In Situ Reduction of Silver by Polydopamine: A Novel Antimicrobial

Aug 1, 2016 - Singapore Centre for Environmental Life Sciences Engineering, Nanyang ... permeability, which provides a new dimension for membrane surf...
1 downloads 3 Views 2MB Size
Subscriber access provided by Northern Illinois University

Article

In situ reduction of silver by polydopamine: a novel anti-microbial modification of thin-film composite polyamide membrane Zhe Yang, Yichao Wu, Jianqiang Wang, Bin Cao, and Chuyang Y. Tang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b01867 • Publication Date (Web): 01 Aug 2016 Downloaded from http://pubs.acs.org on August 8, 2016

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 free 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 accessible to all readers and 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.

Environmental Science & Technology 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 29

Environmental Science & Technology

1

In situ reduction of silver by polydopamine: a novel anti-microbial modification

2

of thin-film composite polyamide membrane

3 Zhe Yanga, Yichao Wub,c, Jianqiang Wanga, Bin Caob,c, Chuyang Y. Tanga,*

4 5 6

a

Department of Civil Engineering, the University of Hong Kong, Pokfulam, Hong Kong

7

b

School of Civil and Environmental Engineering, Nanyang Technological University, 50 Nanyang Avenue,

8

Singapore 639798

9

c

Singapore Centre for Environmental Life Sciences Engineering, Nanyang Technological University, 60 Nanyang

10

Avenue, Singapore 637551

11

*

Corresponding Author. Tel: +852 2859 1976, Fax: +852 2559 5337, E-mail address: [email protected]

12 13

ACS Paragon Plus Environment

Environmental Science & Technology

14

Abstract

15 16 17 18 19 20 21 22 23 24 25 26 27 28 29

We report a facile method for the anti-microbial modification of a thin-film composite polyamide reverse osmosis (RO) membrane. The membrane surface was first coated with polydopamine (PDA), whose reducing catechol groups subsequently immobilized silver ions in situ to form uniformly-dispersed silver nanoparticles (AgNPs) inside the coating layer. Agglomeration of AgNPs was not observed despite a high silver loading of 13.3 ± 0.3 µg/cm2 (corresponding to a surface coverage of 18.5% by the nanoparticles). Both diffusion inhibition zone tests and colony formation unit tests showed clear anti-microbial effects of the silver loaded membranes on model bacteria Bacillus subtills and Escherichia coli. Furthermore, the silver immobilized membrane had significantly enhanced salt rejection compared to the control PDA coated membrane, which is attributed to the preferential formation of AgNPs at defect sides within the PDA layer. This self-healing mechanism can be used to prepare anti-microbial RO membranes with improved salt rejection without scarifying the membrane permeability, which provides a new dimension for membrane surface modification.

30 31

TOC art

32 33

ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29

Environmental Science & Technology

34



35

Water scarcity is a grand challenge to human beings.1 This challenge can be

36

potentially addressed by desalination using reverse osmosis (RO) technology.2 A key

37

obstacle of the RO technology is membrane biofouling, i.e., the development of a

38

cohesive biofilm on membrane surfaces. Severe biofouling can lead to significant loss

39

of water permeability and product water quality.

40

issues, physicochemical pretreatment methods and disinfection (e.g., chlorine or UV)

41

of feed water are applied.4 Although membrane biofouling can be mitigated by

42

chlorinating the feed water, some microorganisms can survive through the

43

chlorination treatment and eventually colonize on membrane surfaces.5 In addition,

44

chlorine-based disinfectants can cause chemical damage to RO membranes.6, 7 For

45

example, it has been well documented that thin-film composite (TFC) polyamide (PA)

46

RO membranes8 have a low tolerance to free chlorine (< 0.1 ppm).9 Therefore,

47

alternative disinfection methods are needed to avoid membrane damage.

INTRODUCTION

3

To mitigate membrane biofouling

48 49

In recent years, researchers have developed a variety of anti-adhesion10, 11 and anti-

50

microbial12-15 surface modification methods. Among the various anti-microbial agents

51

used, silver nanoparticles (AgNPs) have attracted growing interest due to their strong

52

disinfection power.16 AgNPs are commonly blended with polymers or monomers

53

during membrane formation.17, 18 An inherent disadvantage of this approach is the

54

agglomeration of AgNPs, which can lead to the formation of defects and even the loss

ACS Paragon Plus Environment

Environmental Science & Technology

55

of membrane rejection at high loading of particles.19 Yin et al.12 reported the grafting

56

of AgNPs onto a TFC membrane surface by using a bridging chemical agent

57

(cysteamine) to form covalent bonding with AgNPs. Nevertheless, the chemical

58

modification led to a reduced membrane rejection, which is attributed to the swelling

59

of the rejection layer. Ben-Sasson et al.15 prepared a uniform coverage of AgNPs on a

60

TFC RO membrane by reducing silver ions from a solution phase. Although this

61

approach avoids the agglomeration of AgNPs, a strong reducing agent (sodium

62

borohydride) has to be used. Greener and more environmentally benign methods for

63

AgNPs incorporation are yet to be developed.

64 65

We are inspired by the recent literature on using polydopamine (PDA) for the

66

preparation of antifouling membranes.20, 21 PDA is a polymer derived from dopamine,

67

a “bio-glue” that has been isolated from mussels and that is responsible for their

68

attachment to different surfaces in natural environment (e.g., rock).20 The sticky

69

nature of dopamine and its ability to form polymers enable researchers to prepare

70

PDA coating on a variety of solid substrates for applications in environmental, energy,

71

and biomedical fields.22, 23 Furthermore, PDA is highly hydrophilic due to its catechol,

72

quinone, and amine functional groups24, which makes PDA an ideal surface

73

modification agent for preparing anti-adhesion membranes.25, 26 In a recent study on

74

synthesizing metal/organic hybrid nanomaterials, Hong et al.21 presented a simple and

75

elegant method to fix silver in situ by exposing PDA to an AgNO3 solution. Silver

ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29

Environmental Science & Technology

76

ions were reduced spontaneously by the catechol groups of PDA to form well-

77

dispersed AgNPs on a polymer nanofiber substrate. A similar approach has been

78

reported by Tang et al. for biofouling control for a porous ultrafiltration (UF)

79

membrane.27 These authors demonstrated strong anti-adhesive and anti-biofouling

80

properties of PDA/Ag coated UF membranes. Up to date, such PDA/Ag coating has

81

not been applied to membranes with dense rejection layers (such as RO). Compare to

82

porous membranes, a PDA/Ag coating on RO will likely exhibit very different impact

83

on the membrane’s separation properties.

84 85

These existing studies prompt us to use PDA as a bioinspired scaffold to prepare anti-

86

microbial RO membranes. Specifically, PDA-AgNPs composite coatings were

87

prepared on a commercial RO membrane, and their effects on the separation

88

performance and anti-microbial behavior are systematically.

89 90



91

Materials and chemicals.

92

A commercial RO membrane (XLE), received from Dow FilmTec, was used as the

93

base membrane for further modification. XLE is a brackish water TFC RO

94

membrane.28 Its dense polyamide rejection layer is supported by a porous polysulfone

95

(PSF) layer, and a polyester (PET) nonwoven fabric is used to provide the required

96

mechanical strength.28

MATERIALS AND METHODS

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 29

97 98

Unless specified otherwise, all chemicals were of ACS reagent grades. All solutions

99

were prepared using deionized water supplied from a Milli-Q system (Millipore,

100

Billerica,

MA).

Dopamine-hydrochloride

powder,

Tris

(hydroxymethyl)

101

aminomethane (Tris-HCl, ≥ 99.0 %), and hydrochloric acid (HCl) were obtained from

102

Sigma Aldrich and were used for the preparation of PDA coatings. Silver nitrate

103

(AgNO3, Sigma Aldrich) was used as the silver source for the in situ formation of

104

AgNPs. Sodium chloride (NaCl, Sigma Aldrich) was used in membrane rejection

105

tests. N,N-Dimethylformamide (DMF, ReagentPlus®, ≥99%, Sigma Aldrich) was

106

used as a solvent for dissolving the PSF support to prepare isolated polyamide films

107

used for transmission electron microscopy (TEM) characterization.

108 109

PDA coating and in situ formation of AgNPs on TFC RO membranes.

110

The main surface modifications in the current study include a PDA coating step

111

followed by in situ formation of AgNPs (Figure 1). Prior to PDA coating, the base

112

membrane XLE was thoroughly rinsed with deionized water and was dried in air. A

113

membrane coupon of 20 × 12 cm was placed in a custom-made container in such a

114

way that only its rejection layer was exposed to the coating solution. The coating

115

solution was prepared by adding 0.4 g dopamine hydrochloride into a 200 ml buffer

116

solution of 10 mM Tris-HCl. The solution pH was adjusted to 8.5 to allow optimized

117

self-polymerization of PDA.26, 29 During the entire coating step, the coating solution

ACS Paragon Plus Environment

Page 7 of 29

Environmental Science & Technology

118

was continuously stirred to minimize the aggregation of PDA. The PDA-coated

119

membrane is denoted as PDAn, where n is the duration of coating (n = 0.5, 1, and 2 h

120

in the current study).

121 122 123 124 125

Fig. 1. Schematic diagram of in situ formation of AgNPs on a thin film composite membrane. The base membrane XLE was first coated with polydopamine, followed by in situ reduction of silver by immersing the PDA coated membrane into an AgNO3 solution.

126 127

The freshly coated membrane was rinsed with deionized water for 30 minutes before

128

it was placed in a 200 ml solution of 4 gL-1 AgNO3 for the immobilization of silver.

129

Briefly, the PDA-coated rejection layer was exposed to the silver solution in the same

130

custom-made container that was used for PDA coating. The container was wrapped

131

with an aluminum foil to avoid sunlight and was placed on a shaking bath under room

132

temperature for 5 h (adopted based on existing literature30). The resulting silver

133

incorporated membrane is named as PDAnAg.

134 135

Membrane characterization.

136

X-ray photoelectron spectroscopy (XPS) was performed using an SKL-12

ACS Paragon Plus Environment

Environmental Science & Technology

137

spectrometer (Leybold, Sengyang, China) modified with a VG CLAM 4 MCD

138

electron energy analyzer. An Al Kα gun (1496.3 eV) operated at 10 kV and 15 mA

139

was used as the x-ray source. Membrane samples were thoroughly rinsed several

140

times and dried before XPS characterization. Survey spectra over 0-1000 eV were

141

obtained at a scanning resolution of 0.1 eV.

142 143

Membrane structure and morphology were characterized by scanning electron

144

microscopy (SEM) and TEM. To isolate the membrane rejection layer for TEM

145

characterization, the polyester fabric layer was first removed and the PSF layer was

146

subsequently dissolved using DMF as the solvent. The use of DMF for isolating

147

rejection layers of RO and NF membranes has been reported in the existing

148

literature.31 The isolated rejection layer was float on the DMF solvent and was picked

149

up by a carbon-coated copper TEM grid. The sample was allowed to dry in air at

150

room temperature. TEM sample characterization was performed with Philips CM100

151

TEM (Philips, Eindhoven, Netherlands) operating at 100 kV.

152 153

SEM characterization was conducted using a Field Emission Gun Scanning Electron

154

Microscope (LEO1530 FEG SEM, UK) equipped with an energy dispersive

155

spectroscopy (EDS) detector. Vacuum dried membrane samples of approximately 0.5

156

cm × 0.5 cm were sputter-coated with a uniform layer of gold and platinum (SCD

157

005, BAL-TEC, NYC) to avoid sample charging. SEM micrographs were obtained at

ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29

Environmental Science & Technology

158

an accelerating voltage of 5 kV. EDS was also performed to determine the elemental

159

composition of the membrane sample. For EDS analysis, a voltage of 20 kV was

160

used.

161 162

Water contact angle measurements were performed using a goniometer equipped with

163

a video capture device (Powereach®, China). Prior to each test, a membrane sample

164

was dried in vacuum at room temperature for 24 h. Each deionized water droplet with

165

a volume of approximately 5 µL was introduced to the membrane surface, and a

166

stabilizing time of 10 seconds was allowed. For each membrane coupon, contact angle

167

was measured at five different locations and the average value was reported.

168 169

Image analysis.

170

An image analysis software (MediaCybernetics, Inc.) was utilized to analyze the size

171

distribution of AgNPs and their surface coverage over the silver incorporated

172

membranes. Prior to the analysis, TEM micrographs were first saved as 8-bit

173

greyscale images in Tagged-Image File Format (TIFF). These greyscale images were

174

then converted into binary black-and-white images. Background noise was alleviated

175

by applying threshold value adjustment following a previous study.32 In addition to

176

the particle size determination, the percentage of membrane area covered by AgNPs

177

in the 2D images were determined and reported as the surface coverage. The analysis

178

was performed based on three replicate measurements.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 29

179 180

Separation performance tests.

181

A high pressure cross-flow filtration system, similar to the one reported by Tang et

182

al.33, was used to evaluate water flux and solute rejection of membranes under a

183

constant pressure mode. The temperature was maintained constant at around 25 °C

184

using an immersion thermostat (J.P. Selecta S.A., Barcelona, Spain). For each test, a

185

membrane coupon with an effective area of 42 cm2 was placed in a cross-flow cell

186

(CF042, Delrine acetal, Sterlitech). The coupon was pre-compacted using deionized

187

water at the set pressure of 2 MPa for 24 h in order to achieve a stable water flux. The

188

pure water flux was then determined by measuring the mass of the permeate water

189

collected over a specified time interval according to Equation (1):

190

Jv 

191

where Jv (Lm-2h-1) is the water flux, w (kg) is the mass of permeate water collected

192

over a time period of t (h), A (m2) is the effective membrane area, and  (kgL-1) is

193

the density of permeate water.

w t  A  

(1)

194 195

Salt rejection was measured using a 2000 ppm NaCl solution as the feed water. An

196

Ultrameter II (Myron L company, Carlsbad, CA) was used to determine the

197

conductivity of the feed water (Cf) and that of the permeate (Cp), respectively.

198

Membrane rejection R was calculated by:

ACS Paragon Plus Environment

Page 11 of 29

199

Environmental Science & Technology

R  (1 

Cp Cf

) 100%

(2)

200 201

Antibacterial assessment.

202

All membranes were immersed in deionized water for 24 h before the assessment of

203

their anti-microbial activity. A Gram-positive Bacillus subtilis 168 (ATCC 27370) and

204

a Gram-negative Escherichia coli K12 (ATCC 10798) were as the model bacteria

205

Membrane coupons were immersed in the cell suspension that was incubated on a

206

rotary shaker (150 rpm) at room temperature (10 h for B. subtilis and 24 h for E. coli).

207

Viable cells in the suspension were quantified using the colony forming unit (CFU)

208

method (CLSI M07-A935). Experimental results were obtained from three

209

independent replicates.

34

.

210 211

Diffusion inhibition zone (DIZ) tests were also performed to assess the bactericidal

212

effects of the membranes following a modified CLSI M07-A9 method.35 Briefly,

213

aliquots (100 µL) of bacterial culture were spread onto an LB agar plate. Membrane

214

disk samples (diameter = 12.7 mm) were then placed onto the plate with their active

215

layers in contact with the agar surface. After incubation at optimal temperature (37 ºC

216

for E. coli and 30 ºC for B. subtilis) for 24 h, the colonies formed under the membrane

217

samples were examined.

218

ACS Paragon Plus Environment

Environmental Science & Technology

219

Silver loading quantification and silver leaching tests.

220

To measure the total amounts of silver immobilized on the membrane, AgNPs

221

incorporated membrane coupons (3.8 cm2) were placed in glass vials containing 0.2

222

ml 70% HNO3 in 20 ml deionized water. The vials were placed on a reciprocal

223

shaking bath (Unitronic Reciprocal 6032011, J.P. Selecta, S.A., Barcelona, Spain)

224

under 200 rpm for one day. Subsequently, the dissolved silver concentration was

225

quantified using an inductive coupled plasma optical emission spectrometer (ICP-

226

OES, Optima 8x00, Perkin Elmer).

227 228

Ag leaching tests were performed to evaluate the stability of the AgNPs. Membrane

229

coupons (3.8 cm2) were immersed in 20 ml deionized water agitated at 200 rpm. The

230

soaking solution was replaced daily, and the collected water samples were filtered

231

through a 0.45 m membrane and acidified by 1% HNO3 before ICP-OES analysis.

232

ACS Paragon Plus Environment

Page 12 of 29

Page 13 of 29

Environmental Science & Technology

233



234

Membrane characterization.

235

Figure 2 presents the SEM micrographs of the control membrane XLE, the coated

236

membranes PDAn, and the silver-immersed membranes PDAnAg. The control

237

membrane XLE had a ‘ridge and valley’ roughness structure. PDA coating did not

238

have obvious effect on the surface morphology, except for the relatively long coating

239

duration of 2 h at which some aggregates of PDA were apparent. Similar phenomena

240

of PDA aggregation have been observed at relative long coating duration in the

241

existing literature.36-39

RESULTS AND DISCUSSION

242

243 244 245 246 247 248

Fig. 2. SEM micrographs (plan view) of control membrane XLE, the PDA coated membranes (PDA0.5, PDA1 and PDA2), and the silver incorporated membranes (PDA0.5Ag, PDA1Ag, and PDA2Ag). For the silver incorporated membranes, the corresponding EDS spectrum is shown in the insert of each micrograph. The scale bar of all micrographs is 1 µm.

249

ACS Paragon Plus Environment

Environmental Science & Technology

250

After the PDA-coated membranes were immersed in the AgNO3 solution, additional

251

fine particles appeared on the membrane surfaces (Figure 2). Elemental analysis by

252

EDS (inserts of Figure 2) and XPS (Supporting Information S1) suggests that these

253

nano-sized particles were AgNPs. The spontaneous immobilization of silver can be

254

attributed to the reducing catechol groups contained in PDA.21 According to Yang et

255

al.24, electrons released by the oxidation of catechol to quinone can reduce silver ions

256

in the solution phase. At the same time, the O- and N-based ligand sites in PDA could

257

serve as anchors for the resulting AgNPs.40, 41

258 259

Figure 3a-d presents the TEM micrographs (plan view) of the control membrane XLE

260

and the silver incorporated membranes. Consistent with the SEM micrographs (Figure

261

2), the TEM micrographs also show the ‘ridge and valley’ structure of the polyamide

262

rejection layer. For the silver incorporated membranes, spherical AgNPs with

263

relatively uniform size can be clearly observed. These particles were homogeneously

264

distributed over the membrane surface. In addition, increasing the coating time of

265

PDA led to increased loading of AgNPs, which can be attributed to the enhanced PDA

266

deposition and thus the increased concentration of catechol groups. TEM cross-

267

sections of XLE, PDA2, and PDA2Ag are presented in Figure 3e-g (see additional

268

cross sections in Supporting Information SI3). The PDA coating thickness was on the

269

order of 20-30 nm after 2-h coating, which agrees with the literature20 that the coating

270

layer grows at an initial rate of approximately 10 nm/h. We were not able to observe

ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29

271

Environmental Science & Technology

major changes in PDA thickness or conformation on the basis of TEM cross-sections.

272 (e)

(f)

(g)

273 274 275 276 277

Fig. 3. TEM micrographs. Plan view (a-d): control membrane XLE (a) and the silver incorporated membranes PDA0.5Ag (b), PDA1Ag (c), and PDA2Ag (d). Crosssectional view (e-g): control membrane XLE (e), PDA-coated membrane PDA2 (f), and silver incorporated membrane PDA2Ag (g). The scale bar of all micrographs is 200 nm.

ACS Paragon Plus Environment

Environmental Science & Technology

278

The TEM micrographs were further analyzed to determine the size and quantity of the

279

AgNPs immobilized on the membrane surface. As shown in Figure 4a, these in situ

280

reduced AgNPs had a relatively narrow size distribution, with particle size generally

281

on the order of 10-20 nm. The particle number density, i.e., the number of particles

282

found within a unit area, increased from 6.8 × 107 to 12.9 × 107 #/cm2 as the PDA

283

coating time increased from 0.5 to 2 h. The increase in particle number density led to

284

a significantly increased surface coverage by AgNPs (from 7.9 ± 1.9 % for PDA0.5Ag

285

to 18.5 ± 0.6 % for PDA2Ag, see Figure 4b). Such high surface coverage by AgNPs

286

has been seldom reported for conventional silver loading methods including simple

287

blending.17

288 289 290 291

Fig. 4. Characterization of AgNPs. The size distribution of AgNPs (a) was obtained by image analysis of the corresponding TEM micrograph (shown in the insert). The particle number density and surface coverage of AgNPs (b) were also based on image

ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29

292 293 294

Environmental Science & Technology

analysis of TEM micrographs. The elemental percentage of silver and the silver mass loading (c) were obtained by EDS analysis and silver dissolution experiments, respectively.

295 296

Further characterization of silver loading was performed using EDS analysis and

297

silver dissolution tests (Figure 4c). Consistent with TEM observations, the silver

298

dissolution tests revealed a greatly increased mass loading of silver (from 4.5 ± 0.2

299

µg/cm2 for PDA0.5Ag to 13.3 ± 0.3 µg/cm2 for PDA2Ag). A similar trend was also

300

identified on the basis of the EDS elemental analysis.

301 302

Membrane hydrophilicity and separation performance.

303

The membrane contact angle was strongly affected by the PDA coating and silver

304

loading (Figure 5a). A 2-h PDA coating effectively reduced the contact angle from

305

45.1 ± 2.2º (XLE) to 24.7 ± 5.4º (PDA2), which can be attributed to the hydrophilic

306

catechol, quinone, and amine groups in PDA.42,

307

AgNPs further reduced the contact angle (13.5± 2.4º for PDA2Ag). The effect of

308

AgNPs on improving membrane hydrophilicity has been reported in the literature.18,

309

44, 45

43

The in situ immobilization of

According to a recent review on nanocomposite membranes,19 the inclusion of

310

AgNPs can significantly improve the wettability of membrane surface as a result of

311

their hydrophilic nature.

312

ACS Paragon Plus Environment

Environmental Science & Technology

313 314 315 316 317 318

Fig. 5. Effect of PDA coating and silver loading on membrane properties. (a) Water contact angle, (b) water flux and salt rejection. Membrane filtration was performed at 20 MPa. Water flux was obtained using deionized feed water, and rejection tests were performed for feed solutions containing 2000 ppm NaCl, (c) The resulting membrane contains uniformly dispersed silver nanoparticles.

319

No major effect by the PDA coating was observed with short duration ( 1 h) within

320

the experimental uncertainty (Figure 5b). However, increasing coating time to 2 h

321

resulted in a flux reduction of nearly 20%, which is likely due to the formation of a

322

relatively thick PDA layer. The excessive formation of PDA aggregates (Figure 2)

323

may also partially explain the reduced membrane flux. The incorporation of AgNPs

324

into the PDA layer decreased water flux from 54.3 ± 1.4 L/m2h (PDA2) to 32.1 ± 2.8

325

L/m2h (PDA2Ag). In comparison, the effect of silver loading on water flux was much

326

milder at lower PDA coating time. The trend observed in the current study may be

ACS Paragon Plus Environment

Page 18 of 29

Page 19 of 29

Environmental Science & Technology

327

explained by the competing effects of (1) the increased loading of impermeable

328

AgNPs (Figure 4b) that blocks water pathways and (2) the improved membrane

329

hydrophilicity (Figure 5a) that promotes water passage.

330 331

The PDA coating reduced the NaCl rejection, with greater loss observed at longer

332

coating time (Figure 5b). Although charge interaction is a plausible cause for the

333

reduced rejection, there were no major differences in the measured zeta potential of

334

the control and PDA-coated membranes (Supporting Information S4). The current

335

result suggests that the PDA layer was more permeable to NaCl compared to the base

336

membrane, which is consistent to the existing literature reports on the use of PDA as

337

nanofiltration (NF) rejection layers.46 Indeed, the combination of a relatively salt-

338

permeable PDA layer and a high rejection polyamide layer can lead to the

339

accumulation of rejected NaCl within the PDA layer and thus the loss of salt rejection,

340

a phenomenon known as cake enhanced concentration polarization.47 Regardless of

341

the PDA coating time, the AgNPs incorporated membranes consistently showed

342

improved rejection compared to their respective counterpart without silver loading.

343

Such enhancement of rejection implies a reduction of salt flux in addition to the loss

344

of water flux, which is likely due to the simultaneous blockage of salt flux and water

345

pathways (i.e., defects in the PDA layer) by the impermeable AgNPs. The effect of

346

defect sealing upon in situ immobilization of silver can be explained by a self-healing

347

mechanism (Figure 5c): a defect region that is more permeable to NaCl is also more

ACS Paragon Plus Environment

Environmental Science & Technology

348

accessible to silver ions, which leads to enhanced localized AgNPs formation to heal

349

the defect.

350 351

To facilitate the further understanding of the rejection behavior of the coated

352

membrane, a simple resistance-in-series model is developed in Supporting

353

Information S5. According to the model, whether the coating improves or decreases

354

the rejection of the coated membrane depends on the A/B value of the coating in

355

comparison to the A/B value of the base membrane, where A is the water permeability

356

and B is the salt permeability. Rejection improves if (A/B)coating > (A/B)base. Sealing the

357

defects in the coating layer can greatly increase (A/B)coating due to greatly decreased B

358

value of the coating, thus leading to improved overall rejection. On the other hand,

359

excessive silver loading may lead to severe loss of water permeability (low Acoating),

360

leading to a reduced (A/B)coating and thus a reduction in rejection compared to the base

361

membrane (e.g., PDA2Ag vs. XLE).

362 363

Antibacterial tests.

364

Diffusion inhibition zone tests clearly demonstrate the antibacterial effect of the

365

AgNPs incorporated membranes (Figure 6). For both B. subtilis and E. coli, colonies

366

developed under the silver-free control membranes XLE and PDA2. In contrast, no

367

apparent bacterial growth under the membrane was observed for all the silver

368

incorporated membranes (PDA0.5Ag, PDA1Ag, and PDA2Ag). CFU tests also

ACS Paragon Plus Environment

Page 20 of 29

Page 21 of 29

Environmental Science & Technology

369

showed compelling antibacterial effect of the silver loaded membranes (Figure 7).

370

The exposure to PDA2Ag led to a reduction of 62.7 ± 9.3% for viable B. subtilis and

371

42.4 ± 5.7% for E. coli.

372

373 374 375 376 377

Fig. 6. Diffusion inhibition zone test for Gram-positive B. subtilis and Gram-negative E. coli. Membrane disks (diameter = 12.7 mm, rejection layer facing downwards) were placed onto agar plate spread with bacterial culture. A 24 h incubation time was used in the tests.

378

ACS Paragon Plus Environment

Environmental Science & Technology

379 380 381 382 383 384

Fig. 7. Colony forming units (CFU) tests for Gram-positive B. subtilis and Gramnegative E. coli. Membrane disks (diameter = 12.7 mm) were placed into cell suspension (cell density of approximately 3000 cells/mL for B. subtilis and 2.0 × 108 cells/mL for E. coli). After culturing at room temperature (10 h with B. subtilis or 24 h with E. coli), the cultivated cells in suspension were quantified using CFU method. The reported data is the average value of three replicates.

385 386

ACS Paragon Plus Environment

Page 22 of 29

Page 23 of 29

Environmental Science & Technology

387

Implications.

388

We developed a green and facile method to prepare anti-microbial RO membrane.

389

Silver ions from an AgNO3 solution was spontaneously immobilized by the reducing

390

catechol groups in a PDA coating layer without the need to use additional strong

391

reductants and aggressive chemicals. The ability of PDA to coat onto a wide variety

392

of substrates20 implies that this method can be potentially used to modify membranes

393

with different surface chemistries and morphological features. Future studies may

394

further explore its use for modifying other membrane types (e.g., nanofiltration and

395

pervaporation membranes) as well as different membrane base materials (e.g.,

396

inorganic membranes in addition to polymeric membranes). The conditions for PDA

397

coating and silver incorporation may be further optimized to provide a fine control on

398

the silver loading (e.g., the size and number density of AgNPs). The recharge of

399

AgNPs upon their depletion can be a potential challenge. Future studies shall address

400

this problem, for example, by the use of catechol-containing polymers that can

401

sacrificed and re-coated to allow silver recharge. In addition, long-term stability of

402

separation performance needs to be assessed.

403 404

We further report a self-healing mechanism that involves the preferential formation of

405

AgNPs at the defect sites of a PDA coating layer due to their greater accessibility to

406

silver ions. The sealing of the salt pathway is responsible for the enhanced membrane

407

salt rejection. At the meantime, the water permeability of the PDA-AgNPs modified

ACS Paragon Plus Environment

Environmental Science & Technology

408

membranes was not significantly affected at short PDA coating time (e.g., 0.5 h). The

409

current study demonstrated the feasibility of using surface modification to achieve

410

improved salt rejection without sacrificing the water permeability. This method may

411

be further extended to direct defects sealing of RO and NF membranes where

412

damages can be repaired locally to maintain the membrane integrity,48 and a wide

413

range of other metals or metal oxides based nanoparticles can be potentially utilized

414

for such purpose.

415 416



417

Support Information. S1. X-ray photoelectron spectroscopy; S2. Silver leaching

418

tests; S3. TEM cross sectional images; S4. Zeta potential measurements; S5.

419

Resistance-in-series model. This material is available free of charge via the Internet at

420

http://pubs.acs.org.

ASSOCIATED CONTENT

421 422



423

Corresponding Author

424

*Phone: (+852) 2859 1976; e-mail: [email protected]

425

Notes

426

The authors declare no completing financial interest.

AUTHOR INFORMATION

427

ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29

Environmental Science & Technology

428



429

The study receives financial support from the Innovation and Technology

430

Commission of the Hong Kong Government (Project # ITS/208/14) and the General

431

Research Fund of the Research Grants Council (Project # 17207514). The partial

432

funding support from Strategic Research Theme on Clean Energy at the University of

433

Hong Kong is also appreciated. We acknowledge the help on ICP-OES measurements

434

by Ms. Hanlu Yan and Dr. Kaimin Shih at the Department of Civil Engineering, the

435

University of Hong Kong. The XPS used in this work was supported by the Institute

436

for Advanced Materials (IAM) with funding support by the Special Equipment Grant

437

from the University Grants Committee of the Hong Kong Special Administrative

438

Region, China (SEG_HKBU06).

ACKNOWLEDGMENTS

439

ACS Paragon Plus Environment

Environmental Science & Technology

440

Reference

441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480

1. Seckler, D.; Barker, R.; Amarasinghe, U., Water scarcity in the twenty-first century. Int. J. Water Res. Dev. 1999, 15, (1-2), 29-42. 2. Elimelech, M.; Phillip, W. A., The future of seawater desalination: energy, technology, and the environment. Science 2011, 333, (6043), 712-717. 3. Matin, A.; Khan, Z.; Zaidi, S.; Boyce, M., Biofouling in reverse osmosis membranes for seawater desalination: phenomena and prevention. Desalination 2011, 281, 1-16. 4. Prihasto, N.; Liu, Q.-F.; Kim, S.-H., Pre-treatment strategies for seawater desalination by reverse osmosis system. Desalination 2009, 249, (1), 308-316. 5. Kochkodan, V.; Hilal, N., A comprehensive review on surface modified polymer membranes for biofouling mitigation. Desalination 2015, 356, 187-207. 6. Do, V. T.; Tang, C. Y.; Reinhard, M.; Leckie, J. O., Effects of chlorine exposure conditions on physiochemical properties and performance of a polyamide membrane--mechanisms and implications. Environ. Sci. Technol. 2012, 46, (24), 13184-92. 7. Raval, H.; Trivedi, J.; Joshi, S.; Devmurari, C., Flux enhancement of thin film composite RO membrane by controlled chlorine treatment. Desalination 2010, 250, (3), 945-949. 8. Tang, C. Y.; Kwon, Y.-N.; Leckie, J. O., Probing the nano-and micro-scales of reverse osmosis membranes—a comprehensive characterization of physiochemical properties of uncoated and coated membranes by XPS, TEM, ATR-FTIR, and streaming potential measurements. J. Membr. Sci. 2007, 287, (1), 146-156. 9. Glater, J.; Hong, S.-k.; Elimelech, M., The search for a chlorine-resistant reverse osmosis membrane. Desalination 1994, 95, (3), 325-345. 10. Dong, L.-x.; Yang, H.-w.; Liu, S.-t.; Wang, X.-m.; Xie, Y. F., Fabrication and anti-biofouling properties of alumina and zeolite nanoparticle embedded ultrafiltration membranes. Desalination 2015, 365, 70-78. 11. Low, Z.-X.; Wang, Z.; Leong, S.; Razmjou, A.; Dumée, L. F.; Zhang, X.; Wang, H., Enhancement of the antifouling properties and filtration performance of poly (ethersulfone) ultrafiltration membranes by incorporation of nanoporous titania nanoparticles. Ind. Eng. Chem. Res. 2015, 54, (44), 11188-11198. 12. Yin, J.; Yang, Y.; Hu, Z.; Deng, B., Attachment of silver nanoparticles (AgNPs) onto thin-film composite (TFC) membranes through covalent bonding to reduce membrane biofouling. J. Membr. Sci. 2013, 441, 73-82. 13. Ben-Sasson, M.; Zodrow, K. R.; Genggeng, Q.; Kang, Y.; Giannelis, E. P.; Elimelech, M., Surface functionalization of thin-film composite membranes with copper nanoparticles for antimicrobial surface properties. Environ. Sci. Technol. 2013, 48, (1), 384-393. 14. Perreault, F. o.; Tousley, M. E.; Elimelech, M., Thin-film composite polyamide membranes functionalized with biocidal graphene oxide nanosheets. Enviro. Sci.

ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29

481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522

Environmental Science & Technology

Technol. Lett. 2013, 1, (1), 71-76. 15. Ben-Sasson, M.; Lu, X.; Bar-Zeev, E.; Zodrow, K. R.; Nejati, S.; Qi, G.; Giannelis, E. P.; Elimelech, M., In situ formation of silver nanoparticles on thin-film composite reverse osmosis membranes for biofouling mitigation. Water Res. 2014, 62, 260-70. 16. Zhang, C.; Hu, Z.; Deng, B., Silver nanoparticles in aquatic environments: Physiochemical behavior and antimicrobial mechanisms. Water Res. 2016, 88, 403427. 17. Lee, S. Y.; Kim, H. J.; Patel, R.; Im, S. J.; Kim, J. H.; Min, B. R., Silver nanoparticles immobilized on thin film composite polyamide membrane: characterization, nanofiltration, antifouling properties. Polym. Adv. Technol. 2007, 18, (7), 562-568. 18. Kim, E.-S.; Hwang, G.; El-Din, M. G.; Liu, Y., Development of nanosilver and multi-walled carbon nanotubes thin-film nanocomposite membrane for enhanced water treatment. J. Membr. Sci. 2012, 394, 37-48. 19. Yin, J.; Deng, B., Polymer-matrix nanocomposite membranes for water treatment. J. Membr. Sci. 2015, 479, 256-275. 20. Lee, H.; Dellatore, S. M.; Miller, W. M.; Messersmith, P. B., Mussel-inspired surface chemistry for multifunctional coatings. Science 2007, 318, (5849), 426-430. 21. Hong, S.; Lee, J. S.; Ryu, J.; Lee, S. H.; Lee, D. Y.; Kim, D.-P.; Park, C. B.; Lee, H., Bio-inspired strategy for on-surface synthesis of silver nanoparticles for metal/organic hybrid nanomaterials and LDI-MS substrates. Nanotechnology 2011, 22, (49), 494020. 22. Liu, Y.; Ai, K.; Lu, L., Polydopamine and its derivative materials: synthesis and promising applications in energy, environmental, and biomedical fields. Chem. Rev. 2014, 114, (9), 5057-5115. 23. Ho, C.-C.; Ding, S.-J., Structure, properties and applications of musselinspired polydopamine. J. Biomed. Nanotechnol. 2014, 10, (10), 3063-3084. 24. Yang, H.-C.; Luo, J.; Lv, Y.; Shen, P.; Xu, Z.-K., Surface engineering of polymer membranes via mussel-inspired chemistry. J. Membr. Sci. 2015, 483, 42-59. 25. Miller, D. J.; Araújo, P. A.; Correia, P. B.; Ramsey, M. M.; Kruithof, J. C.; van Loosdrecht, M. C.; Freeman, B. D.; Paul, D. R.; Whiteley, M.; Vrouwenvelder, J. S., Short-term adhesion and long-term biofouling testing of polydopamine and poly (ethylene glycol) surface modifications of membranes and feed spacers for biofouling control. Water Res. 2012, 46, (12), 3737-3753. 26. McCloskey, B. D.; Park, H. B.; Ju, H.; Rowe, B. W.; Miller, D. J.; Chun, B. J.; Kin, K.; Freeman, B. D., Influence of polydopamine deposition conditions on pure water flux and foulant adhesion resistance of reverse osmosis, ultrafiltration, and microfiltration membranes. Polymer 2010, 51, (15), 3472-3485. 27. Tang, L.; Livi, K. J. T.; Chen, K. L., Polysulfone membranes modified with bioinspired polydopamine and silver nanoparticles formed in situ to mitigate biofouling. Environ. Sci. Technol. Lett. 2015, 2, (3), 59-65.

ACS Paragon Plus Environment

Environmental Science & Technology

523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564

28. F. Wicaksana; A.G. Fane; C.Y. Tang; Wang, R., Nature Meets Technology: Forward Osmosis Membrane Technology In Biomimetic Membrane for Sensor and Separation Applications, Helix-Nielsen, C., Ed. Springer: 2012; pp 21-42. 29. Arena, J. T.; McCloskey, B.; Freeman, B. D.; McCutcheon, J. R., Surface modification of thin film composite membrane support layers with polydopamine: Enabling use of reverse osmosis membranes in pressure retarded osmosis. J. Membr. Sci. 2011, 375, (1-2), 55-62. 30. Zhang, L.; Wu, J.; Wang, Y.; Long, Y.; Zhao, N.; Xu, J., Combination of bioinspiration: a general route to superhydrophobic particles. J. Am. Chem. Soc 2012, 134, (24), 9879-9881. 31. Freger, V., Swelling and morphology of the skin layer of polyamide composite membranes: an atomic force microscopy study. Environ. Sci. Technol. 2004, 38, (11), 3168-3175. 32. Wang, Y.-N.; Wei, J.; She, Q.; Pacheco, F.; Tang, C. Y., Microscopic characterization of FO/PRO membranes–a comparative study of CLSM, TEM and SEM. Environ. Sci. Technol. 2012, 46, (18), 9995-10003. 33. Tang, C. Y.; Fu, Q. S.; Robertson, A.; Criddle, C. S.; Leckie, J. O., Use of reverse osmosis membranes to remove perfluorooctane sulfonate (PFOS) from semiconductor wastewater. Environ. Sci. Technol. 2006, 40, (23), 7343-7349. 34. Liu, X.; Jin, X.; Cao, B.; Tang, C. Y., Bactericidal activity of silver nanoparticles in environmentally relevant freshwater matrices: influences of organic matter and chelating agent. J. Environ. Chem. Eng. 2014, 2, (1), 525-531. 35. Chen, C.-Y.; Nace, G. W.; Irwin, P. L., A 6× 6 drop plate method for simultaneous colony counting and MPN enumeration of Campylobacter jejuni, Listeria monocytogenes, and Escherichia coli. J. Microbiol. Methods 2003, 55, (2), 475-479. 36. Li, M.; Xu, J.; Chang, C.-Y.; Feng, C.; Zhang, L.; Tang, Y.; Gao, C., Bioinspired fabrication of composite nanofiltration membrane based on the formation of DA/PEI layer followed by cross-linking. J. Membr. Sci. 2014, 459, 62-71. 37. Li, Y.; Su, Y.; Li, J.; Zhao, X.; Zhang, R.; Fan, X.; Zhu, J.; Ma, Y.; Liu, Y.; Jiang, Z., Preparation of thin film composite nanofiltration membrane with improved structural stability through the mediation of polydopamine. J. Membr. Sci. 2015, 476, 10-19. 38. Zhao, J.; Su, Y.; He, X.; Zhao, X.; Li, Y.; Zhang, R.; Jiang, Z., Dopamine composite nanofiltration membranes prepared by self-polymerization and interfacial polymerization. J. Membr. Sci. 2014, 465, 41-48. 39. Zhang, R.; Su, Y.; Zhao, X.; Li, Y.; Zhao, J.; Jiang, Z., A novel positively charged composite nanofiltration membrane prepared by bio-inspired adhesion of polydopamine and surface grafting of poly(ethylene imine). J. Membr. Sci. 2014, 470, 9-17. 40. Bekalé, L.; Barazzouk, S.; Hotchandani, S., Nanosilver Could Usher in Next‐Generation Photoprotective Agents for Magnesium Porphyrins. Part. Part.

ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591

Environmental Science & Technology

Syst. Charact. 2014, 31, (8), 843-850. 41. Murphy, S.; Huang, L.; Kamat, P. V., Charge-transfer complexation and excited-state interactions in porphyrin-silver nanoparticle hybrid structures. J. Phys. Chem. C 2011, 115, (46), 22761-22769. 42. Tiraferri, A.; Kang, Y.; Giannelis, E. P.; Elimelech, M., Highly hydrophilic thin-film composite forward osmosis membranes functionalized with surface-tailored nanoparticles. ACS Appl. Mater. Interfaces 2012, 4, (9), 5044-5053. 43. Liang, S.; Kang, Y.; Tiraferri, A.; Giannelis, E. P.; Huang, X.; Elimelech, M., Highly hydrophilic polyvinylidene fluoride (PVDF) ultrafiltration membranes via postfabrication grafting of surface-tailored silica nanoparticles. ACS Appl. Mater. Interfaces 2013, 5, (14), 6694-6703. 44. Zodrow, K.; Brunet, L.; Mahendra, S.; Li, D.; Zhang, A.; Li, Q.; Alvarez, P. J., Polysulfone ultrafiltration membranes impregnated with silver nanoparticles show improved biofouling resistance and virus removal. Water Res. 2009, 43, (3), 715-723. 45. Prabhawathi, V.; Sivakumar, P. M.; Doble, M., Green synthesis of protein stabilized silver nanoparticles using Pseudomonas fluorescens, a marine bacterium, and its biomedical applications when coated on polycaprolactam. Ind. Eng. Chem. Res. 2012, 51, (14), 5230-5239. 46. Li, X. l.; Zhu, L. p.; Jiang, J. h.; Yi, Z.; Zhu, B. k.; Xu, Y. y., Hydrophilic nanofiltration membranes with self-polymerized and strongly-adhered polydopamine as separating layer. Chin. J. Polym. Sci. 2012, 30, 152-163. 47. Hoek, E. M.; Elimelech, M., Cake-enhanced concentration polarization: a new fouling mechanism for salt-rejecting membranes. Environ. Sci. Technol. 2003, 37, (24), 5581-5588. 48. Tang, C. Y.; Kwon, Y.-N.; Leckie, J. O., Fouling of reverse osmosis and nanofiltration membranes by humic acid—effects of solution composition and hydrodynamic conditions. J. Membr. Sci. 2007, 290, (1), 86-94.

592 593

ACS Paragon Plus Environment