and CODout are the influent and effluent COD, ∆t is the time interval

AC C

151

7

ACCEPTED MANUSCRIPT

156

(HRT), and M = 32 is the molecular weight of oxygen (Logan et al. 2006). Total COD was

157

measured following standard methods (method 5220, HACH company, Loveland, CO).

159

3. Results and discussion

160

3.1. Power generation as a function of theoretical HRT

RI PT

158

When the MFCs were started up at the longest HRT of 8.8 h (200 Ω external resistance),

162

there was inconsistent power produced for the similar reactor configurations (Fig. 2). For

163

example, the S2C upstream MFC produced the most power, but the downstream S2C produced

164

the least power. When operation was switched to shorter HRTs of 4.4 and 2.2 h, there was more

165

consistent power production, with the S2C MFCs consistently producing more power (1.36 mW

166

at 4.4 h, and 1.30 mW at 2.2 h) than the N1C MFCs (1.01 mW at 4.4 h, and 1.22 mW at 2.2 h).

167

The upstream MFCs generally produced greater power than the downstream MFCs, likely due to

168

the higher COD concentrations in the upstream MFCs. After these tests at the two shortest HRTs,

169

the reactors were switched back to the longer 8.8 h HRT, and a fresh wastewater sample was

170

used. This time at this longer HRT, the MFCs produced more consistent power densities based on

171

reactor configuration, with power by the two-cathode MFCs (1.14 mW, S2C) larger than that

172

produced by the single-cathode MFCs (0.85 mW, N1C).

EP

TE D

M AN U

SC

161

Power and polarization curves (Fig. 3) also demonstrated that higher power could be

174

produced by the S2C MFCs than the single cathode N1C MFCs at all HRTs. For example, at the

175

shortest HRT of 2.2 h, the maximum power of the two S2C MFCs was 1.67 mW, compared to

176

1.20 mW for the N1C MFCs. In most cases, similar power densities were produced by both the

177

upstream and downstream MFCs. At the shortest theoretical HRT, however, the maximum power

AC C

173

8

ACCEPTED MANUSCRIPT

produced by the downstream MFCs based on these polarization data was equal to or slightly

179

higher than that of the upstream MFCs. The largest difference was obtained at an HRT of 2.2 h

180

where the downstream S2C produced 0.89 ± 0.03 mW compared to 0.78 ± 0.04 mW for the

181

upstream reactor (Fig. 3c). For the N1C reactors, the downstream reactor produced 0.65 ± 0.03

182

mW, compared to 0.55 ± 0.02 mW for the upstream MFC. Together, these averaged 209 mW/m2

183

for the S2C and 300 mW/m2 for the N1C. There were smaller differences between the upstream

184

and downstream MFCs at the longer HRTs, with average power produced of 1.72 mW (S2C) and

185

1.12 mW (N1C) at a 4.4 h HRT, and 1.61 mW (S2C) and 1.06 mW (N1C) at an 8.8 h HRT. There

186

are several possible reasons for the slightly higher maximum power densities in the downstream

187

MFCs. Polarization data represent the capability of the MFCs for power production based on

188

short-term tests, but with longer acclimation the power densities may have been more similar.

189

For example, in Figure 2 it was shown that under steady flow conditions, and at the same

190

external resistance, the upstream reactor produced higher voltages than the downstream reactors.

191

Power overshoot, where the power density curve doubles back to lower current densities

192

oftentimes before a higher power density that is possible with further reactor acclimation, could

193

also be a factor as overshoot was observed in several tests (Fig. 3). Differences in the

194

composition and quality of the wastewater in the upstream and downstream reactors could also

195

be important. Domestic wastewater contains both soluble and particulate COD, and it thus there

196

may be more fermentation and particle hydrolysis occurring in the upstream reactor, with more

197

labile and biodegradable organics in the downstream reactor.

AC C

EP

TE D

M AN U

SC

RI PT

178

198

The higher power production of the S2C MFCs with two cathodes is consistent with previous

199

results showing that increasing cathode area can result in higher power densities (Cheng and 9

ACCEPTED MANUSCRIPT

Logan 2011). The S2C reactors had 43% more surface area (57 m2/m3) than the N1C reactors (40

201

m2/m3), and produced 39-53% more power on average at the three different flow rates based on

202

polarization data. In contrast, only a difference of 16% was observed previously when the three

203

anodes were placed closer to a single cathode (Ahn et al. 2014). Thus, the closer placement of

204

the anode to the cathode in the N1C reactors, which would reduce ohmic resistance, was less

205

important for higher power production than the use of more cathode surface area.

RI PT

200

207

SC

206 3.2. COD removals as a function of HRTs

COD removals with the two-cathode S2C MFCs decreased in proportion to HRTs, with

209

removals of 64.8 ± 1.7% at an HRT of 8.8 h, decreasing to 48.3 ± 1.0% at 4.4 h, and 32.8 ± 1.9%

210

at 2.2 h (Fig. 4). It was expected that the S2C MFC would have greater COD removal than the

211

N1C MFC due to higher current densities (Zhang et al. 2015) and greater cathode surface area

212

per volume of reactor (which would allow more leakage of oxygen into the reactor) at the same

213

set HRT. However, the COD removals with the single-cathode N1C MFC were higher than those

214

of the S2C MFCs, with removals of 69.0 ± 0.4% at 8.8 h, 54.2 ± 2.3% at 4.4 h, and 40.7 ± 2.7%

215

at 2.2 h. This is 6.4 to 24.4% greater COD removal for the single cathode MFCs than the two-

216

cathode MFCs at each of these set HRTs.

218

TE D

EP

AC C

217

M AN U

208

3.3. Effect of reactor configuration on HRTs in continuous flow mode MFCs

219

The unexpected results showing better COD removals for single cathode MFCs compared to

220

the two-cathode suggested that configuration of the MFCs may have produced HRTs different

221

than the theoretical ones (calculated from reactor empty bed volume and flow rate). Therefore, 10

ACCEPTED MANUSCRIPT

salt tracer tests were conducted to measure the actual residence time of solutes through the two

223

types (S2C and N1C) of reactors. The results showed that the actual HRTs of S2C reactors were

224

less than or equal to the theoretical HRTs, with actual:theoretical HRTs of 7.4:8.8 h, 3.8:4.4 h,

225

and 2.2:2.2 h (Table 1). A slight reduction in HRT due to short circuiting in these MFCs was

226

likely a result of the reactor volume occupied by the brushes as there may not have been

227

advective flow through the brushes. However, the HRTs of the two N1C reactors were longer

228

than those measured for the S2C reactors by 39% (7.4:10.3 h for S2C:N1C), 37% (3.8:5.2 h), and

229

32% (2.2:2.9 h). Thus, the greater removal by the N1C reactors was likely due to the much

230

longer HRTs produced using this configuration compared to that obtained with the S2C

231

configuration.

M AN U

SC

RI PT

222

A comparison of the COD removals on the basis of the measured HRTs shows that there was

233

generally good agreement with the observed COD removals and measured HRTs despite

234

differences in the two reactor configurations (Fig. 5). The COD removals for both types of

235

reactors show good agreement with measured HRTs based on a non-linear regression. However,

236

there was little agreement between the two configurations in terms of COD removal on the basis

237

of the theoretical HRTs, with slightly better COD removals indicated to occur for the N1C MFC

238

than the S2C MFC. These results suggest that the COD removal was more a function of the

239

HRTs, and thus how the HRT was altered by the use of the two different brush configurations,

240

rather other factors such as cathode surface area. The main impact of the reactor geometry was

241

on power production, with more power generated when using two cathodes than one cathode

242

with closely-spaced electrodes.

243

AC C

EP

TE D

232

The low dispersion numbers and high Peclet numbers indicate that flow through the reactors 11

ACCEPTED MANUSCRIPT

was consistent with plug flow reactors with dispersion. For both reactor configurations there was

245

little dispersion, with the dispersion increasing slightly with flow rate (Table 1). Given the high

246

Peclet numbers at the longest theoretical HRT, the reactors could be modeled as a series of

247

completely mixed, constant flow reactors (CSTRs). For example, the S2C MFCs (Pe = 36) could

248

be modeled as 19 CSTRs in series, with a rate constant of –0.0066 h–1 (Logan 1999).

RI PT

244

249 3.4. CEs as a function of the HRTs

SC

250

The highest CE obtained was 36 ± 2% at theoretical HRT of 8.8 h for the two N1C MFCs in

252

series (Fig. 6). The CEs ranged from 18 ± 2 to 36 ± 2% for the N1C MFCs, compared to 18 ± 5

253

to 29 ± 3% for the S2C MFCs. While the CEs were generally higher for N1C MFCs for each

254

pairwise comparison of the two reactor configurations on the basis of theoretical HRT, as noted

255

above the measured HRTs were longer for the N1C MFCs than the S2C MFCs. The downstream

256

MFCs consistently had higher CEs than the upstream MFCs (Fig. 6) although there was a much

257

greater percentage of COD removal in the upstream than downstream MFCs (Fig. 4). This

258

indicated that more of the COD removed in the upstream MFC was not used for electricity

259

generation, and was lost to alternate electron acceptors, such as in dissolved oxygen. The CEs are

260

known to be much lower in MFCs with domestic wastewater than with defined media such as

261

acetate in a phosphate buffer solution.

263

TE D

EP

AC C

262

M AN U

251

4. Conclusions

264

Brush anode and activated carbon cathodes can be used as effective and inexpensive

265

electrode materials in MFCs. The use of closely spaced electrodes and high cathode specific 12

ACCEPTED MANUSCRIPT

surface areas have been shown to improve power generation in MFCs, but little attention has

267

been paid to treatment performance in these systems. Higher power production was achieved

268

using reactors with greater cathode specific surface area than reactors with a configuration based

269

on close electrode spacing. However, there was little difference in COD removals in these two

270

reactor designs on the basis of measured HRTs. The MFC with the closely spaced electrodes had

271

longer measured HRTs, and thus slightly greater COD removals than MFCs with spaced

272

electrodes despite the same set HRT. This different results in terms of power and COD removal

273

suggest that there is a tradeoff in MFC design. The use of more cathode surface area per volume

274

of reactor will increase power production, but that design would have a higher capital cost due to

275

the use of more cathodes. It might be that the use of less cathode area per reactor would be more

276

cost effective than the recovery of more electrical power. Thus, the benefits of power production

277

will need to be considered relative capital costs and advantages of the different reactor designs.

M AN U

SC

RI PT

266

TE D

278 Acknowledgments

280

This research was supported by the grant from the Strategic Environmental Research and

281

Development Program (SERDP).

282

EP

279

References

284

Ahn, Y. and Logan, B. (2013) Domestic wastewater treatment using multi-electrode continuous

285

flow MFCs with a separator electrode assembly design. Appl. Microbiol. Biotechnol. 97, 409-

286

416.

287

Ahn, Y., Hatzell, M.C., Zhang, F. and Logan, B.E. (2014) Different electrode configurations to

AC C

283

13

ACCEPTED MANUSCRIPT

optimize performance of multi-electrode microbial fuel cells for generating power or treating

289

domestic wastewater. J. Power Sources 249, 440-445.

290

Akman, D., Cirik, K., Ozdemir, S., Ozkaya, B. and Cinar, O. (2013) Bioelectricity generation in

291

continuously-fed microbial fuel cell: Effects of anode electrode material and hydraulic retention

292

time. Bioresour. Technol. 149, 459-464.

293

Cheng, S. and Logan, B.E. (2011) Increasing power generation for scaling up single-chamber air

294

cathode microbial fuel cells. Bioresour. Technol. 102(6), 4468-4473.

295

Cheng, S. and Wu, J. (2013) Air-cathode preparation with activated carbon as catalyst, PTFE as

296

binder and nickel foam as current collector for microbial fuel cells. Bioelectrochemistry 92, 22-

297

26.

298

Chernicharo, C.D. (2006) Post-treatment options for the anaerobic treatment of domestic

299

wastewater. Rev. Environ. Sci. Bio/Technol. 5(1), 73-92.

300

Hays, S., Zhang, F. and Logan, B.E. (2011) Performance of two different types of anodes in

301

membrane electrode assembly microbial fuel cells for power generation from domestic

302

wastewater. J. Power Sources 196(20), 8293-8300.

303

Kim, J., Kim, K., Ye, H., Lee, E., Shin, C., McCarty, P.L. and Bae, J. (2011) Anaerobic fluidized

304

bed membrane bioreactor for wastewater treatment. Environ. Sci. Technol. 45(2), 576-581.

305

Kim, K.-Y., Yang, E., Lee, M.-Y., Chae, K.-J., Kim, C.-M. and Kim, I.S. (2014) Polydopamine

306

coating effects on ultrafiltration membrane to enhance power density and mitigate biofouling of

307

ultrafiltration microbial fuel cells (UF-MFCs). Water Res. 54, 62-68.

308

Li, W.-W., Yu, H.-Q. and He, Z. (2014) Towards sustainable wastewater treatment by using

309

microbial fuel cells-centered technologies. Energy Environ. Sci. 7(3), 911-924.

AC C

EP

TE D

M AN U

SC

RI PT

288

14

ACCEPTED MANUSCRIPT

Liu, H., Cheng, S. and Logan, B.E. (2005) Power generation in fed-batch microbial fuel cells as

311

a function of ionic strength, temperature, and reactor configuration. Environ. Sci. Technol.

312

39(14), 5488-5493.

313

Logan, B.E. (1999) Environmental Transport Processes, Wiley, New York.

314

Logan, B.E., Hamelers, B., Rozendal, R., Schröder, U., Keller, J., Freguia, S., Aelterman, P.,

315

Verstraete, W. and Rabaey, K. (2006) Microbial fuel cells: methodology and technology. Environ.

316

Sci. Technol. 40(17), 5181-5192.

317

Logan, B.E., Cheng, S., Watson, V. and Estadt, G. (2007) Graphite fiber brush anodes for

318

increased power production in air-cathode microbial fuel cells. Environ. Sci. Technol. 41(9),

319

3341-3346.

320

Min, B., Evans, P.J., Chu, A. and Logan, B.E. (2004) Perchlorate removal in sand and plastic

321

media bioreactors. Water Res. 38(1), 47-60.

322

Min, B. and Logan, B.E. (2004) Continuous electricity generation from domestic wastewater and

323

organic substrates in a flat plate microbial fuel cell. Environ. Sci. Technol. 38(21), 5809-5814.

324

Nam, J.-Y., Kim, H.-W., Lim, K.-H. and Shin, H.-S. (2010) Effects of organic loading rates on

325

the continuous electricity generation from fermented wastewater using a single-chamber

326

microbial fuel cell. Bioresour. Technol. 101(1), S33-S37.

327

Puig, S., Serra, M., Coma, M., Balaguer, M. and Colprim, J. (2011) Simultaneous domestic

328

wastewater treatment and renewable energy production using microbial fuel cells (MFCs). Water

329

Sci. Technol. 64(4), 904-909.

330

Rabaey, K. and Verstraete, W. (2005) Microbial fuel cells: novel biotechnology for energy

331

generation. Trends Biotechnol. 23(6), 291-298.

AC C

EP

TE D

M AN U

SC

RI PT

310

15

ACCEPTED MANUSCRIPT

Ren, L., Ahn, Y., Hou, H., Zhang, F. and Logan, B.E. (2014a) Electrochemical study of multi-

333

electrode microbial fuel cells under fed-batch and continuous flow conditions. J. Power Sources

334

257, 454-460.

335

Ren, L., Ahn, Y. and Logan, B.E. (2014b) A two-stage microbial fuel cell and anaerobic fluidized

336

bed membrane bioreactor (MFC-AFMBR) system for effective domestic wastewater treatment.

337

Environ. Sci. Technol. 48(7), 4199-4206.

338

Yang, W., He, W., Zhang, F., Hickner, M.A. and Logan, B.E. (2014) Single-step fabrication using

339

a phase inversion method of poly (vinylidene fluoride)(PVDF) activated carbon air cathodes for

340

microbial fuel cells. Environ. Sci. Technol. Lett. 1(10), 416-420.

341

Zhang, F., Cheng, S., Pant, D., Bogaert, G.V. and Logan, B.E. (2009) Power generation using an

342

activated carbon and metal mesh cathode in a microbial fuel cell. Electrochem. Commun. 11(11),

343

2177-2179.

344

Zhang, F., Ge, Z., Grimaud, J., Hurst, J. and He, Z. (2013) Long-term performance of liter-scale

345

microbial fuel cells treating primary effluent installed in a municipal wastewater treatment

346

facility. Environ. Sci. Technol. 47(9), 4941-4948.

347

Zhang, F., Ahn, Y. and Logan, B.E. (2014a) Treating refinery wastewaters in microbial fuel cells

348

using separator electrode assembly or spaced electrode configurations. Bioresour. Technol. 152,

349

46-52.

350

Zhang, H., Bruns, M.A. and Logan, B.E. (2006) Biological hydrogen production by Clostridium

351

acetobutylicum in an unsaturated flow reactor. Water Res. 40(4), 728-734.

352

Zhang, X., Xia, X., Ivanov, I., Huang, X. and Logan, B.E. (2014b) Enhanced activated carbon

353

cathode performance for microbial fuel cell by blending carbon black. Environ. Sci. Technol.

AC C

EP

TE D

M AN U

SC

RI PT

332

16

ACCEPTED MANUSCRIPT

48(3), 2075-2081.

355

Zhang, X., He, W., Ren, L., Stager, J., Evans, P.J. and Logan, B.E. (2015) COD removal

356

characteristics in air-cathode microbial fuel cells. Bioresour. Technol. 176, 23-31.

357

Zhuang, L., Zheng, Y., Zhou, S., Yuan, Y., Yuan, H. and Chen, Y. (2012) Scalable microbial fuel

358

cell (MFC) stack for continuous real wastewater treatment. Bioresour. Technol. 106, 82-88.

RI PT

354

AC C

EP

TE D

M AN U

SC

359

17

ACCEPTED MANUSCRIPT

1

Figure captions

2 Figure 1. (A) Full brush anodes spaced evenly between two cathodes (S2C) and trimmed brush

4

anodes near a single cathode (N1C), not drawn to scale. (B) Experimental set-up used in this

5

study. Pairs of S2Cs and N1Cs were arranged in a series, and domestic wastewater was

6

continuously provided through the inlet channel to each of the paired reactors.

7

Figure 2. Power generation (external resistor of 200 Ω for each MFC) produced by S2Cs and

8

N1Cs (U = upstream, D = downstream) during continuous operation under different theoretical

9

HRTs (8.8, 4.4 and 2.2 h).

M AN U

SC

RI PT

3

Figure 3. Polarization and power curves of 4 MFCs with different theoretical HRTs. (a)(d): 8.8 h;

11

(b)(e): 4.4 h; (c)(f): 2.2 h.

12

Figure 4. COD concentrations in the influent and effluent of the S2Cs and N1Cs (U = upstream,

13

D = downstream) and percentage of COD removal at theoretical HRTs of 8.8, 4.4 and 2.2 h.

14

Figure 5. COD removals (%) by S2C and N1C MFCs based on the theoretical and measured

15

HRTs. The regression line (black dashed) was drawn to fit the COD removals based on the

16

measured HRTs of both MFC configurations.

17

Figure 6. CEs of S2Cs and N1Cs (U = upstream, D = downstream) based on theoretical HRTs

18

(8.8, 4.4 and 2.2 h).

AC C

EP

TE D

10

1

ACCEPTED MANUSCRIPT

Table 1. Comparison of theoretical HRTs and measured HRTs with different anode

2

configurations (S2C and N1C). Dispersion and Peclet numbers at each HRT were obtained from

3

salt tracer tests. HRT (h)

Q (L/m2·h)

Configuration

N1C

0.028

36

8.8 (9.9)

7.4 (8.4)

18.2

4.4 (4.9)

3.8 (4.3)

36.3

0.035

28

2.2 (2.5)

2.2 (2.5)

72.7

0.077

13

8.8 (10.5)

10.3 (12.4)

18.2

0.020

49

4.4 (5.3)

5.2 (6.2)

36.5

0.015

68

2.2 (2.6)

2.9 (3.5)

73.0

0.043

23

SC

Measured*

TE D

*HRTs shown are corrected to include only the working reactor volume. Values in parentheses include the whole reactor volume (includes the entrance and exit volumes).

EP

7

Pe

Theoretical*

AC C

4 5 6

Dispersion number (d)

M AN U

S2C

RI PT

1

1

SC

RI PT

ACCEPTED MANUSCRIPT

M AN U

377

Figure 1. (A) Full brush anodes spaced evenly between two cathodes (S2C) and trimmed brush

379

anodes near a single cathode (N1C), not drawn to scale. (B) Experimental set-up used in this

380

study. Pairs of S2Cs and N1Cs were arranged in a series, and domestic wastewater was

381

continuously provided through the inlet channel to each of the paired reactors.

EP AC C

382

TE D

378

19

SC

RI PT

ACCEPTED MANUSCRIPT

M AN U

383 384

Figure 2. Power generation (external resistor of 200 Ω for each MFC) produced by S2Cs and

385

N1Cs (U = upstream, D = downstream) during continuous operation under different theoretical

386

HRTs (8.8, 4.4 and 2.2 h).

EP AC C

388

TE D

387

20

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

389

Figure 3. Polarization and power curves of 4 MFCs with different theoretical HRTs. (a)(d): 8.8 h;

391

(b)(e): 4.4 h; (c)(f): 2.2 h.

EP AC C

392

TE D

390

21

SC

RI PT

ACCEPTED MANUSCRIPT

M AN U

393 394

Figure 4. COD concentrations in the influent and effluent of the S2Cs and N1Cs (U = upstream,

395

D = downstream) and percentage of COD removal at theoretical HRTs of 8.8, 4.4 and 2.2 h.

AC C

EP

TE D

396

22

SC

RI PT

ACCEPTED MANUSCRIPT

M AN U

397 398

Figure 5. COD removals (%) by S2C and N1C MFCs based on the theoretical and measured

399

HRTs. The regression line (black dashed) was drawn to fit the COD removals based on the

400

measured HRTs of both MFC configurations.

AC C

EP

TE D

401

23

SC

RI PT

ACCEPTED MANUSCRIPT

M AN U

402 403

Figure 6. CEs of S2Cs and N1Cs (U = upstream, D = downstream) based on theoretical HRTs

404

(8.8, 4.4 and 2.2 h).

AC C

EP

TE D

405

24

ACCEPTED MANUSCRIPT



Compared performance of MFCs with different electrode configurations.

2



MFC with two cathodes and space between the electrodes produced highest power.

3



Measured HRTs were altered by different brush anode configurations.

4



COD removal was primarily a function of the measured HRTs and not configuration.

AC C

EP

TE D

M AN U

SC

RI PT

1

1

ACCEPTED MANUSCRIPT

Supporting Information Impact of Brush Electrode Configuration on Retention Time and Continuous Domestic Wastewater Treatment Efficiency Using Microbial Fuel Cells

RI PT

Kyoung-Yeol Kim, Wulin Yang, and Bruce E. Logan*

Department of Civil and Environmental Engineering, The Pennsylvania State University, 231Q Sackett Building, University Park, PA 16802, USA

EP

TE D

M AN U

SC

*Corresponding author: e-mail: [email protected]; phone: +1-814-863-7908; fax: +1-814-863-7304

Figure S1. Diagrams of two MFC reactors demonstrating the total reactor volume; S2C (A),

AC C

N1C (B). Both reactors have an entrance and exit region (10 mL each) to evenly distribute the flow-in and -out.

S1

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure S2. Salt tracer curve at a 8.8 h theoretical HRT. Arrows indicate the measured HRTs (8.4

AC C

EP

TE D

h for S2C, 12.4 h for N1C, based on the total reactor volume).

Figure S3. Salt tracer curve at a 4.4 h theoretical HRT. Arrows indicate the measured HRTs (4.3 h for S2C, 6.2 h for N1C, based on the total reactor volume).

S2

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure S4. Salt tracer curve at a 2.2 h theoretical HRT. Arrows indicate the measured HRTs (2.5

AC C

EP

TE D

h for S2C, 3.5 h for N1C, based on the total reactor volume).

S3

Impact of electrode configurations on retention time and domestic wastewater treatment efficiency using microbial fuel cells.

Efficient treatment of domestic wastewater under continuous flow conditions using microbial fuel cells (MFCs) requires hydraulic retention times (HRTs...
2MB Sizes 1 Downloads 11 Views