Stable graphene based membrane with pH-responsive gates for

Aug 7, 2019 - Atomic force microscopy (AFM) was used to dynamically explore the internal structure altering of GPM in the pH range from 3 to 11. It wa...
0 downloads 0 Views 1017KB Size
Subscriber access provided by UNIV AUTONOMA DE COAHUILA UADEC

Remediation and Control Technologies

Stable graphene based membrane with pHresponsive gates for advanced molecular separation Lina Zhang, Abdul Ghaffar, Xiaoying Zhu, and Baoliang Chen Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b03662 • Publication Date (Web): 07 Aug 2019 Downloaded from pubs.acs.org on August 8, 2019

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 24

Environmental Science & Technology

1

Stable graphene based membrane with pH-responsive gates for

2

advanced molecular separation

3

Lina Zhang1,2, Abdul Ghaffar1,2, Xiaoying Zhu1,2,* and Baoliang Chen1,2

4 5 6

1. Department of Environmental Science, Zhejiang University, Hangzhou, Zhejiang 310058, China.

7

2. Zhejiang Provincial Key Laboratory of Organic Pollution Process and Control, Hangzhou 310058, China.

8 9 10

*Corresponding author:

11

Dr. Xiaoying Zhu, [email protected]

12 13

Tel.: +86-571-88982651

14

Fax: +86-571-88982651

15 16 17

Co-authors:

18

Lina Zhang, [email protected]

19

Abdul Ghaffar, [email protected]

20

Baoliang Chen, [email protected]

21

1

ACS Paragon Plus Environment

Environmental Science & Technology

22

TOC

23

2

ACS Paragon Plus Environment

Page 2 of 24

Page 3 of 24

25

Environmental Science & Technology

Abstract

26

Graphene based stable pH-responsive membranes (GPMs) were developed by alternative

27

deposition of graphene oxide (GO) with polyethylenimine (PEI) in a layer-by-layer manner.

28

Different from the conventional pore-blocking pH-responsive membranes, the size-adjustable gaps

29

among the GO sheets were firstly designed to response to the surrounding pH. Atomic force

30

microscopy (AFM) was used to dynamically explore the internal structure altering of GPM in the pH

31

range from 3 to 11. It was found that the PEI molecules not only cross-linked the GO sheets through

32

amide bonds to ensure the membrane stability but also reversibly altered the gate size of GPM in a

33

certain extent according to the surrounding pH. In filtration, the gates of GPM were widening with

34

the decreasing pH of the feed and vice-versa. As a result, the permeate flux of GPM was increasing

35

with the decreasing feed pH. More importantly, the molecular weight cut-off (MWCO) of GPM

36

could be continuously regulated by the feed pH in a certain range; in filtration of the PVP and PEO2

37

mixed solution, only PVP (58 kDa) could penetrate GPM at pH: 11, while the left PEO2 (600 kDa)

38

would penetrate GPM at pH: 3. The controlled penetration through GPM led to a complete

39

separation and recovery of the molecules in different sizes, which is highly desirable for advanced

40

molecular separation in environmental applications.

41 42

Keywords:

43

pH-responsive membrane; graphene oxide; polyethylenimine; layer-by-layer assembly; molecular

44

separation

3

ACS Paragon Plus Environment

Environmental Science & Technology

45

Page 4 of 24

1. Introduction

46

The stimuli-responsive membranes are of great interests in both academia and industry, whose

47

performances such as sieving effect and fouling resistance are changing with the surrounding

48

conditions such as pH, temperature, light irradiation, electric field, concentration of chemical

49

species, and ionic strength of the feed.1-4 As an example, pore size of a thermal responsive membrane

50

prepared by assembling graphene oxide (GO) sheets grafted with poly(N-isopropylacrylamide) was

51

adjusted by temperature of the feed.5 However, the consumption of energy will be tremendous, if

52

either the feed solution or the membrane module was heated up from 25 oC to 50 oC in a typical

53

wastewater treatment plant (50,000 m3/day). On the other hand, pH adjusting is a simple and routine

54

procedure in water and wastewater treatments.4 Thus, the membrane which is responsive to pH rather

55

than the other stimuli indicated much greater potential to be practically applied in plants. It has been

56

reported that polyelectrolytes containing weak acidic or basic functional groups were incorporated

57

into polymeric membranes through blending, grafting and coating to achieve the pH-responsive

58

function.4 Since pH value of the feed solution determines ionization degree of weak acidic or basic

59

groups in the polyelectrolyte chain, the polymeric chain conformation which affects channel size and

60

surface property of the prepared membrane would be changed accordingly.4, 6 However, stability of

61

the pH-responsive membranes containing only polyions is of the main concern that swelling of

62

polyions is uncontrollable in these cases. Even cross-linking might improve the membrane stability,

63

the pH-responsive effect of the membrane will be limited because of the disordered and intertwined

64

chains in the polymeric matrix. In addition, pore blocking caused by conformation change of

65

pH-responsive polymeric chains was the mechanism of the sieving effect altering in the reported

66

pH-responsive membranes, which however may result a wide pore size distribution because of the

67

high flexibility of the polymeric chains.

68

Graphene oxide (GO) is one of the most important graphene derivatives which not only inherits

69

the desirable properties such as atomic thickness, superior chemical and mechanical stability from

70

graphene but also contains abundant oxygen-containing functional groups such as hydroxyl, epoxy

71

and carboxyl providing more possibilities of further modification and application.7-10 Moreover, GO

72

nanosheets can be easily prepared in large scale, which makes the fabrication of GO based

73

membrane (GOM) for practical applications possible.9, 11-14 GOMs were usually prepared by stacking 4

ACS Paragon Plus Environment

Page 5 of 24

Environmental Science & Technology

74

GO nanosheets through vacuum filtration,15,

16

layer-by-layer (LbL) assembly,17,

18

75

liquid crystalline.12 As a versatile method, the LbL assembly could well control thickness, interlayer

76

space and interaction of GOMs.19-21 GOMs with controlled nanochannel size by spacers were widely

77

applied in molecular separation during gas, organic solvent and water purifications.13, 22-24 However,

78

the sizes of the sieving channels of the reported GOMs were usually constant and not changing with

79

the environmental stimuli.

or casting GO

80

As a typical weak acidic group, carboxylic group is abundant in the edges and defects of GO

81

sheets.8 Obviously, GOMs may be endowed with the GO initiated pH-responsive feature.25 However,

82

GOMs constructed from GO only were suffering the same stability issue as polymeric pH-responsive

83

membranes.26, 27 Thus, GO was usually collaborated with polycations to prepare more stable GOMs

84

or control interlayer space.21 It has been reported that the ionization of carboxylic groups in GO took

85

response for the channel size shifting to a very limited extend in a GO/poly(allylamine

86

hydrochloride) multilayer membrane system because only very limited amine groups existed in the

87

GOM.17 Polyethylenimine (PEI) contains high density weakly basic amine groups which not only

88

may react with carboxylic group to form stable amide bond but also may change the conformation of

89

the PEI chain with the surrounding pH.28-30 It has been reported that PEI could either improve

90

stability of GOM31 or shift surface charge of GOM from negative to positive.32 It is of great interest

91

to study the polycation initiated pH-responsive feature of GOM. However, the pH-responsive feature

92

of the spacer molecule has never been adopted in any GOM systems.

93

In this study, novel graphene based pH-responsive membranes denoted as GPMs were

94

constructed by alternatively assembling GO nanosheets with two types of PEI in LbL manner,

95

respectively. To study the PEI initiated pH-responsive effect of GPM, both branched and linear PEI

96

molecules were used in the membrane preparation. The prepared GPMs were characterized by SEM,

97

XPS and FTIR. Afterwards, AFM and XRD were adopted to reveal the internal structure altering of

98

GPM in the pH range from 3 to 11. Subsequently, the pH-responsive performances such as the

99

permeate flux and the sieving of GPM were studied in the pH range from 3 to 11. The barrier of

100

GPM regulated by pH was applied to separate molecules with different molecular weights. It was

101

found that one single piece of GPM may respectively separate a series of molecules with different

102

molecular weights by simply adjusting pH of the mixture, which indicated great potential for

103

advanced molecular separation in practical applications such as water purification. 5

ACS Paragon Plus Environment

Environmental Science & Technology

104

2. Methods

105

2.1 Materials

Page 6 of 24

106

Natural graphite flakes (325 meshes, 99.8%) were purchased from Alfa-Aesar. Poly(vinylidene

107

fluoride) (PVDF, Mw: ~530,000) and Poly(2-ethyl-2-oxazoline) (PEOX, Mw: ~50,000, 1H NMR

108

(D2O) δH: 3.4 ppm (m), 2.3 ppm (m) and 1 ppm (m)) were provided by Sigma-Aldrich. Branched

109

polyethylenimine (PEI, Mw: ~10,000 Da), polyvinylpyrrolidone (PVP, Mw: ~58,000 Da),

110

N-N-dimethylacetamide (DMAc, >99%) were obtained from Aladdin Reagent Company (Shanghai,

111

China) and polyethylene oxides (PEO1, Mw: ~300,000 Da and PEO2, Mw: ~600,000 Da) were

112

provided by Macklin. Silver nanoparticles (Ag NP, size: ~50nm) were provided by Xianfeng Nano

113

Company (Nanjing, China). All chemicals were used without further purification.

114

2.2 Synthesis and characterization of the linear PEI

115

The linear PEI was synthesized by hydrolyzing PEOX (Figure S1). In this process, PEOX (1 g)

116

was added to a 50 mL round flask containing 20 mL of HCl solution (6 mol/L); the mixture was

117

heated to 100 oC for 4 h in a microwave digester (Hanon SH230). After cooling the mixture to room

118

temperature, the generated white turbid suspensions were collected and purified through dialysis

119

(Scientific Research Special, Mw: 3500) against DI water for a few days. Finally, the pure linear PEI

120

molecules were concentrated and freeze dried for further use.

121

Synthesis of the linear PEI was characterized by Nuclear Magnetic Resonance (NMR, Agilent

122

600) and Fourier Transform Infrared Spectrometer (FTIR, Thermo, Nicolet 6700). Molecular weight

123

of the linear PEI was characterized with gel permeation chromatography (GPC, Waters 1525/2414).

124

The chromatography conditions were as followed: using water as solvent and eluent; column:

125

PL1149-6830, PL1149-6840 and PL1149-6850 in series; flow rate: 0.8 mL/min; temperature: 30 oC.

126

Information of the linear PEI was summarized below [GPC determined Mw: 8956 Da. 1H NMR

127

(D2O) δH: 2.9 ppm (s); FTIR: 3425 cm-1, 1609 cm-1, 1139 cm-1 and 755 cm-1].

128

2.3 Preparation and characterization of GPMs

129

GO nanosheets was synthesized from natural graphite flakes by a modified Hummers’ method

6

ACS Paragon Plus Environment

Page 7 of 24

Environmental Science & Technology

130

and detailed earlier.33, 34 The support membrane was prepared by a phase inversion technique. First,

131

GO (100 mg) was sonicated in 40.9 g of DMAc for 30 min to obtain a homogeneous dispersion.

132

Thereafter, PVDF (8 g) and PVP (1 g) were added and dissolved in the dope solution at 70 °C under

133

continuous stirring in sequence to achieve a homogeneous solution. After centrifuging (3000 rpm, 5

134

min) to degas, the dope solution was uniformly casted on glass plates with a thickness of 200 µm by

135

using an automatic film applicator (BEVS1811) and a doctor blade (BEVS 1806/150). Subsequently,

136

the glass plates were immediately immersed in a coagulation bath containing DI water. The

137

solidified membranes were stored in a fresh pure water bath for overnight to ensure a complete phase

138

inversion and then dried in air.

139

For LbL deposition, a GO solution (1 mg/mL) was prepared by dissolving GO in DI water and

140

sonicating for 30 min to get a homogeneous solution; the branched and linear PEI solutions (1

141

mg/mL) were prepared by dissolving PEI in DI water. The pH values of the GO and PEI solutions

142

were all adjusted to 3.5 using NaOH (1 mol/L) and HCl (1 mol/L) solutions, respectively. Firstly, the

143

negatively charged support membrane was immersed in the PEI solution for 10 min; subsequently,

144

the membrane was taken out and rinsed with DI water to remove residual PEI. After drying under

145

nitrogen steam, the above membrane was immersed into the GO solution for 10 min; and then the

146

membrane was taken out and rinsed with DI water to remove residual GO; afterwards the membrane

147

was dried again under nitrogen stream. This cycle was repeated seven times to obtain a membrane

148

with seven bilayers of GO&PEI. Finally, the membrane was heated in an oven at 60 oC for overnight

149

to obtain the GPM membrane (Figure 1). The LbL assembly was also applied on a silicon wafer

150

surface for the subsequent AFM characterizations. The silicon wafer was pretreated with oxygen

151

plasma at 80 w for 2 min (Saot (Beijing) Optoelectronic Technology Co, Ltd.), resulting negatively

152

charged surface (Figure S2) to initiate the LbL assembly. GPMs prepared from the branched PEI

153

(Mw: 6639 Da determined by GPC) and the linear PEI (Mw: 8956 Da determined by GPC) will be

154

donated as MB and ML, respectively.

155

Surface morphology of the membranes was observed using a field emission scanning electron

156

microscope (FESEM, Hitachi S-4800, Tokyo Japan) and atomic force microscopy (AFM, Bruker

157

Dimension Icon). Membrane surface scanning by AFM was performed in air using the Bruker

158

ScanAsyst mode using a probe (ScanAsyst in air) provided by Bruker. Surface roughness of

159

membrane was calculated from the AFM height images (1×1 μm) through the software (NanoScope 7

ACS Paragon Plus Environment

Environmental Science & Technology

160

Page 8 of 24

Analysis, Bruker).

161

Thickness of the LbL film on silicon wafer was characterized by AFM. A scratch made by a

162

clean and sharp blade on the silicon wafer with the LbL film was scanned in the ScanAsyst air mode

163

using probe (ScanAsyst in air) in air and the ScanAsyst fluid mode using tip (ScanAsyst in fluid) in

164

liquid. Thickness of the LbL films can be determined by measuring the height of the step from the

165

silicon wafer substrate to top of the film in the AFM images. To minimize the influence of the

166

membrane surface unevenness in the thickness determination, the lowest points of the higher and

167

lower steps were selected to measure their vertical distance which was considered as the membrane

168

thickness. For the measurements in liquid, the membrane was fixed in a tank filled with liquid and

169

thickness shifting was monitored by scanning the same scratched step while changing the immersion

170

solutions at different pH values (from 3 to 11 and back to 3). The thickness was determined by the

171

average of five measurements at different positions for each membrane.

172

Water contact angle of GPMs was evaluated through the captive bubble method in liquid. A

173

surface analyzer (OSA200 Optical) provided by Ningbo NB Scientific Instruments Co., Ltd. was

174

used for the contact angle measurements. A membrane sample (5 cm × 1cm) was fixed in a rack and

175

immersed in a cubic tank filled with the HCl and NaOH water solutions at pH 3, 6 and 11 for 30 min

176

until equilibrium, respectively. Subsequently, an air bubble was released by a syringe with a hook

177

needle to the membrane surface; image of the bubble was taken and analyzed by the software

178

(Surface Meter) to give the water contact angle of the sample at a certain pH. Five measurements

179

were taken for each membrane at different positions, and the average value was recorded as the

180

contact angle of the membrane at the specific pH.

181

Surface chemical compositions of the membrane samples were analyzed by X-ray photoelectron

182

spectroscopy (XPS, Thermo-Fisher Scientific) and FTIR (Thermo, Nicolet 6700). Interlayer spacing

183

of the dry GPM membranes were analyzed by X-ray diffraction (XRD, Bruker D8 Advance),

184

equipped with Cu Kα radiation source (λ = 0.15418 nm). The interlayer spacing was calculated based

185

on the Bragg's law. ζ potentials of GPMs were measured using a Surpass 3 electro-kinetic analyzer

186

from Anton Paar (GmbH, Austria). GPMs were cut into 1×2 cm sliders and attached to the sample

187

holders. A 1 mM KCl aqueous background solution was used for determination of the ζ potential of

188

the membranes at different pH values from 3 to 11 adjusted by HCl (0.05 mol/L) or NaOH (0.05

189

mol/L). 8

ACS Paragon Plus Environment

Page 9 of 24

190

Environmental Science & Technology

2.4 Filtration tests of GPM at different pH values

191

The filtration experiments were carried out by using a dead-end filtration system. The detailed

192

setup of the system can be found in our pervious paper.28 A serial of working solutions were

193

prepared by adjusting pH of DI water to 3, 5, 7, 9 and 11, using the HCl and the NaOH solutions (1

194

M). In the membrane flux determination, each membrane disk (diameter: 2.2 cm) was equilibrated in

195

the working solutions for 30 min at pH: 3, 5, 7, 9 and 11, respectively, before filtration. In between

196

the filtrations at different pHs, the membrane sample was rinsed with DI water. For each filtration,

197

the membrane was pressed at 20 psi for 10 min; and then, the flux was recorded for 30 min when the

198

transmembrane pressure was stabilized at 14.5 psi (1 bar). The stable flux in the last 10 min filtration

199

was recorded as the membrane flux. Flux of each type of membrane was the average stable fluxes of

200

three membrane samples filtration at the specific pH.

201

Stability of the GPM was evaluated through measuring TOC of the permeate during the

202

filtration of the working solutions at pH: 3, 5, 7, 9, and 11. One piece of GPM was set in the filtration

203

cell that, firstly, the side with the GO/PEI multilayer was facing up; the volume of permeate

204

collected at each pH point was about 10 mL; subsequently, the same piece of GPM was turned over

205

with the GO/PEI multilayer side facing down; the same working solution series was filtrated through

206

the membrane again. TOC of the collected permeates were measured using a total organic carbon

207

analyzer (TOC-Vcph, SHIMADZU). Three membrane samples were tested for each GPM.

208

In the single component retention experiments, each of the feed solutions including 100 mg/L

209

PVP (58 kDa, its chemical structure can be seen in Figure S3a), 100 mg/L PEO1 (300 kDa, its

210

chemical structure can be seen in Figure S3b), 100 mg/L PEO2 (600 kDa) and 100 mg/L silver

211

nanoparticles (Ag NP, size: 50 nm) were prepared using the work solutions at pH: 3, 5, 7, 9 and 11,

212

respectively. The prepared solutions (10 mL) at different pH values were filtrated through the

213

membrane sample, while the permeates were collected and the solute concentrations were

214

determined accordingly. Three membrane samples were tested for each GPM to determine the

215

retention rate to the specific object (PVP, PEO1, PEO2 or Ag NP) at the specific pH value (3, 5, 7, 9

216

or 11).

217

The concentration of PVP and Ag NP was analyzed using a UV-visible spectrophotometer

218

(UV-2550 SHIMADZU) at wavelength of 214.5 and 408 nm, respectively. The concentration of 9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 24

219

PEOs was measured through a total organic carbon analyzer (TOC-Vcph, SHIMADZU).

220

Retention rate (R) of a membrane was calculated using the following equation:

221

R= (C0-Cp)/C0*100%

222

where C0 is the solute concentration of the feed solution and Cp is the solute concentration of

223

the permeate solution.

224

In the molecular mixture separation experiment, the feed solution was prepared by dissolving

225

both PVP (58 kDa) and PEO2 (600 kDa) in the working solution at pH: 11 to give concentrations

226

both at 100 mg/L. After equilibrium for 30 min, half volume (5 mL) of the feed was filtrated through

227

the membrane; subsequently, another 5 mL of the working solution at pH: 11 was added into the

228

feed; the PVP and PEO2 concentrations of the permeate were constantly monitored; the filtration and

229

refill cycle was repeated until the PVP concentration in the permeate dropped to 0. In the second

230

phase, pH of the feed was adjusted to 3 using a diluted HCl solution (1 M) and volume of the feed

231

was fixed at 10 mL; after equilibrium for 30 min, half volume (5 mL) of the feed was filtrated

232

through the membrane; subsequently, another 5 mL of the working solution at pH: 3 was added into

233

the feed; the same filtration and refill cycle was repeated until the PEO2 concentration dropped to 0.

234

Three membrane samples were tested for each GPM.

235

3. Results and discussion

236

3.1 Construction of stable GPMs with controllable gates

237

GPMs were constructed through LbL deposition of GO nanosheets and PEI molecules on a

238

negatively charged supporting membrane (Figure S2) which was prepared by imbedding GO sheets

239

in the bulk PVDF matrix, as shown in Figure 1. Obviously, GO and PEIs bore opposite charges

240

(Figure S4) at pH: 3.5, which ensured the surface charge inversion in every LbL deposition cycle.

241

Since GO sheets framed the main structure of GPM, it is necessary to integrate more GO sheets in

242

GPMs to ensure the membrane functions and reduce defects. It has been reported that the quantity of

243

weak polyelectrolyte in each layer could be precisely controlled by tuning pH of the dipping solution

244

during assembly.35, 36 Since the carboxylic groups of GO were only partially ionized at pH: 3.5, more

245

GO nanosheets were needed to compensate the positive charge of the previous PEI layer. Thus,

246

GPMs with less defects were purposely assembled at pH: 3.5.

247

The GO sheets framed structure of GPM, while the PEI molecules were playing two important 10

ACS Paragon Plus Environment

Page 11 of 24

Environmental Science & Technology

248

roles including cross-linker and spacer in the GPM system. To study the contributions of PEIs in

249

different shapes to the membrane stability and pore tuning, two types of PEI molecules in branched

250

and linear shapes were used in the GPM fabrication. The branched PEI (Mw: 6639 Da, determined

251

by GPC) was commercially available, while the linear PEI was synthesized by hydrolyzing PEOX

252

(Figure S1, more information can be seen in Figure S5 and S6). The GPM membrane prepared from

253

GO and the linear PEI or the branched PEI was denoted as ML or MB, respectively.

254

255 256

Figure 1. Schematic diagram of the stable GPMs.

257 258

Morphologies of the GPM membranes were studied through SEM and AFM. As shown in

259

Figure 2a and 2c, the surfaces of the supporting membrane (Figure S7) were homogeneously covered

260

by the GO/PEI assembly. In addition, due to the flexibility of GO sheets, the assembled GO sheets

261

were folded to show wrinkles (Figure S8). These wrinkles would be the entrances and the buffering

262

spaces of the permeation species during filtration, resulting enhanced permeability of the GO based

263

membrane.37 Similarly, the AFM images of MB (Figure 2b and S9a) and ML (Figure 2d and S9b)

264

showed smooth surfaces with low roughness (Ra) at 4.8 nm and 5.2 nm, respectively. This could be

265

attributed to the coverage of smooth planar GO nanosheets on the supporting membrane. In addition,

266

the cross-sections of MB and ML were also observed through SEM. As shown in Figure S10, the

267

dense skinny separation layer and the sponge-like supporting structure were presented both in the

268

MB and ML membranes. Owing to the strong connection between the GO/PEI multilayer and the

269

supporting membrane, it is difficult to identify the boundary between them from the SEM images

270

(Figure S10). Since the ending layers of MB and ML were both GO, the ζ potential curves of the MB

271

and ML membranes indicated negative charge in the pH range from 3 to 11 (Figure S11).

11

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 24

272 273

Figure 2. SEM (above) and AFM (below) images of (a)-(b) MB and (c)-(d) ML.

274 275

To examine the interaction between GO and PEI, chemical compositions of MB and ML were

276

characterized by FTIR and XPS. As shown in the FTIR spectra (Figure 3a), the C=O stretching

277

vibration at 1720 cm-1 of GO completely disappeared, while new peaks at 1654, 1572 and 1433 cm-1

278

which could be assigned to the amide bond (HNC=O), the N-H bond and the stretching of C-N bond,

279

respectively, showed up in the spectra of MB and ML.38 These results indicated the formation of

280

amide bonds and the existence of amine groups in MB and ML. Surface chemical compositions of

281

MB and ML were further analyzed by XPS. The XPS scans of MB and ML in Figure 3b both

282

presented the N1s peak (400 eV) rather than GO (Figure S12), indicating the introduction of

283

nitrogen-containing functional groups to GPM. In addition, the C1s spectrum of MB in Figure 3c

284

presented two new peaks at 285.5 eV and 288.6 eV belonging to C-N and N-C=O, respectively, as

285

compared with GO (Figure S12). The N1s spectrum of MB indicated two peaks at 400.1 eV and

286

398.4 eV belonging to the NC=O and C-N group, respectively, as shown in Figure S13a.33, 39, 40 The

287

C1s and N1s spectra of ML were similar to those of MB, as shown in Figure 3d and S13b, 12

ACS Paragon Plus Environment

Page 13 of 24

Environmental Science & Technology

288

respectively. Obviously, the XPS spectra confirmed the amide bond formation between GO and PEI

289

in MB and ML. The formation of covalent amide bond or cross-linking ensured the GPM membranes

290

could sustain the pH range from 3 to 11 in application. Even the amide bond formation may consume

291

a certain amount of amine groups; there were still abundant amine groups in GPM indicated by both

292

FTIR and XPS data. In addition, after calculation ratios of carbon containing functional groups based

293

on the XPS spectra of MB and ML, it was found that the percentages of carbon in NC=O were

294

10.3% and 9.3% in MB and ML, respectively, as shown in Table S1. It seems the branched PEI

295

formed more amide bonds with GO nanosheets than the linear PEI. It is possible that MB with more

296

anchoring points might present a narrower extension range by responding to the pH stimuli than ML.

297

Stability of GPMs was evaluated by measuring TOC of the permeate during filtration of the

298

working solutions in the pH range from 3 to 11. As shown in Figure S14a, after filtrating the serial

299

working solutions through MB and ML with the multilayer side facing up (feed), the TOC data of the

300

permeates were very close to DI water, when pH of the feed was in the range from 3-9; TOC of the

301

permeates were still below 1.5 mg/L, when pH of the feed was raised to 11. To further investigate

302

the membrane stability, both MB and ML were filtrated again by the above working solutions but

303

with the multilayer side facing down (permeate). Most of the TOC values measured in the tested pH

304

range were lower than 1 mg/L, as shown in Figure S14b. In general, GO nanosheets without

305

cross-linking were extremely unstable at high pH because of the ionization of their carboxylic

306

groups.41 Fortunately, both the MB and ML membranes indicated highly stable structures in the

307

working pH range from 3 to 11.

308

Stability of GPM is an important concern of this study, which was determined in two aspects:

309

first, the interaction between the supporting membrane and the LbL assembly; second, the bonding

310

among the layers of the LbL assembly. Besides the negative charges provided by both PVDF and

311

GO might attract PEI, GO contains abundant carboxylic groups which may easily react with amine

312

groups of PEI to form stable amide bond.28, 31 Both the chemical composition and the permeate TOC

313

data proved that, in the bottom-up GPM system, the GO sheets blended in the supporting membrane

314

were functioning as the anchoring points to firmly fix the multilayer GPM on top of the support;

315

furthermore, the GO nanosheets were the bricks for the GPM construction, while the PEI molecules

316

acted as the glue and spacer in the laminated GPM structure.

13

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 24

317 318

Figure 3. (a) FTIR and (b-d) XPS spectra of MB and ML.

319

3.2 Characterization of GPMs in the varied pH conditions

320

The pH-responsive performance of GPM during filtration is the main concern of this study. The

321

permeate fluxes which may be affected by the internal structure altering of GPMs were evaluated in

322

the working pH range from 3 to 11. As illustrated in Figure 4a, the permeate fluxes of MB and ML

323

were both linearly decreasing with the increasing pH of the feed solution. The flux of MB was 30.9 ±

324

2.1 L m-2 h-1 bar-1 at pH: 3, which dropped to 4.3 ± 0.6 L m-2 h-1 bar-1 at pH: 11. In addition, the flux

325

of ML was decreasing from 47.5 ± 2.8 L m-2 h-1 bar-1 to 5.8 ± 0.5 L m-2 h-1 bar-1, when pH of the feed

326

was adjusted from 3 to 11. Obviously, the permeate flux of GPM was responding to the pH changing

327

of the feed. It seems the water channels of GPM were linearly enlarged according to the pH decrease

328

of the feed during filtration. Moreover, the flux of ML was higher than that of MB at pH: 3, but

329

similar to MB at pH: 11. To investigate reversibility of the pH-responsive flux shifting of GPM, the

330

permeate flux of GPM was measured, when pH of the feed was adjusted from 11 to 3 and back to 11

331

for four cycles. As shown in Figure 4b, in the four filtration cycles, the flux variation of MB at pH: 3 14

ACS Paragon Plus Environment

Page 15 of 24

Environmental Science & Technology

332

and 11 were 7% and 7%, respectively; the flux variation of ML at pH: 3 and 11 were 2% and 4%,

333

respectively. Obviously, in the four pH adjusting cycles, the permeate fluxes of GPMs either at pH: 3

334

or at pH: 11 were relatively consistent. In short, GPM indicated stable structure and reversible

335

pH-responsive permeate flux shifting performance, which ensured reliability of the membrane in the

336

subsequently molecular separation applications.

337

To reveal mechanism of the GPM's pH-responsibility, AFM collaborated with XRD was used to

338

statically and dynamically explore the internal structure of GPM in dry and wet states. The

339

thicknesses of MB and ML in dry form, directly measured through AFM (Figure S15), were 16.1 ±

340

0.6 nm and 20.4 ± 0.4 nm (Figure 4d), respectively. Since the GPM system contained seven bilayers

341

of PEI and GO, the average thickness of each bilayer for MB and ML was 2.3 nm or 2.9 nm,

342

respectively. In addition, the thickness of the GO nanosheets was about 1 nm (Figure S16), which

343

was also determined through AFM. After deduction, the average distance among the assembled GO

344

sheets of MB was 1.3 nm which is shorter than that of ML at 1.9 nm. Interlayer distances of the

345

laminated GPMs were also determined via XRD. As shown in Figure 4c and S17, the intense 2θ

346

peaks of GO were at 9.92o for MB and 9.49o for ML, indicating d-values of 0.89 nm and 0.94 nm,

347

respectively. Even the interlayer distances estimated from the results of XRD were lower than the

348

average values measured through AFM, these data ensured that the interlayer space created by

349

inserting the branched PEI molecules among the GO sheets was smaller than that created by

350

inserting the linear PEI molecules. Moreover, thickness shifting of GPM in liquid was monitored by

351

scanning the same scratched step on the sample using AFM while changing pH of the immersion

352

solution. As shown in Figure 4d, when the pH was varied from 3 to 11, the thickness of MB or ML

353

was changing from 25.6 ± 1.4 nm to 20.3 ± 1.1 nm or from 32.2 ± 2.1 nm to 22.7 ± 2.0 nm,

354

respectively; in addition, when the pH was decreased from 11 to 3 again, the thickness of GPM

355

increased to the initial value as well. Figure S18 and S19 indicated the representative thicknesses of

356

ML and MB at pH 3 and 11, respectively.

357

Interestingly, besides the thickness of GPM, the protonation extent of the amine groups from

358

PEI were also increasing with the decreasing pH. Obviously, the conformation change of PEI was

359

linked to the pH variation through the amine group protonation extent.30 It seems that the shape of

360

the PEI chain determined by the surrounding pH manipulated the internal structure of GPM

361

reversibly. The GPM system was constructed from GO and PEI, while the PEI molecules were not 15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 24

362

only the cross-linkers to ensure the structure stability but also the spacers to control the movement of

363

the GO sheets vertically and horizontally in the laminated structure. The PEI molecule would extend

364

at low pH because of the electrostatic repulsion contributed by its protonated amine groups, which

365

might result the enlarged distance among the adjacent GO sheets. On the contrary, the PEI chain

366

would compress at high pH, which might reduce the distance among the adjacent GO sheets,

367

showing in Figure 4e. It is highly possible that the reversible thickness altering as well as water flux

368

changing of GPM triggered by surrounding pH could be attributed to the reversible transformation of

369

PEI determined by ionization/deionization of the amine groups. Even the carboxylic groups of GO

370

may be responsive to pH as well, they were not taking the main roles in the GPM system. This might

371

be attributed to the low quantity of carboxylic group in GPM and the long distance among the GO

372

sheets, which may restrict the electrostatic effects.42

373

The conformation change of PEI might make the GPM internal structure altering and water flux

374

changing possible; furthermore, the formed stable amide bonds between PEI and GO ensured the

375

vertical and horizontal movement of the GO sheets in GPM was limited in a certain range. The

376

thickness data indicated that MB was thinner than ML both in dry and wet conditions; moreover, the

377

thickness of MB could only change in a narrower range than that of ML. These phenomena should

378

be attributed to the different structures (branched and linear) of the PEI molecules. The XPS spectra

379

of MB and ML indicated that the branched PEI formed more amide bonds with GO nanosheets than

380

the linear PEI. It is possible that more anchoring points were formed in the MB membrane than ML.

381

Consequently, MB not only indicated a more compacted structure but also can swell to a smaller

382

extent than ML because the existence of more anchoring points in the MB system. These results

383

proved the flux data showing wider water flux range of ML than MB.

16

ACS Paragon Plus Environment

Page 17 of 24

Environmental Science & Technology

384 385

Figure 4. pH-responsive performances of GPMs, (a) permeate flux of MB and ML in the pH range

386

from 3 to 11; (b) permeate flux of MB and ML when pH was shifting from 11 to 3 and back to 11 for

387

4 cycles; (c) XRD spectra of MB and ML; (d) film thickness of MB and ML at different pH values;

388

(e) schematic of potential conformation change of GPM at pH: 3 and 11.

389 390

Surface hydrophilicity of MB and ML at different pH values were evaluated through the captive

391

bubble method for the equilibrium membrane samples in the working solutions. As shown in Figure

392

5, the water contact angles of MB at pH: 3, 6 and 11 were 48 ± 2o, 56 ± 3o and 50 ± 2o, respectively;

393

40 ± 5o, 45 ± 3o and 41 ±5o were the water contact angles of ML at pH: 3, 6 and 11, respectively. 17

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 24

394

Even the contact angle variation of ML at pH: 3, 6 and 11 was not very significant, MB indicated the

395

highest water contact angle (statistically significant) or the lowest hydrophilicity at pH: 6, but more

396

hydrophilic surface of MB was observed in both acidic and basic conditions. This could be attributed

397

to the ionization of the amine groups at low pH or the carboxylic groups at high pH on the GPM

398

surface. It has been reported that the protonation of amine groups in acidic condition might promote

399

the hydrophilicity of PEI.43 On the other hand, the deprotonation of GO's carboxylic groups in basic

400

condition would enhance GO's hydrophilicity as well.44 In addition, the ML membrane showed more

401

hydrophilic surface than MB at the three tested pH values, which should be contributed by the lower

402

cross-linking extent of ML (indicated by XPS data) leaving more available hydrophilic functional

403

groups. It is well known that the hydrophilicity enhancement may promote the membrane's permeate

404

flux,45 which however was not the main contributor as compared with the membrane physical

405

structure changing to the permeate flux shifting in the GPM system.

406

In short, owing to the reversible internal structure altering at varied pH values, the gates of

407

GPM could be regulated by pH accordingly for permeate flux control and precise molecular

408

separation.

409

410 411

Figure 5. Contact angle of MB and ML at different pH values.

412

3.3 pH-responsive sieving of GPMs

413

Besides the permeate flux, the sieving effect of GPM was also evaluated in the working pH

414

range from 3 to 11. Three macromolecules with different molecular weights including PVP (58 kDa),

415

PEO1 (300 kDa) and PEO2 (600 kDa) as well as one silver nanoparticle (Ag NP, 50 nm), which are 18

ACS Paragon Plus Environment

Page 19 of 24

Environmental Science & Technology

416

stable in the working pH range from 3 to 11, were adopted as the indicators in different sizes to

417

evaluate the retention and separation performances of GPMs. The highly hydrophilic substances

418

were purposely chosen to prevent the potential adsorption of the indicators to GPM. As shown in

419

Figure 6, the retention rates for the different indicators were all increasing with pH of the feed during

420

filtrations using GPM. When pH of the feed solution was changed from 3 to 11, the retention rates

421

for PVP using MB and ML were increasing from < 1% to 45.5% and from 1.5% to 26.4%,

422

respectively; the retention of PEO1 was changing from 29.6% to 88.7% and from 32.6% to 81.7%

423

for MB and ML, respectively; the retained PEO2 was shifting from 56.7% to 96.6% and from 53.9%

424

to 97.4% for MB and ML, respectively. For the largest Ag NP, both MB and ML could reach

425

retention rates which were higher than 90% in the whole working pH range from 3 to 11. The high

426

retention and recovery rate of Ag NP implied the gate size of GPM was smaller than 50 nm in the pH

427

range. The retention results further indicated that the sieving effects of ML and MB were similar and

428

both responding to pH of the feed that the gaps of the barriers to retain the indicators were narrowing

429

when the feed pH was increasing. This trend is consistent with the permeate flux changing of GPM

430

in the working pH range. Since some of the penetrated components were larger than the average

431

interlayer distance among the assembled GO sheets in GPM, the 3D channels formed by stacking the

432

2D edge-to-edge gaps beside the GO sheets in each layer, as shown in Figure 4e, were the barriers

433

for sieving, which was also reported in the literature.46 Nonetheless, both the interlayer spaces and

434

the edge-to-edge gap channels were the water channels taking response for the permeate flux of

435

GPM, as shown in Figure 4e. It seems the gates of GPM including the water channels and the

436

barriers were following the same changing trend with the surrounding pH. However, the shape

437

difference of PEI (branched and linear) resulted varied permeate fluxes but similar retention

438

performance of GPMs. Moreover, the results indicated that MB could only retain PEO2 at pH: 5,

439

while PEO1 could be well retained by MB at pH: 11. In other words, GPM may retain molecules

440

with a lower molecular weight at a higher pH. It is possible that size of the gates for sieving was

441

linearly decreasing with the increasing pH of the feed. Based on the retention rates of PEO1 and

442

PEO2, the estimated molecular weight cut-off (MWCO) range of GPM was from 750 kDa to 150

443

kDa, when pH of the feed was changing from 3 to 11. Therefore, theoretically, MWCO of GPM

444

could be continuously adjusted in the above certain range by pH of the feed. In other words, one

445

piece of GPM might separate/release substances with different sizes in the certain range respectively 19

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 24

446

by simply adjust pH of the mixture, which is desirable advanced molecular separation that cannot be

447

achieved by conventional membranes with fixed pore sizes.

448 449

Figure 6. Retention rate of single components (Ag NP, PEO2, PEO1 and PVP) by (a) MB and (b)

450

ML.

451

To evaluate the performance of GPM in molecular separation, a mixture containing PVP and

452

PEO2 was filtrated through ML at pH: 11 and subsequently at pH: 3. As shown in Figure 7, only the

453

PVP molecules could penetrate GPM, when pH of the mixture was at 11; after nine cycles of half

454

volume (5 mL) filtration and refill at pH: 11, the dosed PVP was fully recovered (recovery rate:

455

107%) (Table S2). Subsequently, the PEO2 molecules begun to permeate, when pH of the feed was

456

adjusted to 3; similarly, PEO2 from the mixture was fully recovered (recovery rate: 111%) after ten

457

cycles of half volume (5 mL) filtration and refill. These results confirmed the sieving effect of GPM

458

was responding to pH that the gaps of the barrier were widening with the decreasing pH. The

459

components (PVP and PEO2) with different molecular weights in the feed would penetrate GPM in

460

sequence, when pH of the feed was adjusted from 3 to 11, as shown in Figure 7. The molecule

461

release controlled by pH led to a complete separation of the PVP and PEO2 mixture. In addition, the

462

complete recovery of PVP and PEO2 proved the retention of PVP and PEO was contributed by

463

sieving rather than adsorption of GPM. Theoretically, the penetration of any molecules with sizes in

464

the tunable MWCO range of GPM (150-750 kDa) could be wisely controlled by the feed pH. In

465

other words, separation of multiple components with different sizes from a complicated mixture,

466

which was only possible by filtrating the mixture through numbers of membranes with fixed pore

467

sizes in series, could be accomplished by using one single piece of GPM.

468 20

ACS Paragon Plus Environment

Page 21 of 24

Environmental Science & Technology

469 470

Figure 7. Concentration of PVP and PEO2 in permeate at pH: 11 and pH: 3 during filtration of their

471

mixture through ML (the above image is a schematic diagram of the molecular separation).

472 473

In this study, the novel pH-responsive membranes (GPMs) were constructed by alternatively

474

assembling the GO nanosheets with the PEI molecules on a supporting PVDF membrane imbedded

475

with GO. The PEI molecules not only cross-linked the GO sheets through amide bonds to ensure the

476

membrane stability but also determined the pH-responsive effects of GPM. The reversible

477

conformation change of PEI contributed by the protonation of amine groups at different pH values

478

made the GPM internal structure altering possible; furthermore, the formed stable amide bonds

479

between PEI and GO ensured the vertical and horizontal movement of the GO sheets in GPM was

480

limited in a certain range. It seems the gates of GPM including the water channels and the barriers

481

were narrowing with the increasing pH of the feed and vice-versa. As a result, the permeate flux

482

governed by the water channels was decreasing but the retention rate determined by the barriers was

483

increasing, when the feed pH was increasing. Nevertheless, the branched PEI formed more amide

484

bonds with GO resulting a more compacted membrane (MB) which could swell to a smaller extent

485

than ML prepared from the linear PEI and GO in the pH range from 3 to 11. Thus, the permeate flux

486

of MB was lower than ML, when the PEI molecules were in the extended state at pH: 3. Moreover,

487

the MWCO of one GPM may be continuously regulated by the feed pH in a certain range, which was

488

adopted to allow the penetration of PVP (58 kDa) at pH: 11 and then PEO2 (600 kDa) at pH: 3 from

489

their mixture resulting a completely separation and recovery. Theoretically, the penetration of any 21

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 24

490

molecules with sizes in the tunable MWCO range of GPM could be wisely controlled by the feed

491

pH, which indicated a sustainable and efficient way of advanced molecular separation in

492

environmental applications such as water purification and resource recovery.

493 494

Supporting Information

495

Chemical bond content of MB and ML based on the XPS spectra, concentration of PVP and

496

PEO2 in permeate at pH:11 and pH:3 after different filtration cycles using ML, synthesis of the

497

linear PEI, ζ potential of the substrates and the membranes, chemical structure of PVP and PEO, ζ

498

potential of PEI and GO, NMR and FTIR spectra of linear PEI and PEOX, SEM and AFM images of

499

the membranes, XPS spectra of GO, MB and ML, TOC of the permeate after filtration of the

500

working solutions at different pHs, thickness image determined of GO, MB and ML by AFM, XRD

501

spectra of MB and ML.

502 503

Acknowledgments

504

This work was supported by the National Natural Science Foundation of China (21425730,

505

21621005 and 21607124), the National Key Research and Development Program of China

506

(2017YFA0207000) and the Qianjiang River Talent Plan.

507 508

Reference

509 510 511 512 513 514 515 516 517 518 519 520 521 522 523

1.

Nunes, S. P.; Behzad, A. R.; Bobby Hooghan; Sougrat, R.; Karunakaran, M.; Pradeep, N.; Vainio, U.; Peinemann,

K.-V., Switchable pH-responsive polymeric membranes prepared via block copolymer micelle assembly. ACS Nano 2011, 5, 3516-3522. 2.

Dutta, K.; De, S., Smart responsive materials for water purification: an overview. J. Mater. Chem. A 2017, 5, (42),

22095-22112. 3.

Wandera, D.; Wickramasinghe, S. R.; Husson, S. M., Stimuli-responsive membranes. J. Membr. Sci. 2010, 357, (1-2),

6-35. 4.

Zhao, C.; Nie, S.; Tang, M.; Sun, S., Polymeric pH-sensitive membranes—A review. Prog. Polym. Sci. 2011, 36, (11),

1499-1520. 5.

Liu, J.; Wang, N.; Yu, L. J.; Karton, A.; Li, W.; Zhang, W.; Guo, F.; Hou, L.; Cheng, Q.; Jiang, L.; Weitz, D. A.; Zhao, Y.,

Bioinspired graphene membrane with temperature tunable channels for water gating and molecular separation. Nat Commun 2017, 8, (1), 2011. 6.

Lee, D.; Nolte, A. J.; Kunz, A. L.; Rubner, M. F.; Cohen, R. E., pH-Induced hysteretic gating of track-etched

polycarbonate membranes: swelling/deswelling behavior of polyelectrolyte multilayers in confined geometry. J. Am. Chem. Soc. 2006, 128, 8521-8529. 22

ACS Paragon Plus Environment

Page 23 of 24

524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567

7.

Environmental Science & Technology

Sun, P.; Zhu, M.; Wang, K.; Zhong, M.; Wei, J.; Wu, D.; Xu, Z.; Zhu, H., Selective ion penetration of graphene oxide

membranes. ACS Nano 2013, 7, 428-437. 8.

Perreault, F.; Fonseca de Faria, A.; Elimelech, M., Environmental applications of graphene-based nanomaterials.

Chem. Soc. Rev. 2015, 44, 5861-5896. 9.

Joshi, R. K.; Carbone, P.; Wang, F. C.; Kravets, V. G.; Su, Y.; Grigorieva, I. V.; H. A. Wu; Geim, A. K.; Nair, R. R.,

Precise and ultrafast molecular sieving through graphene oxide membranes. Science 2014, 343, 752-754. 10. Ji Won Suk, R. D. P., Jinho An, and Rodney S. Ruoff, Mechanical Properties of Monolayer Graphene Oxide. ACS Nano 2010, 4, 6557-6564. 11. Sun, P.; Zheng, S.; Zhu, M.; Song, Z.; Wang, K.; Zhong, M.; Wu, D.; Little, R. B.; Xu, Z.; Zhu, H., Selective trans-membrane transport of alkali and alkaline earth cations through graphene oxide membranes based on cation-π interactions. ACS Nano 2014, 8, 850-859. 12. Akbari, A.; Sheath, P.; Martin, S. T.; Shinde, D. B.; Shaibani, M.; Banerjee, P. C.; Tkacz, R.; Bhattacharyya, D.; Majumder, M., Large-area graphene-based nanofiltration membranes by shear alignment of discotic nematic liquid crystals of graphene oxide. Nat Commun 2016, 7, 10891. 13. 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), 442-4. 14. Hu, W.; Peng, C.; Luo, W.; Lv, M.; Li, X.; Li, D.; Huang, Q.; Fan, C., Graphene-based antibacterial paper. ACS Nano 2010, 4, 4317-4323. 15. Xu, W. L.; Fang, C.; Zhou, F.; Song, Z.; Liu, Q.; Qiao, R.; Yu, M., Self-Assembly: A facile way of forming ultrathin, high-performance graphene oxide membranes for water purification. Nano Lett 2017, 17, 2928-2933. 16. Han, Y.; Xu, Z.; Gao, C., Ultrathin graphene nanofiltration membrane for water purification. Adv. Funct. Mater. 2013, 23, (29), 3693-3700. 17. Oh, Y.; Armstrong, D. L.; Finnerty, C.; Zheng, S.; Hu, M.; Torrents, A.; Mi, B., Understanding the pH-responsive behavior of graphene oxide membrane in removing ions and organic micropollulants. J. Mem. Sci. 2017, 541, 235-243. 18. Song, X.; Zambare, R. S.; Qi, S.; Sowrirajalu, B. N.; James Selvaraj, A. P.; Tang, C. Y.; Gao, C., Charge-gated ion transport through polyelectrolyte intercalated amine reduced graphene oxide membranes. ACS Appl. Mater. Interfaces 2017, 9, (47), 41482-41495. 19. Xiao, F. X.; Pagliaro, M.; Xu, Y. J.; Liu, B., Layer-by-layer assembly of versatile nanoarchitectures with diverse dimensionality: a new perspective for rational construction of multilayer assemblies. Chem. Soc. Rev. 2016, 45, (11), 3088-121. 20. Richardson, J. J.; Bjornmalm, M.; Caruso, F., Multilayer assembly. Technology-driven layer-by-layer assembly of nanofilms. Science 2015, 348, (6233), aaa2491. 21. Mi, B., Graphene oxide memranes for ionic and molecular sieving. Science 2014, 343, 740-742. 22. Sun, P.; Wang, K.; Zhu, H., Recent developments in graphene-based membranes: structure, mass-transport mechanism and potential applications. Adv. Mater. 2016, 28, 2287-2310. 23. Lee, C. S.; Choi, M. K.; Hwang, Y. Y.; Kim, H.; Kim, M. K.; Lee, Y. J., Facilitated water transport through graphene oxide membranes functionalized with aquaporin-mimicking peptides. Adv Mater 2018, 30, (14), e1705944. 24. Huang, L.; Chen, J.; Gao, T.; Zhang, M.; Li, Y.; Dai, L.; Qu, L.; Shi, G., Reduced graphene oxide membranes for ultrafast organic solvent nanofiltration. Adv Mater 2016, 28, (39), 8669-8674. 25. Huang, H.; Mao, Y.; Ying, Y.; Liu, Y.; Sun, L.; Peng, X., Salt concentration, pH and pressure controlled separation of small molecules through lamellar graphene oxide membranes. Chem. Commun. 2013, 49, (53), 5963-5965. 26. Zhu, X.; Yang, K.; Chen, B., Membranes prepared from graphene-based nanomaterials for sustainable applications: a review. Environ. Sci.: Nano 2017, 4, (2267), 2267-2285. 27. Zheng, S.; Tu, Q.; Urban, J. J.; Li, S.; Mi, B., Swelling of graphene oxide membranes in aqueous solution: 23

ACS Paragon Plus Environment

Environmental Science & Technology

568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606

Page 24 of 24

characterization of interlayer spacing and insight into water transport mechanisms. ACS Nano 2017, 11, (6), 6440-6450. 28. Zhang, L.; Chen, B.; Ghaffar, A.; Zhu, X., Nanocomposite membrane with polyethylenimine-grafted graphene oxide as a novel additive to enhance pollutant filtration performance. Environ. Sci. Technol. 2018, 52, (10), 5920-5930. 29. Sharma, K. P.; Choudhury, C. K.; Srivastava, S.; Davis, H.; Rajamohanan, P. R.; Roy, S.; Kumaraswamy, G., Assembly of polyethyleneimine in the hexagonal mesophase of nonionic surfactant: effect of pH and temperature. J. Phys. Chem. B 2011, 115, (29), 9059-9069. 30. Choudhury, C. K.; Roy, S., Structural and dynamical properties of polyethylenimine in explicit water at different protonation states: a molecular dynamics study. Soft Matter 2013, 9, (7), 2269. 31. Nam, Y. T.; Choi, J.; Kang, K. M.; Kim, D. W.; Jung, H. T., Enhanced stability of laminated graphene oxide membranes for nanofiltration via interstitial amide bonding. ACS Appl. Mater. Interfaces 2016, 8, 27376-27382. 32. Nan, Q.; Li, P.; Cao, B., Fabrication of positively charged nanofiltration membrane via the layer-by-layer assembly of graphene oxide and polyethylenimine for desalination. Appl. Surf. Sci. 2016, 387, 521-528. 33. Wang, J.; Chen, Z.; Chen, B., Adsorption of polycyclic aromatic hydrocarbons by graphene and graphene oxide nanosheets. Environ. Sci. Technol. 2014, 48, 4817-4825. 34. Chen, X.; Chen, B., Direct observation, molecular structure, and location of oxidation debris on graphene oxide nanosheets. Environ. Sci. Technol. 2016, 50, 8568-8577. 35. Zhu, X.; Janczewski, D.; Guo, S.; Lee, S. S. C.; Velandia, F. J. P.; Teo, S. L.-M.; He, T.; Puniredd, S. R.; Vancso, G. J., Polyion Multi layers with Precise Surface Charge Control for Antifouling. ACS Appl. Mater. Interfaces 2015, 7, (1), 852-861. 36. Zhu, X.; Guo, S.; He, T.; Jiang, S.; Janczewski, D.; Vancso, G. J., Engineered, robust polyelectrolyte multilayers by precise control of surface potential for designer protein, cell, and bacteria adsorption. Langmuir 2016, 32, 1338-1346. 37. Qi, B.; He, X.; Zeng, G.; Pan, Y.; Li, G.; Liu, G.; Zhang, Y.; Chen, W.; Sun, Y., Strict molecular sieving over electrodeposited 2D-interspacing-narrowed graphene oxide membranes. Nat. Commun. 2017, 8, (1), 825. 38. Erno, P.; Philippe, B.; Martin, B., Structure determination of organic compounds. 2009. 39. Liu, H.; Kuila, T.; Kim, N.; Ku, B.; Lee, J., In situ synthesis of the reduced graphene oxide–polyethyleneimine composite and its gas barrier properties. J. Mater. Chem. A 2013, 1, 3739-3746. 40. Crist, B. V., Handbook of The Elements and Native Oxides. 1999. 41. 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. 2014, 7, (2), 166-170. 42. Amadei, C. A.; Montessori, A.; Kadow, J. P.; Succi, S.; Vecitis, C. D., Role of oxygen functionalities in graphene oxide architectural laminate subnanometer spacing and water transport. Environ. Sci. Technol. 2017, 51, 4280-4288. 43. Xia, F.; Feng, L.; Wang, S.; Sun, T.; Song, W.; Jiang, W.; Jiang, L., Dual-responsive surfaces that switch between superhydrophilicity and superhydrophobicity. Adv. Mater. 2006, 18, (4), 432-436. 44. Wan, S.; Pu, J.; Zhang, X.; Wang, L.; Xue, Q., The tunable wettability in multistimuli-responsive smart graphene surfaces. Appl. Phys. Lett. 2013, 102, (1), 011603. 45. Zhu, X.; Loo, H.; Bai, R., A novel membrane showing both hydrophilic and oleophobic surface properties and its non-fouling performances for potential water treatment applications. J. Membr. Sci. 2013, 436, 47-56. 46. Wei, N.; Peng, X.; Xu, Z., Understanding water permeation in graphene oxide membranes. ACS Appl. Mater. Interfaces 2014, 6, (8), 5877-5883.

607

24

ACS Paragon Plus Environment