Bacterial

Sep 14, 2018 - Reactive, Self-Cleaning Ultrafiltration Membrane Functionalized with Iron Oxychloride Nanocatalysts. Environmental Science & Technology...
0 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Sustainability Engineering and Green Chemistry

Photothermally-active Reduced Graphene Oxide/Bacterial Nanocellulose Composites as Biofouling-resistant Ultrafiltration Membranes Qisheng Jiang, Deoukchen Ghim, Sisi Cao, Sirimuvva Tadepalli, Keng-Ku Liu, Hyuna Kwon, Jingyi Luan, Yujia Min, Young-Shin Jun, and Srikanth Singamaneni Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b02772 • Publication Date (Web): 14 Sep 2018 Downloaded from http://pubs.acs.org on September 14, 2018

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

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

Page 1 of 33

Environmental Science & Technology

Photothermally-active Reduced Graphene Oxide/Bacterial Nanocellulose Composites as Biofouling-resistant Ultrafiltration Membranes Qisheng Jiang1,#, Deoukchen Ghim2,#, Sisi Cao1, Sirimuvva Tadepalli1, Keng-Ku Liu1, Hyuna Kwon2,3, Jingyi Luan1, Yujia Min2, Young-Shin Jun2* and Srikanth Singamaneni1*

1

Institute of Materials Science and Engineering and Department of Mechanical Engineering and

Materials Science, Washington University in St. Louis, St Louis, MO 63130 2

Department of Energy, Environmental and Chemical Engineering, Washington University, St.

Louis, MO 63130 3

Department of Energy and Resources Engineering, Seoul National University, Seoul, South

Korea

Address: One Brookings Drive, Campus Box 1099 (SS) and 1180 (YSJ) *E-mail: [email protected] (SS) and [email protected] (YSJ) Phone: (314) 935-5407 (SS) and (314)935-4539 (YSJ) #

Q. Jiang and D, Ghim contributed equally

Submitted: May 2018 Revised: September 2018

Environmental Science & Technology *To whom correspondence should be addressed

ACS Paragon Plus Environment

Environmental Science & Technology

1

ABSTRACT

2

By severely decreasing water flux and driving up operational costs, biofouling poses one

3

of the most serious challenges to membrane technologies. Here, we introduce a novel anti-

4

biofouling ultrafiltration membrane based on reduced graphene oxide (RGO) and bacterial

5

nanocellulose (BNC), which incoporates GO flakes into BNC in situ during its growth. In

6

contrast to previously reported GO-based membranes for water treatment, the RGO/BNC

7

membrane exhibited excellent aqueous stability under environmentally relevant pH conditions,

8

vigorous mechanical agitation/sonication, and even high pressure. Importantly, due to its

9

excellent photothermal properties, under light illumination, the membrane exhibited effective

10

bactericidal activity, obviating the need for any treatment of the feed water or external energy.

11

The novel design and in situ incorporation of the membranes developed in this study present a

12

proof-of-concept for realizing new highly efficient and environmental-friendly anti-biofouling

13

membranes for water purification.

ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

14

Environmental Science & Technology

INTRODUCTION

15

Water scarcity is recognized as one of the most critical global challenges in the 21st

16

century.1-3 In response to this dire need, various membrane technologies are being actively

17

investigated for water purification and reclamation.4-6 However, fouling and consequent

18

degradation of membranes during performance still remain a ubiquitous problem.7 The three

19

major fouling mechanisms, mineral scaling, organic fouling, and biofouling, all lead to a decline

20

in water flux.8 Among them, biofouling, which accounts for more than 45% of all membrane

21

fouling, is the Achilles heel of the technologies, due to the difficulty of completely removing

22

microorganisms.9-11

23

To inhibit the formation of biofilm, various biofouling controlling strategies, such as

24

adjusting pH, adding disinfectants and biocides, and introducing quorum sensing molecules,

25

have been suggested.12-15 However, most of these strategies introduce considerable operational

26

costs and/or potential hazardous contaminants.12

27

incorporation of nanomaterials (e.g., silver nanoparticles, TiO2 nanoparticles, and graphene oxide

28

nanosheets), polymers (e.g., polyethylene glycol, polyvinyl alcohol, polyvinylpyrrolidone, and

29

zwitterionic

30

biomacromolecules) to engineer fouling-resistant membrane surfaces that can reduce biofilm

31

growth and inactivate bacteria.8, 16-21 However, most of these methods involve additional thermal

32

or chemical treatment steps. Furthermore, most techniques are effective for only a short period

33

of time, because biofilm gradually adapts to imposed harsh environments.9 Even if 99.9% of

34

biofilm is removed, the residual cells are sufficient to grow back and form a new biofilm. Highly

35

efficient and cost-effective methods that overcome biofouling on water purification membranes

polymers),

and

other

materials

Researchers have also investigated the

(e.g.,

1 ACS Paragon Plus Environment

small

organic

molecules

and

Environmental Science & Technology

36

over a long period of time would immediately help meet the grand challenge of providing access

37

to clean water.

38

The photothermal effect of materials can offer a unique solution to biofouling, obviating

39

the need for harsh chemical treatments to achieve bacterial lysis. Photothermally-active materials

40

can effectively absorb light and then convert it into heat. In our recent study, gold nanostars

41

grown on graphene oxide (GO) flakes coating a commercial membrane were utilized as

42

nanoheaters.21 In the study, with laser irradiation, the photothermal properties of the gold

43

nanostars and GO were utilized to quickly kill adjacent Escherichia coli bacteria, inhibiting the

44

formation of biofilm on reverse osmosis membranes. While this study provided a promising

45

example of utilizing the photothermal effect to minimize biofouling on membranes, it would be

46

even more beneficial if the membrane itself were comprised of photothermal materials.

47

GO has been recognized as efficient photothermal material because the closely spaced

48

energy level from loosely bonded π electrons absorb the broad electromagnetic spectrum.22-24

49

The absorbed light energy excites electrons, and its electron relaxes to ground state through non-

50

radiative decay, releasing the energy by heat.22-24 In addition to having an intriguing

51

photothermal property, GO has frequently been employed as a membrane component owing to

52

its excellent mechanical strength, facile synthesis, and ease of attachment by vacuum filtration,

53

spin-coating, and drop-casting.25-28 In contrast to graphene flakes, which have a strong tendency

54

to stack and aggregate in aqueous solutions, GO flakes are easily dispersed, making the

55

membrane preparation process easier in aqueous media.29, 30 This dispersibility comes from rich

56

oxygen-containing functional groups (carboxyl, epoxy, hydroxyl, and carbonyl groups).31

57

However, the stability of current GO-based membranes is compromised by vigorous agitation, 2 ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33

Environmental Science & Technology

58

and pH and ionic strength variations that are within the typical range of feed waters.32 Thus,

59

there is a need to develop new scalable approaches to fabricate stable GO-based membranes.

60

Bacterial nanocellulose (BNC) is a highly pure cellulose produced by bacteria with low-

61

molecular weight sugar as a food source. Through a series of biochemical steps, the bacteria

62

form exterior cellulose nanofibers in aqueous cultures, and these fibers become entangled to

63

form a 3D network hydrogel.33 Similar to other cellulose nanofibers and cellulose nanocrystals

64

(CNC), BNC is highly attractive for membrane technologies in view of its excellent mechanical

65

properties, tunable porosity, chemical functionalizability, easy synthesis, high scalability, and

66

most important, low environmental impact.34 Therefore, BNC is a promising material for

67

fabricating functional composites through in situ growth or by adsorption of pre-synthesized

68

nanostructures on the nanoscale cellulose fibers.35-44

69

Here, we demonstrate a novel and facile approach for fabricating an anti-biofouling

70

ultrafiltration membrane, involving in situ incorporation of GO flakes into BNC during its

71

growth. The reduced graphene oxide (RGO) incorporated BNC membrane not only exhibited

72

outstanding mechanical and chemical stability under environmentally-relevant pH conditions and

73

vigorous mechanical agitation/sonication, but also showed stable water flux under high pressure.

74

Particularly, owing to its photothermal properties, the membrane exhibited light-enabled

75

bactericidal activity, avoiding the need for any treatment of the feed water or any external

76

energy. The novel design and preparation method introduced here are suggested the first steps

77

toward realizing highly efficient, environmental-friendly, and biofouling-resistant membranes for

78

water purification.

79 3 ACS Paragon Plus Environment

Environmental Science & Technology

80

EXPERIMENTAL SECTION

81

Fabrication of RGO/BNC membranes. Gluconacetobacter hansenii (ATCC®53582) was

82

employed to synthesize the cellulose nanofibers. To produce a dense bacterial suspension, the

83

bacteria were cultured in test tubes containing 16 ml of #1765 medium at 30 °C for 3 days under

84

shaking at 250 rpm. The #1765 growth medium is composed of 2% (w/v) glucose, 0.5% (w/v)

85

yeast extract, 0.5% (w/v) peptone, 0.27% (w/v) disodium phosphate, and 0.5% (w/v) citric acid.

86

To synthesize GO, we employed an efficient oxidation process, reported by Tour and co-

87

workers.45 For in situ incorporation of GO sheets in a BNC membrane, GO solution (150 mL of

88

0.0725wt%) was sonicated (2 hours), centrifuged, re-dispersed in #1765 medium, and then

89

centrifuged again to concentrate a wet mixture of GO and medium after supernatant was

90

decanted. Then the densely cultured Gluconacetobacter hansenii suspension was added to the

91

GO/medium wet mixture, making a total of 150 ml with 0.0725 wt% GO contents. The solution

92

was subsequently transferred to a Pyrex bakeware dish (18 cm × 18 cm) and incubated at room

93

temperature without disturbance. After 2 days, a thin hydrogel of GO/BNC had formed, and it

94

was harvested from the bakeware for cleaning. To remove the residual bacteria and growth

95

medium, the hydrogel was boiled in 2.5 L of 0.1 M NaOH aqueous solution for 2 h. The obtained

96

RGO/BNC hydrogel was then dialyzed in de-ionized water for 1 day. The purified RGO/BNC

97

hydrogel was dried to obtain an RGO/BNC membrane. In our preliminary test, the size and

98

contents of GO in membranes and membrane thickness affected the water flux and solute

99

rejection of RGO/BNC membranes. The specific GO concentration (0.0725 wt%), employed for

100

preparing RGO/BNC membranes, provides the best flux performance and good solute rejection.

4 ACS Paragon Plus Environment

Page 6 of 33

Page 7 of 33

Environmental Science & Technology

101

Microstructure characterization and property measurements. Scanning electron microscopy

102

(SEM, FEI Nova 2300 Field Emission SEM at an accelerating voltage of 10 kV) provided

103

micron-scale images of RGO/BNC and pristine BNC. Atomic force microscopy (AFM) images

104

were obtained for determining the thickness of GO flakes, using a Dimension 3000 (Bruker Inc.)

105

instrument in light tapping mode. To investigate the relative oxygen and carbon ratio of GO and

106

RGO flakes, carbon 1s peak was analyzed by X-ray photoelectron spectroscopy (XPS, a Physical

107

Electronics® 5000 VersaProbe II Scanning ESCA Microprobe). The pore size distribution of

108

RGO/BNC membranes was measured by the Brunauer-Emmett-Teller (BET) method using an

109

Autosorb-1C (AX1C-MP-LP) at 298 K.

110

Stability tests of RGO/BNC membrane. To study the stability of the RGO/BNC membrane, we

111

placed it in Petri dishes filled with solutions at pH 4, 7, and 9 and sonicated them for 5 hours

112

(483 W, 8892, Cole-Parmer). The pH values were chosen because they occur in many natural and

113

engineering aqueous systems.46 Subsequently, the release of RGO from the membrane was

114

quantified from the UV-Vis absorbance spectra (Shimadzu UV-1800 spectrophotometer, 400 nm

115

to 1000 nm) of the solutions. SEM was used to monitor the surface morphologies of the

116

RGO/BNC film after sonication. To study the mass change of RGO/BNC membrane before and

117

after sonication, thermo-gravimetric analysis (TGA) was performed using a TA Instruments

118

Q5000 IR Thermogravimetric Analyzer in air (at rate of 5 °C/min).

119

Because the GO membranes were frequently prepared by vacuum filtration, we made two

120

types of control samples for comparison.26-28 First, a similar amount of base-washed RGO flakes

121

was deposited on top of a BNC hydrogel, using the vacuum-assisted method, and dried to obtain

122

a dry film. Second, BNC dispersions were pre-mixed with GO solutions and were then filtered to 5 ACS Paragon Plus Environment

Environmental Science & Technology

123

make membranes. Both control membranes were subjected to the same aqueous stability tests as

124

the RGO/BNC membrane (Figure S1 in Supporting Information).

125

Evaluation of mass transport performance of RGO/BNC membranes. The mass transport

126

performance of RGO/BNC membranes was evaluated to estimate pore size of RGO/BNC by

127

using a side-by-side diffusion cell system because it can use much smaller sample volume (~10

128

mL) than that of benchtop cross-flow system (~15 L). An RGO/BNC membrane was first

129

mounted between the two cells (Adams & Chittenden Scientific Glass, 5 mL volume). Then,

130

ethanol and water were used to rinse the membrane several times to avoid subsequent air bubble

131

formation. To test diffusion-driven transport through an RGO/BNC membrane, 0.5 mM of

132

rhodamine 6G (R6G, ~ 1 nm, 479 Da) and lysozyme (3.8–4 nm, 14300 Da) were used. The

133

observation of transport behavior of two model solutes (R6G and lysozyme) across RGO/BNC

134

membrane helps to reveal approximate pore size of RGO/BNC membrane, and consequently,

135

determines filtration type of RGO/BNC. R6G is cationic dye, thus it is positively charged.47 The

136

isoelectric points for lysozyme are 10.5, and its pH value at 0.5 mM concentration is 3.53,

137

indicating that lysozyme is positively charged under our experimental condition.48 The solute

138

was introduced on the feed side, while the dialysate side was just DI water. Solutions in both

139

cells were subjected to vigorous stirring to minimize concentration polarization effects close to

140

the membrane. The diffusing concentration of the solute was monitored in the wavelength range

141

of 300−700 nm, using a Shimadzu UV-1800 spectrometer. We utilized the RGO/BNC

142

membranes grown in independent batches, and diffusion tests were conducted with three

143

replicates.

6 ACS Paragon Plus Environment

Page 8 of 33

Page 9 of 33

Environmental Science & Technology

144

Ultrafiltration efficiency and flux tests. The water fluxes of the RGO/BNC membrane and

145

commercial ultrafiltration membrane (YMGESP3001, GE) were tested using a benchtop cross-

146

flow system, and then compared those permeate flux. The commercial ultrafiltration membrane,

147

used for comparison, was designed for pre-treatment, dye reduction, and purification with 1,000

148

Dalton cut-off sizes. The benchtop cross-flow system included a crossflow membrane cell

149

(CF042D, Sterlitech Corp.) and a Hydracell pump (M03S, Wanner Engineering, Inc). During the

150

measurement of the permeate flux, we set the pressure and the feed flux at 100 psi and 0.66

151

L/min, with 25 °C water. For the experiments, we utilized the RGO/BNC membranes grown in

152

independent batches and measured the water flux with three replicates. Furthermore, gold

153

nanoparticles (AuNPs) with a diameter around 5 nm were synthesized using the seed-mediated

154

growth method,49 and their size distribution was determined from TEM images. AuNP solutions

155

were then filtered by RGO/BNC membranes, using above cross-flow system under 100 psi.

156

Before and after filtration, AuNP concentrations in filtration/permeate solutions were measured

157

by UV-Vis spectrometry (400 nm to 1000 nm). The rejection rate (RR) was calculated using the

158

equation below:

159

ܴܴ =

ா೑ ா೛

× 100,

160

where ‫ܧ‬௙ is the optical extinction of the feed solution, and ‫ܧ‬௣ is the optical extinction of the

161

permeate solution.

162

Photothermal and bactericidal performance of RGO/BNC membranes under illumination.

163

The Photothermal performance of the membranes was tested using a solar simulator (Newport

164

66921, Arc Lamp). Both the RGO/BNC membrane and BNC membrane were illuminated at a

7 ACS Paragon Plus Environment

Environmental Science & Technology

165

power density of 2.9 kW/m2 for 180 sec. The temperature map of the surface of both membranes

166

under water was monitored by an IR camera (Ti 100, FLUKE). To test bactericidal activity, MG

167

1655 E. coli was grown in Luria-Bertani liquid medium at 37 °C. All cultures were in 125 mL

168

baffled shake flasks (25 mL working volume, shaking at 225 rpm). Cells in log phase (>108 live

169

cells/ mL) were harvested after 12 h of incubation and then used for bactericidal tests. A layer of

170

MG 1655 E. coli biofilm was grown on the RGO/BNC and BNC membrane surfaces, and then

171

exposed to simulated sunlight for 180 secs at 2.9 kW/m2. Before/after light illumination, the

172

biofilms were exposed to fluorescent dyes (Molecular Probes Live/Dead Bacterial cell viability

173

kit) for 30 min, and then imaged under a Leica microscope to identify live (blue fluorescent

174

filter, 340-380 nm) and dead (green fluorescent filter, 450-490 nm) cells.50

175

RESULTS AND DISCUSSION

176

Characterization of the RGO/BNC membranes

177

The RGO/BNC membranes were fabricated by in situ incorporation of GO flakes within

178

the BNC network via bacteria-mediated growth under aerobic and static conditions (Figure 1).

179

To determine the average thickness of synthesized GO flakes, they were deposited on a silicon

180

substrate and measured with AFM. The thicknesses were ~ 1.0 ± 0.2 nm, corresponding to a

181

bilayer of GO (the thickness of a monolayer is ~ 0.7 nm ) (Figure 2A).51 After washing, GO

182

flakes were dispersed in broth solution with bacteria at an optimized concentration to achieve a

183

desired BNC growth rate (Figure 1A). The GO dispersed solution was left undisturbed under

184

ambient conditions to obtain GO/BNC hydrogels. To remove bacteria and residual broth solution

185

from in situ grown GO/BNC hydrogel, it was immersed in NaOH solution (0.1 M) at boiling

8 ACS Paragon Plus Environment

Page 10 of 33

Page 11 of 33

Environmental Science & Technology

186

temperature, which partially reduced the GO flakes (discussed in detail below). The cleaned

187

RGO/BNC hydrogel was dried to obtain a large, robust RGO/BNC membrane (Figure 1B).

188

To understand the chemical reduction of GO during cleaning, GO flakes were added to a

189

boiling temperature 0.1 M NaOH solution, which turned uniformly distributed GO particles into

190

black aggregated particles, indicating the partial reduction and restacking of the RGO. To

191

confirm the reduction of GO after exposure to the basic solution, X-ray photoelectron

192

spectroscopy (XPS) was utilized (Figure 2B, 2C). The high resolution 1s spectra of carbon were

193

obtained and were deconvoluted into three peaks, corresponding one sp2 domain (C=C with a

194

binding energy of 284.6 eV) and two oxidized sp3 domains (C–O with a binding energy of 286.6

195

eV, and C=O with a binding energy of 288.2 eV).52 The relative carbon and oxygen ratio was

196

calculated based on the peak area, and this ratio was utilized to estimate the reduction extent of

197

GO. For synthesized GO, the C/O ratio was 1.7, suggesting that ~ 58% of the GO surface was

198

oxidized (Figure 2B). After an extensive base wash, the C/O ratio increased to 4.6, indicating

199

that ~37% of the oxygen functional groups were reduced (Figure 2C). This result confirms that

200

base washing to kill residual bacteria also reduces GO flakes in the BNC matrix.

201

In the absence of GO flakes, bacteria-mediated growth results in a white, translucent

202

BNC membrane with outstanding flexibility and mechanical strength (Figure 2D).53 Because

203

bacteria-mediated synthesis of nanocellulose requires oxygen, a dense network of nanocellulose

204

fibers (20 –100 nm in diameter) forms near the air/liquid interface, where abundant oxygen is

205

available (Figure 2E). As the oxygen diffuses deeper into the medium, the first dense layer sinks,

206

and makes way for the formation of subsequent BNC layers, which, upon sinking, stack together

207

to form a 3D BNC network.33, 44 Due to this “layer-by-layer” formation, the cellulose nanofibrils 9 ACS Paragon Plus Environment

Environmental Science & Technology

208

are preferentially oriented parallel to the surface (i.e., normal to the thickness) of the membrane,

209

which results in denser physical entanglements of the cellulose nanofibers parallel to the surface

210

(Figure 2F). To form RGO/BNC membranes, GO flakes were added during in situ growth, and a

211

subsequent base washing process was conducted (Figure 2G). Compared to pristine BNC

212

membrane, the RGO/BNC membrane is smoother and has fewer fibrils protruding from the

213

surface, due to the presence of 2D RGO sheets (Figure 2H). For quantitative comparisons,

214

nanoscale surface roughness of BNC and RGO/BNC membranes were measured by AFM, and

215

the results showed the surface roughness of 32.4 ± 9.46 nm and 25.0 ± 14.2 nm, respectively,

216

indicating the rougher surface of BNC membranes (Figure S6). Cross-sectional SEM images

217

showed that the membrane was ~8 µm thick (Figure 2I). The images also show the embedded

218

RGO flakes between BNC layers due to the “layered” formation of BNC, starting from the

219

liquid/air interface as described above (Figure 2I). Because the hydrophilicity is an important

220

property to influence the water flux of membrane, the contact angles of RGO/BNC and BNC

221

membranes were measured (Figure S6). Due to reduction of hydrophilic GO flakes during

222

inevitable cleaning process, RGO/BNC membrane showed higher contact angle (65 ± 13°) than

223

that of BNC membrane (23 ± 5°).

224

Mechanical and chemical stability of the RGO/BNC membranes

225

To investigate the mechanical and chemical stability of the membrane, we exposed it to

226

ultrasonic agitation for 5 hours in solutions at pH 4, 7, and 9 (see Figure S1 in Supporting

227

Information). Even after this vigorous mechanical agitation at environmentally-relevant pH

228

conditions, the membrane did not exhibit any signs of disintegration or loss of RGO flakes

229

(Figure 3A and its inset). This result was further supported by SEM images of the membrane

10 ACS Paragon Plus Environment

Page 12 of 33

Page 13 of 33

Environmental Science & Technology

230

surface, which did not show a discernable change in the morphology after sonication (Figure

231

3B). Then, to evaluate the thermal stability of the RGO/BNC membrane, TGA was conducted.

232

The membrane showed a first mass loss of ~2 wt.% at ~100 °C, due to the loss of absorbed

233

water. A second mass loss of ~3 wt.% at ~200 °C was attributed to the decomposition of

234

functional groups of GO.54 A third mass loss of ~46 wt.% began at 280°C, and was due to the

235

degradation of cellulose. A fourth mass loss (~49 wt.%) at 390 °C was attributable to the

236

continued decomposition of cellulose residual and sublimation or burning of damaged graphitic

237

regions.55 Based on the mass loss profiles of the RGO/BNC membrane, RGO, and pristine BNC

238

membrane, the mass loading of RGO in RGO/BNC was calculated to be ~45 wt.%, and it

239

suggested excellent thermal stability of the RGO/BNC membrane up to 200 °C. The RGO/BNC

240

membrane after ultrasonic treatment also showed an identical mass loss profile, implying that the

241

embedded RGO flakes within the BNC matrix remained intact (Figure 3C).

242

As mentioned above, GO-based membranes have been extensively investigated in the

243

past several years. Most of these membranes were fabricated by vacuum filtration of GO flakes

244

onto a supporting membrane, but this coating method always leaves doubt about its long-term

245

aqueous stability.32 Here, we compared the mechanical stability of in situ embedded RGO/BNC

246

with that of an RGO/BNC membrane prepared by depositing RGO particles (base-washed) on

247

top of a BNC membrane using vacuum filtration. After 5 h of ultrasonic agitation in solutions at

248

pH 4, 7, and 9, the RGO particles had disintegrated completely, and the solution exhibited broad

249

absorbance, a feature of RGO flakes in solution (Figure 3D). This disintegration was further

250

confirmed by surface SEM images and quantified by TGA. After sonication, the membrane

251

prepared through vacuum filtration showed an initial mass loss (~5%) at 100 °C, attributed to

11 ACS Paragon Plus Environment

Environmental Science & Technology

252

absorbed water, and the decomposition of cellulose at ~280 °C (~63%) and at ~340 °C (~32%).

253

All of these findings indicated the absence of RGO flakes (Figure 3F).

254

Conventional GO-based membranes rely on hydrogen bonding or metal ion incorporation

255

for mechanical stability. However, mechanical agitation during water filtration or cleaning

256

procedures, as well as pH variations in the feed water, can compromise the mechanical stability

257

of these membranes (Figure S1 in Supporting Information). In contrast, in situ RGO/BNC

258

membranes are robust because during the BNC growth, plate-like GO particles are physically

259

locked into the layered BNC matrices, which provides its excellent mechanical and chemical

260

stability. Moreover, intensive vortexing has been employed to wash used in situ RGO/BNC

261

membranes, and did not cause any discernable damage on these membranes, suggesting the

262

mechanical stability under the mechanical stress (Figure S1D). This finding further suggests the

263

durability of the in situ RGO/BNC membrane during potential cleaning process.

264

Mass transport performance and water flux tests

265

To probe the diffusive transport capability of small molecules across the RGO/BNC

266

membranes, we employed a two-cell setup (Figure 4A) and tested it with 0.5 mM of two model

267

solute systems, having different sizes and molecular weights: rhodamine 6G (R6G, ~ 1 nm, 479

268

Da) and lysozyme (3.8–4 nm, 14300 Da).56 Because ultrafilters have pore sizes between 1 nm to

269

100 nm, and thus remove contaminants via a size exclusion mechanism, these two different

270

solutes helped to determine the pore size and filtration capability of RGO/BNC membranes.57 A

271

UV-vis spectrometer was used to monitor the concentration of model solutes from the feed side

272

to the permeate side (see experimental section for details). For pristine BNC membrane, all two

273

solutes rapidly diffused through because the BNC fiber network is composed of micro-scale

12 ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33

Environmental Science & Technology

274

pores even if the nanofibers are densely packed (Figure 4B). By contrast, the addition of

275

graphene oxide flakes within the matrix completely blocked the diffusion of lysozyme (3.8–4

276

nm), although R6G (~1 nm) passed through (Figure 4B). These results are in accordance with

277

the BET measurements which indicate the pore size of the RGO/BNC membrane is around 2.2

278

nm (radius), which falls into the range of ultrafiltration membranes (Figure S3).58 Moreover, the

279

unique mass transport properties of GO-based membranes originate from the nanocapillary

280

network formed by lamellar stacking of GO, and the mass-transport behavior can be adjusted by

281

tuning functional groups or inserting external species with desired dimensions. 25,28 In this work,

282

the presence of bacterial cellulose nanofibers between RGO flakes will lead to an overall

283

tortuous network of pores in the membrane, even though no visible pores can be seen in a SEM

284

image of the surface of RGO/BNC membranes.

285

To further demonstrate the potential of the novel RGO/BNC membrane for an

286

ultrafiltration system, we performed flux tests and particle rejection tests using size-controlled

287

gold nanoparticles (AuNPs) via a benchtop cross-flow system (setup diagram in Figure 4C). For

288

a particle rejection test, spherical gold nanoparticles with diameters of 5.15 ± 0.4 nm were

289

synthesized using the seed-mediated method.49 The prepared AuNPs showed a very narrow size

290

distribution (RSD < 8 %), which made the particle rejection study accurate (Figure S2B in

291

Supporting Information). The particle rejection rates were calculated by measuring the UV-vis

292

extinction spectra of solutions before/after filtration through the RGO/BNC membranes (Figure

293

4D). This result suggested that AuNP of 5 nm diameter were ~100% rejected (inset of Figure

294

4D). In the same way, the rejection test was also performed for a commercial ultrafiltration

13 ACS Paragon Plus Environment

Environmental Science & Technology

295

membrane (pore size of around 1.66 nm), which also showed ~ 100% rejection for 5 nm gold

296

nanoparticles (Figure S2C in Supporting Information).

297

Under 100 psi, the RGO/BNC membranes showed higher water fluxes than a commercial

298

ultrafiltration (UF) membrane. Because the pore size for both membrane types belongs to the

299

range between UF and nanofiltration (NF), we intentionally tested water flux at a higher

300

operating pressure than the usual operating pressure for UF (7.3–73 psi). Over a five-hour-long

301

flux test after stabilization, the water flux of the RGO/BNC membrane was found to be 52.6 ±

302

2.5 L/m2 h, and that of the commercial ultrafiltration membrane was 21.6 ± 0.8 L/m2 h (Figure

303

4E). Importantly, the RGO/BNC membrane withstood an operating pressure as high as 100 psi

304

without any supporting membrane. This performance emphasized the remarkable mechanical

305

strength of RGO/BNC, considering that most of the GO-based membranes reported in the

306

literature require a support membrane or a carefully designed apparatus due to their limited

307

mechanical strength.59-63

308

Photothermal and bactericidal performance under illumination

309

Next, we examined the photothermal and bactericidal ability of the RGO/BNC membrane

310

using a Newport 66921 Arc Lamp with a power density of 2.9 kW/m2. We used IR imaging to

311

monitor the temperature profile of the RGO/BNC membrane in an aqueous environment during

312

illumination (Figure 5B). The reduction of GO flakes during inevitable sample preparation

313

(cleaning) process provides the benefits for better photothermal conversion efficiency because

314

the abundant delocalized π electrons in conjugated sp2-bonded carbon create the closely spaced

315

energy levels.22-24 Upon illumination, the temperatures of the RGO/BNC membrane rapidly

316

increased from room temperature (26 °C) to ~60 °C (Figure 5C). Specifically, the temperature

14 ACS Paragon Plus Environment

Page 16 of 33

Page 17 of 33

Environmental Science & Technology

317

rapidly increased during the first 20 seconds after the onset of irradiation and remained constant

318

over the entire duration (120 sec). In comparison, the pristine BNC membranes showed only a

319

small temperature increase (3 °C). The large rise in temperature was caused by the broad optical

320

absorption of many RGO flakes arranged within the BNC matrix. Once light was absorbed by

321

the RGO flakes, they immediately generated heat, which then dissipated to the surrounding water

322

and BNC.44 Due to the decrease of thermal radiation along the distance between IR camera and

323

surface (~30 cm), the temperature profile obtained from the IR camera underestimated the actual

324

temperature at the surface of the RGO/BNC membrane. Therefore, biological species were

325

exposed to temperature higher than ~60 °C. To test whether the heat generated by light exposure

326

would affect the mass transport and water flux performance of the RGO/BNC membrane, we

327

performed BET and water flux tests before and after long duration of light exposure (2.9

328

kW/m2).

329

membrane remained stable with minor variation after light exposure (Figures S3 and S4).

Both the pore size distribution and the water flux performance of RGO/BNC

330

To test the bactericidal ability, the RGO/BNC membrane was covered with stained E. coli

331

bacteria from a live/dead cell viability assay, and then light was shined on the membrane surface.

332

Before irradiation, both RGO/BNC and pristine BNC membranes showed substantial and well

333

distributed green fluorescence, corresponding to live bacteria, and no sign of red fluorescence,

334

indicating the absence of dead bacteria (Figure 5D). After irradiation (2.9 kW/m2) for 180 secs,

335

the bacteria on the RGO/BNC membrane exhibited predominantly red fluorescence (dead

336

bacteria) and a complete visible absence of green fluorescence. However, the E.coli-covered

337

pristine BNC membrane exhibited green fluorescence corresponding to live bacteria, even after

338

irradiation (Figure 5E). The SEM images showed morphology changes and leakage of bacteria,

339

indicating that the high temperature at the RGO/BNC membrane had disrupted the cell walls and 15 ACS Paragon Plus Environment

Environmental Science & Technology

340

cell membranes (Figure 5F, G). Before light irradiation, live E. coli bacteria on an RGO/BNC

341

membrane showed a rod-like structure, while after irradiation they were significantly shrunken

342

and wrinkled. Because the cell walls of E. coli are known to deteriorate near 70 °C,64 this change

343

implied that upon illumination, the surface of the RGO/BNC membrane had rapidly heated to

344

above 70 °C and killed bacteria within a short time (180 sec). Live E. coli bacteria on an

345

RGO/BNC membrane without illumination showed no dead bacteria cells even after 1 hour

346

exposure, indicating the bactericidal activity of RGO/BNC membrane only occur upon light

347

illumination (Figure S5). Here, we demonstrated the excellent bactericidal performance of the

348

RGO/BNC membrane in harvesting light, and this capability makes the RGO/BNC membrane

349

highly attractive for energy-saving and environmentally-friendly water purification applications.

350

ENVIRONMENTAL IMPLICATIONS

351

In this study, we presented an innovative approach that uses the photothermal effect of

352

RGO by embedding it in BNC structures. This new type of membrane can enhance the stability

353

and durability of a membrane and inhibit or delay microorganism growth on its surface. While

354

the most contemporary approaches to resisting biofouling rely on temporary chemical treatments,

355

combining the photothermal effect with a noble membrane design shows that anti-biofouling can

356

be achieved with sustainable and abundant sunlight.

357

We note that fully utilizing the photothermal property of RGO/BNC membranes can be

358

challenging. A possible implementation involves modifying a spiral-wound module system, as

359

shown in Figure 5A. The inner and outer surfaces of the membrane modules or feed channel

360

spacers can be equipped with low-energy LEDs for illumination, and can potentially be powered

361

by renewable energy sources, such as low-cost photovoltaic devices or triboelectric

16 ACS Paragon Plus Environment

Page 18 of 33

Page 19 of 33

Environmental Science & Technology

362

nanogenerators (TENGs).65 In TENGs, mechanical energy generated through fluid flow in the

363

UF process can be harnessed to produce light and to heat the membrane surface, reducing the

364

overall operational expense. In addition to spiral-wound module system, this suggested system

365

can also be applied to plate sheet membrane module by adding LED-equipped plates between the

366

membrane modules. This would be an exciting future research direction for realizing

367

photothermal membrane water filtration.

368

Here we demonstrated the anti-biofouling property of RGO/BNC membrane, which

369

originated from localized surface heating by photothermal effect of RGO particles. While we

370

successfully inhibit the biofilm growth on membrane by inactivating the bacteria, systematic

371

tests for the effects of extracellular polymeric substance and soluble microbial products

372

potentially produced by microorganism, mineral scaling, and organic fouling on membrane’s

373

performance (e.g., localized heating on the fouling behavior of membranes) should be conducted

374

in the future.

375

The novel fabrication method of incorporating RGO during ‘layer-by-layer’ growth of

376

BNC yields a well stacked the structure, with a pore size in the UF membrane range. In addition,

377

BNC production is considered eco-friendly because it needs only low-molecular weight sugar

378

and oxygen as food sources. Until now, many researchers prepared GO membranes through

379

vacuum filtration or spin coating without a polymer matrix25-28, but these fabrication methods

380

inevitably raised the mechanical stability concerns. However, the RGO/BNC membrane,

381

reported here, exhibits stable water flux under 100 psi loading and maintains chemical stability at

382

solution pH varying from 4 to 9. The water flux is higher than that of commercial UF membranes

383

under identical pressure. In summary, RGO/BNC membrane demonstrated here not only

17 ACS Paragon Plus Environment

Environmental Science & Technology

384

provides a novel anti-biofouling approach powered by solar energy, but also suggests a scalable,

385

eco-friendly, and cost-effective way to fabricate UF membranes for water purification.

386

SUPPORTING INFORMATION

387

The Supporting Information is available free of charge on the ACS Publications Website at DOI:

388

Additional physical and chemical stability test (Figure S1), AuNPs rejection data (Figure S2),

389

nitrogen isotherm and pore size distribution (Figure S3), water flux test under cyclic illumination

390

(Figure S4), bacterial cell viability and fluorescence images (Figure S5), and surface roughness

391

and contact angles (Figure S6) were included.

392

ACKNOWLEDGEMENTS

393

The authors acknowledge support from National Science Foundation Environmental Engineering

394

Program (CBET- 1604542) and Air Force Office of Scientific Research (Award # FA9550-15-1-

395

0228). H. Kwon received a fellowship from Washington University’s McDonnell Academy

396

Global Energy and Environment Partnership. The authors thank Prof. Lihong Wang from the

397

Department of Biomedical Engineering at Washington University for providing access to an IR

398

camera, and the Nano Research Facility (NRF) at Washington University for providing access to

399

electron microscopy facilities. The authors acknowledge Professor James C. Ballard for carefully

400

reviewing the manuscript.

18 ACS Paragon Plus Environment

Page 20 of 33

Page 21 of 33

Environmental Science & Technology

401

List of Figures

402

Figure 1. Fabrication of RGO/BNC membrane. Photographs show (A) GO in bacterial medium

403

and (B) in situ grown RGO/BNC membrane after cleaning and drying.

404

Figure 2. Chemical composition and microstructure of RGO/BNC membrane. (A) AFM image

405

of GO flakes deposited on a silicon substrate. X-ray photoelectron spectra of (B) pristine and (C)

406

base-washed GO. Photograph (D) and SEM images of (E) the surface and (F) a cross-section of a

407

pristine BNC membrane. Photograph (G) and SEM images of (H) the surface and (I) a cross-

408

section of an RGO/BNC membrane.

409

Figure 3. Stability of RGO/BNC membrane. UV-vis absorption spectra of (A) RGO/BNC and

410

(D) RGO-coated BNC immersed solutions (at pH 7) before and after ultrasonic agitation for 5 h.

411

The insets are photographs of an RGO/BNC membrane and an RGO-coated BNC membrane

412

after sonication. SEM images of (B) an RGO/BNC membrane and (E) an RGO-coated BNC

413

membrane after ultrasonic agitation. TGA analyses of (C) an RGO/BNC membrane and (F) an

414

RGO-coated BNC membrane before and after ultrasonic agitation.

415

Figure 4. Mass transport performance and water flux tests. (A) Schematic diagram of a two-cell

416

diffusion setup. (B) Diffusion of model solutes through pristine BNC and RGO/BNC

417

membranes. The average values of concentrations in the diffused part were obtained from three

418

replicates for diffusion studies. (C) Schematic diagram of cross-flow flux test setup. RGO/BNC

419

membranes were placed in between the cross-flow cell and tightly sealed. (D) UV-vis extinction

420

spectra indicating the rejection of 5 nm AuNPs filtered through RGO/BNC membranes using the

421

cross-flow system under 100 psi (inset shows feed and permeate solutions). (E) Water flux of

19 ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 33

422

RGO/BNC membranes (~8 µm thick) and commercial ultrafiltration membranes with three

423

replicates. 100 psi was applied for the flux tests.

424

Figure 5. Photothermal and bactericidal performance under illumination. (A) Schematic showing

425

antifouling mechanism of RGO/BNC membrane and a possible configuration of a spiral-wound

426

UF module coupled with LEDs. (B) IR images showing the temperature of the pristine BNC and

427

the RGO/BNC membrane in water under illumination at various time points. (C) Plot showing

428

the temperature of pristine BNC and the RGO/BNC membrane in water under 2.9 kW/m2

429

illumination, as a function of irradiation time. Fluorescence images of E. coli on BNC and

430

RGO/BNC membranes (D) before and (E) after irradiation.

431

RGO/BNC membranes (F) before and (G) after irradiation.

20 ACS Paragon Plus Environment

SEM images of E. coli on

Page 23 of 33

432

Environmental Science & Technology

Figure 1.

21 ACS Paragon Plus Environment

Environmental Science & Technology

433

Figure 2.

22 ACS Paragon Plus Environment

Page 24 of 33

Page 25 of 33

434

Environmental Science & Technology

Figure 3.

23 ACS Paragon Plus Environment

Environmental Science & Technology

435 436

Figure 4.

24 ACS Paragon Plus Environment

Page 26 of 33

Page 27 of 33

437

Environmental Science & Technology

Figure 5.

25 ACS Paragon Plus Environment

Environmental Science & Technology

438

TOC Art:

439

26 ACS Paragon Plus Environment

Page 28 of 33

Page 29 of 33

Environmental Science & Technology

440

References

441 442

1. Elimelech, M. The global challenge for adequate and safe water. J. Water Supply: Res. Technol.-AQUA, 2006, 55 (1), 3-10.

443 444

2. Gleick, P. H.; et al. The world's Water: The Biennial Report on Freshwater Resources; Island Press: Washington, DC, USA, 2014.

445

3.

446 447 448

4. Shannon, M. A.; Bohn, P. W.; Elimelech, M.; Georgiadis, J. G.; Marinas, B. J.; Mayes, A. M. Science and technology for water purification in the coming decades. Nature 2008, 452 (7185), 301-310.

449 450

5. Singh, R., Hankins, N., Eds. Emerging Membrane Technology for Sustainable Water Treatment; Elsevier: Amsterdam, 2016.

451 452

6. Pendergast, M. M.; Hoek, E. M. V. A review of water treatment membrane nanotechnologies. Energy Environ. Sci. 2011, 4 (6), 1946-1971.

453 454

7. Le-Clech, P.; Chen, V.; Fane, T. A. G. Fouling in membrane bioreactors used in wastewater treatment. J. Membr. Sci. 2006, 284 (1–2), 17-53.

455 456 457

8. Zhang, R.; Liu, Y.; He, M.; Su, Y.; Zhao, X.; Elimelech, M.; Jiang, Z. Antifouling membranes for sustainable water purification: strategies and mechanisms. Chem. Soc. Rev. 2016, 45 (21), 5888-5924.

458 459

9. Flemming, H. C.; Schaule, G.; Griebe, T.; Schmitt, J.; Tamachkiarowa, A. Biofouling— the Achilles heel of membrane processes. Desalination 1997, 113 (2), 215-225.

460 461

10. 28.

462 463

11. Baker, J. S.; Dudley, L. Y. Biofouling in membrane systems - A review. Desalination 1998, 118 (1-3), 81-89.

464 465 466

12. Nguyen, T.; Roddick, F. A.; Fan, L. Biofouling of Water Treatment Membranes: A Review of the Underlying Causes, Monitoring Techniques and Control Measures. Membranes 2012, 2 (4), 804-840.

467 468

13. Mansouri, J.; Harrisson, S.; Chen, V. Strategies for controlling biofouling in membrane filtration systems: challenges and opportunities. J. Mater. Chem. 2010, 20 (22), 4567-4586.

469 470 471

14. Barnes, R. J.; Low, J. H.; Bandi, R. R.; Tay, M.; Chua, F.; Aung, T.; Fane, A. G.; Kjelleberg, S.; Rice, S. A. Nitric Oxide Treatment for the Control of Reverse Osmosis Membrane Biofouling. Appl. Environ. Microbiol. 2015, 81 (7), 2515-2524.

Eliasson, J. The Rising Pressure of Global Water Shortages. Nature 2014, 517 (7532), 6.

Komlenic, R. Rethinking the causes of membrane biofouling. Filtr. Sep. 2010, 47 (5), 26-

27 ACS Paragon Plus Environment

Environmental Science & Technology

472 473

15. Lade, H.; Paul, D.; Kweon, J. H. Quorum Quenching Mediated Approaches for Control of Membrane Biofouling. Inter. J. Biol. Sci. 2014, 10 (5), 550-565.

474 475 476

16. Yang, H.-L.; Lin, J. C.-T.; Huang, C. Application of nanosilver surface modification to RO membrane and spacer for mitigating biofouling in seawater desalination. Water Res. 2009, 43, (15), 3777-3786.

477 478

17. Kumar, A.; Vemula, P. K.; Ajayan, P. M.; John, G. Silver-nanoparticle-embedded antimicrobial paints based on vegetable oil. Nat. Mater. 2008, 7 (3), 236-241.

479 480 481

18. Perreault, F.; Tousley, M. E.; Elimelech, M. Thin-Film Composite Polyamide Membranes Functionalized with Biocidal Graphene Oxide Nanosheets. Environ. Sci. Technol. Lett. 2014, 1 (1), 71-76.

482 483

19. Perreault, F.; Fonseca de Faria, A.; Elimelech, M. Environmental applications of graphene-based nanomaterials. Chem. Soc. Rev. 2015, 44 (16), 5861-5896.

484 485

20. Kochkodan, V.; Hilal, N. A comprehensive review on surface modified polymer membranes for biofouling mitigation. Desalination 2015, 356, 187-207.

486 487 488

21. Ray, J. R.; Tadepalli, S.; Nergiz, S. Z.; Liu, K.-K.; You, L.; Tang, Y.; Singamaneni, S.; Jun, Y.-S. Hydrophilic, Bactericidal Nanoheater-Enabled Reverse Osmosis Membranes to Improve Fouling Resistance. ACS Appl. Mater. Interfaces 2015, 7 (21), 11117-11126.

489 490

22. Liu, G.; Xu, J.; Wang, K. Solar water evaporation by black photothermal sheets. Nano Energy 2017, 40, 269-284.

491 492

23. Bond, T. C.; Bergstrom, R. W. Light absorption by carbonaceous particles: An investigative review. Aerosol Sci. Technol. 2006, 40 (1), 27-67.

493 494 495

24. Wang, P. Emerging investigator series: the rise of nano-enabled photothermal materials for water evaporation and clean water production by sunlight. Environ. Sci.: Nano 2018, 5 (5), 1078-1089.

496 497 498

25. Nair, R. R.; Wu, H. A.; Jayaram, P. N.; Grigorieva, I. V.; Geim, A. K. Unimpeded Permeation of Water Through Helium-Leak–Tight Graphene-Based Membranes. Science 2012, 335 (6067), 442-444.

499 500

26. Joshi, R. K.; Alwarappan, S.; Yoshimura, M.; Sahajwalla, V.; Nishina, Y. Graphene oxide: the new membrane material. Appl. Mater. Today 2015, 1 (1), 1-12.

501 502 503

27. Sun, P.; Wang, K.; Zhu, H., Recent Developments in Graphene-Based Membranes: Structure, Mass-Transport Mechanism and Potential Applications. Adv. Mater. 2016, 28 (12), 2287-2310.

504 505

28. Mi, B. Graphene oxide membranes for ionic and molecular sieving. Science 2014, 343 (6172), 740-742. 28 ACS Paragon Plus Environment

Page 30 of 33

Page 31 of 33

Environmental Science & Technology

506 507

29. Konkena, B.; Vasudevan, S. Understanding aqueous dispersibility of graphene oxide and reduced graphene oxide through pKa measurements. J. Phys. Chem. Lett. 2012, 3 (7), 867-872.

508 509

30. Li, D.; Müller, M. B.; Gilje, S.; Kaner, R. B.; Wallace, G. G. Processable aqueous dispersions of graphene nanosheets. Nat. Nanotechnol. 2008, 3 (2), 101.

510 511

31. Dreyer, D. R.; Park, S.; Bielawski, C. W.; Ruoff, R. S. The chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39 (1), 228-240.

512 513

32. Yeh, C.-N.; Raidongia, K.; Shao, J.; Yang, Q.-H.; Huang, J. On the origin of the stability of graphene oxide membranes in water. Nat Chem. 2015, 7 (2), 166-170.

514 515 516

33. Klemm, D.; Kramer, F.; Moritz, S.; Lindström, T.; Ankerfors, M.; Gray, D.; Dorris, A. Nanocelluloses: A New Family of Nature-Based Materials. Angew. Chem. Int. Ed. 2011, 50 (24), 5438-5466.

517 518

34. Carpenter, A. W.; de Lannoy, C.-F.; Wiesner, M. R. Cellulose Nanomaterials in Water Treatment Technologies. Environ. Sci. Technol. 2015, 49 (9), 5277-5287.

519 520 521

35. Wu, Z.-Y.; Liang, H.-W.; Chen, L.-F.; Hu, B.-C.; Yu, S.-H. Bacterial Cellulose: A Robust Platform for Design of Three Dimensional Carbon-Based Functional Nanomaterials. Acc. Chem. Res. 2016, 49 (1), 96-105.

522 523 524

36. Tian, L.; Jiang, Q.; Liu, K.-K.; Luan, J.; Naik, R. R.; Singamaneni, S. Bacterial Nanocellulose-Based Flexible Surface Enhanced Raman Scattering Substrate. Adv. Mater. Interfaces 2016, 3 (15), 1600214.

525 526 527

37. Tian, L.; Luan, J.; Liu, K.-K.; Jiang, Q.; Tadepalli, S.; Gupta, M. K.; Naik, R. R.; Singamaneni, S. Plasmonic Biofoam: A Versatile Optically Active Material. Nano Lett. 2016, 16 (1), 609-616.

528 529 530

38. Chen, L.-F.; Huang, Z.-H.; Liang, H.-W.; Gao, H.-L.; Yu, S.-H. Three-Dimensional Heteroatom-Doped Carbon Nanofiber Networks Derived from Bacterial Cellulose for Supercapacitors. Adv. Funct. Mater. 2014, 24 (32), 5104-5111.

531 532 533

39. Chen, L.-F.; Zhang, X.-D.; Liang, H.-W.; Kong, M.; Guan, Q.-F.; Chen, P.; Wu, Z.-Y.; Yu, S.-H. Synthesis of Nitrogen-Doped Porous Carbon Nanofibers as an Efficient Electrode Material for Supercapacitors. ACS Nano 2012, 6 (8), 7092-7102.

534 535 536

40. Tong, S.; Zheng, M.; Lu, Y.; Lin, Z.; Zhang, X.; He, P.; Zhou, H. Binder-free carbonized bacterial cellulose-supported ruthenium nanoparticles for Li-O2 batteries. Chem. Commun. 2015, 51 (34), 7302-7304.

537 538 539

41. Wang, B.; Li, X.; Luo, B.; Yang, J.; Wang, X.; Song, Q.; Chen, S.; Zhi, L. Pyrolyzed Bacterial Cellulose: A Versatile Support for Lithium Ion Battery Anode Materials. Small 2013, 9 (14), 2399-2404.

29 ACS Paragon Plus Environment

Environmental Science & Technology

540 541 542

42. Liang, H.-W.; Guan, Q.-F.; Zhu, Z.; Song, L.-T.; Yao, H.-B.; Lei, X.; Yu, S.-H. Highly conductive and stretchable conductors fabricated from bacterial cellulose. NPG Asia Mater. 2012, 4, e19.

543 544 545

43. Olsson, R. T.; Azizi Samir, M. A. S.; Salazar Alvarez, G.; BelovaL; StromV; Berglund, L. A.; IkkalaO; NoguesJ; Gedde, U. W. Making flexible magnetic aerogels and stiff magnetic nanopaper using cellulose nanofibrils as templates. Nat. Nanotechnol. 2010, 5 (8), 584-588.

546 547 548

44. Jiang, Q.; Tian, L.; Liu, K.-K.; Tadepalli, S.; Raliya, R.; Biswas, P.; Naik, R. R.; Singamaneni, S. Bilayered Biofoam for Highly Efficient Solar Steam Generation. Adv. Mater. 2016, 28 (42), 9400-9407.

549 550 551

45. Marcano, D. C.; Kosynkin, D. V.; Berlin, J. M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L. B.; Lu, W.; Tour, J. M. Improved synthesis of graphene oxide. ACS nano 2010, 4 (8), 4806-4814.

552 553

46. Stumm, W.; Morgan, J. J. Aquatic Chemistry: Chemical Equilibria and Rates in Natrual Water, 3rd, ed.; John Wiley & Sons: New York, 1996.

554 555

47. Kansal, S.; Singh, M.; Sud, D., Studies on photodegradation of two commercial dyes in aqueous phase using different photocatalysts. J. Hazard. Mater. 2007, 141 (3), 581-590.

556 557 558

48. Yan, Z.; Xia, M.; Wang, P.; Zhang, P.; Liang, O.; Xie, Y.-H. Selective manipulation of molecules by electrostatic force and detection of single molecules in aqueous solution. J. Phys. Chem. C 2016, 120 (23), 12765-12772.

559 560 561

49. Zheng, Y.; Zhong, X.; Li, Z.; Xia, Y. Successive, Seed-Mediated Growth for the Synthesis of Single-Crystal Gold Nanospheres with Uniform Diameters Controlled in the Range of 5–150 nm. Part. Part. Syst. Charact. 2014, 31 (2), 266-273.

562 563 564

50. Gottenbos, B.; van der Mei, H. C.; Klatter, F.; Nieuwenhuis, P.; Busscher, H. J. In vitro and in vivo antimicrobial activity of covalently coupled quaternary ammonium silane coatings on silicone rubber. Biomaterials 2002, 23 (6), 1417-1423.

565 566

51. Kulkarni, D. D.; Choi, I.; Singamaneni, S. S.; Tsukruk, V. V. Graphene Oxide−Polyelectrolyte Nanomembranes. ACS Nano 2010, 4 (8), 4667-4676.

567 568 569

52. Kulkarni, D. D.; Kim, S.; Chyasnavichyus, M.; Hu, K.; Fedorov, A. G.; Tsukruk, V. V., Chemical Reduction of Individual Graphene Oxide Sheets as Revealed by Electrostatic Force Microscopy. J. Am. Chem. Soc. 2014, 136 (18), 6546-6549.

570 571

53. Gatenholm, P.; Klemm, D. Bacterial Nanocellulose as a Renewable Material for Biomedical Applications. MRS Bull. 2010, 35 (03), 208-213.

572 573 574

54. Rourke, J. P.; Pandey, P. A.; Moore, J. J.; Bates, M.; Kinloch, I. A.; Young, R. J.; Wilson, N. R. The Real Graphene Oxide Revealed: Stripping the Oxidative Debris from the Graphenelike Sheets. Angew. Chem. Int. Ed. 2011, 50 (14), 3173-3177. 30 ACS Paragon Plus Environment

Page 32 of 33

Page 33 of 33

Environmental Science & Technology

575 576 577

55. Poletto, M.; Pistor, V.; Santana, R. M. C.; Zattera, A. J., Materials produced from plant biomass: part II: evaluation of crystallinity and degradation kinetics of cellulose. Mater. Res. 2012, 15, 421-427.

578 579 580

56. Kidambi, P. R.; Jang, D.; Idrobo, J.-C.; Boutilier, M. S. H.; Wang, L.; Kong, J.; Karnik, R. Nanoporous Atomically Thin Graphene Membranes for Desalting and Dialysis Applications. Adv. Mater. 2017, 29 (33), 1700277.

581 582

57. Wang, L. K.; Chen, J. P.; Hung, Y.-T.; Shammas, N. K., Eds. Membrane and desalination technologies; Humana Press: New Jersey 2008.

583 584

58. Baker, R. W. Membrane Technology and Applications; John Wiley & Sons, Ltd: Califonia, USA, 2004.

585 586 587

59. Fang, Q.; Zhou, X.; Deng, W.; Zheng, Z.; Liu, Z. Freestanding bacterial cellulosegraphene oxide composite membranes with high mechanical strength for selective ion permeation. Sci. Rep. 2016, 6, 33185.

588 589 590

60. Joshi, R. K.; Carbone, P.; Wang, F. C.; Kravets, V. G.; Su, Y.; Grigorieva, I. V.; Wu, H. A.; Geim, A. K.; Nair, R. R. Precise and Ultrafast Molecular Sieving Through Graphene Oxide Membranes. Science 2014, 343 (6172), 752-754.

591 592 593 594

61. Chen, L.; Shi, G.; Shen, J.; Peng, B.; Zhang, B.; Wang, Y.; Bian, F.; Wang, J.; Li, D.; Qian, Z.; Xu, G.; Liu, G.; Zeng, J.; Zhang, L.; Yang, Y.; Zhou, G.; Wu, M.; Jin, W.; Li, J.; Fang, H. Ion sieving in graphene oxide membranes via cationic control of interlayer spacing. Nature 2017, 550 (7676), 380.

595 596 597

62. Abraham, J.; Vasu, K. S.; Williams, C. D.; Gopinadhan, K.; Su, Y.; Cherian, C. T.; Dix, J.; Prestat, E.; Haigh, S. J.; Grigorieva, I. V.; Carbone, P.; Geim, A. K.; Nair, R. R. Tunable sieving of ions using graphene oxide membranes. Nat. Nanotechnol. 2017, 12 (6), 546.

598 599

63. An, D.; Yang, L.; Wang, T.-J.; Liu, B. Separation Performance of Graphene Oxide Membrane in Aqueous Solution. Ind. Eng. Chem. Res. 2016, 55 (17), 4803-4810.

600 601 602

64. Lim, D.-K.; Barhoumi, A.; Wylie, R. G.; Reznor, G.; Langer, R. S.; Kohane, D. S. Enhanced Photothermal Effect of Plasmonic Nanoparticles Coated with Reduced Graphene Oxide. Nano Lett. 2013, 13 (9), 4075-4079.

603 604 605

65. Some, S.; Xu, Y.; Kim, Y.; Yoon, Y.; Qin, H.; Kulkarni, A.; Kim, T.; Lee, H. Highly sensitive and selective gas sensor using hydrophilic and hydrophobic graphenes. Sci. Rep. 2013, 3, srep01868.

606

31 ACS Paragon Plus Environment