Metalized Nanocellulose Composites as a Feasible Material for

Mar 20, 2017 - Herein, we study the feasibility of using nanocellulose (NC)-based composites with silver and platinum nanoparticles as additive materi...
2 downloads 5 Views 6MB Size
Subscriber access provided by Fudan University

Article

Metalized nanocellulose composites as a feasible material for membrane supports: Design and applications for water treatment Perla Cruz-Tato, Edwin O. Ortiz-Quiles, Karlene Vega-Figueroa, Liz Noemi Santiago-Martoral, Michael Flynn, Liz M Diaz-Vazquez, and Eduardo Nicolau Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b05955 • Publication Date (Web): 20 Mar 2017 Downloaded from http://pubs.acs.org on March 21, 2017

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 31

Environmental Science & Technology

1

Metalized nanocellulose composites as a feasible material for

2

membrane supports: Design and applications for water treatment

3

Perla Cruz-Tato1,2, Edwin O. Ortiz-Quiles1,2, Karlene Vega-Figueroa1,2, Liz Santiago-Martoral1,2,

4

Michael Flynn3, Liz M. Díaz-Vázquez1and Eduardo Nicolau*1,2

5

1

6

Puerto Rico USA 00931-3346

7

2

8

Juan Puerto Rico USA 00931-3346

9

3

Department of Chemistry, University of Puerto Rico, Rio Piedras Campus, PO Box 23346, San Juan,

Molecular Sciences Research Center, University of Puerto Rico, 1390 Ponce De Leon Ave, Suite 2, San

NASA Ames Research Center, Bioengineering Branch, Moffett Field, California 94036 USA

10

Corresponding Author

11

*Phone: 787-292-9820, fax: 787-522-2150; email: [email protected]

12

13 14 15

16

17 18 19 20 21

ACS Paragon Plus Environment

Environmental Science & Technology

22

Abstract

23

Herein, we study the feasibility of using nanocellulose (NC)-based composites with silver and platinum

24

nanoparticles as additive materials to fabricate the support layer of thin film composite (TFC) membranes

25

for water purification applications. In brief, the NC surface was chemically modified and then was

26

decorated with silver and platinum nanoparticles, respectively by chemical reduction. These metalized

27

nanocellulose composites (MNC) were characterized by several techniques including: FTIR, XPS, TGA,

28

XRD and XANES to probe their integrity. Thereafter, we fabricated the MNC-TFC membranes and the

29

support layer was modified to improve the membrane properties. The membranes were thoroughly

30

characterized and the performance was evaluated in forward osmosis (FO) mode with various feed

31

solutions: nanopure water, urea and wastewater samples. The fabricated membranes exhibited finger-like

32

pore morphologies and varying pore sizes. Interestingly, higher water fluxes and solute rejection was

33

obtained with the MNC-TFC membranes with wastewater samples. The overall approach of this work

34

provides an effort to fabricated membranes with high water flux and enhanced selectivity.

35

KEYWORDS

36

Nanocellulose, water purification, membrane supporting layer, metal nanoparticles

37 38 39 40 41 42 43

ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

Environmental Science & Technology

44

Introduction

45

According to the World Health Organization (WHO) there are 663 million of human beings worldwide

46

with limited access to fresh water for consumption and hygiene. Moreover, water-related diseases are

47

known to affect more than 1.5 billion people every year.1 Thus, finding cost-effective approaches for

48

water recycling, water purification and wastewater reclamation are of urgent need.

49

Current state-of-the-art for water reclamation strategies rely on the use of membrane-based processes. The

50

most widely used membrane processes in water purification applications are microfiltration (MF),2-3

51

ultrafiltration (UF),4-5 nanofiltration (NF)6-7 and reverse osmosis (RO).8-9 The membrane characteristics

52

for these processes are very similar and only vary in the pore size of the membrane for each application.

53

Nevertheless, these technologies employ high pressures to drive the flux of water making the overall

54

process energy-intensive.10 In contrast, osmotically driven membrane processes have the potential to

55

sustainably produce and reclaim water.11 Forward osmosis (FO) is a membrane process driven by the

56

osmotic pressure gradient across a semi-permeable membrane (Figure 1), where water is extracted across

57

the membrane from a regime of low osmotic pressure in the feed solution (FS) to a high-osmotic-pressure

58

in the draw solution (DS).12 Forward osmosis has captured the attention of researchers mainly because it

59

is a technology that helps alleviate the issues encountered with the aforementioned technologies (e.g. high

60

pressure requirements, fouling and concentrate management).13 This membrane-based water separation

61

process is less prone to fouling because it generates a lower transmembrane pressure. Moreover, FO

62

membranes also allow for high water flux with minimum energy consumption, making FO a feasible and

63

cost-effective approach for wastewater treatment. Even when FO exhibits lower membrane fouling, the

64

technology still suffers from some fouling effects. The membrane fouling process in osmotically driven

65

membranes is known to be different in pressure driven processes due to the low hydraulic pressure being

66

employed.13 Interestingly, membrane fouling may improve the solute rejection and permeability of the FO

67

membrane. 14-15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 31

68

Figure 1. Representation of the FO membrane process 69

In the area of FO membranes the use of the thin film composites (TFC) polyamide membranes are

70

considered at the forefront for the actual application. These membranes are composed of a porous

71

support, typically polysulfone (PSF), and a non-porous highly crosslinked polyamide selective layer. TFC

72

polyamide membranes can achieve higher water permeability and comparable salt rejection than

73

asymmetric cellulose-acetate-based membranes, such as the commercial membrane from Hydration

74

Innovation Technologies (HTI-CTA).16 The rational design of TFC membranes for FO processes requires

75

a thin support layer with high porosity and low tortuosity in order to increase the flux of water across the

76

membrane. Moreover, the support layer must be hydrophilic while providing chemical and thermal

77

stability within reasonable mechanical strength.17

78

A variety of technologies and approaches has been studied to improve the membrane performance (e.g.

79

higher water fluxes), most of these are focused on chemical modifications and in the development of

80

mixed matrix membranes, i.e. using additives. Reported chemical modifications and mixed matrix

81

membranes include: polymers18-19 and its variations

20-21

, nanoparticles22-23, organic

ACS Paragon Plus Environment

24

and inorganic

Page 5 of 31

Environmental Science & Technology

82

compounds.25-26 These modifications are performed via chemical or physical interactions on the active

83

side or in the support layer of the membrane.27

84

As previously shown by Tarabara27, the use of nanoparticles in polymeric membranes may have direct

85

influence on important characteristics such as the porosity, the pore size and the morphology of the

86

asymmetric support layer which affects the water permeability, the salt rejection and the interfacial

87

polymerization reaction. Also an increase in the thickness and changes in the contact angle can be

88

obtained, depending on the nature of the nanoparticles. Nevertheless, metallic nanoparticles are

89

thermodynamically unstable and without the use of capping agents, ligands or supports tend to

90

aggregation.28 Nanocellulose (NC) is considered an ideal support for depositing metal nanoparticles

91

owing to their high surface area-to-volume ratio, reductive surface, mechanical strength (i.e. Young’s

92

Module in the range of 100-130GPa), biodegradability, non-toxicity and low cost.29-31 In fact, NC

93

biopolymer has been recently considered as a candidate for novel water treatment technologies, and have

94

been shown that low loadings of NC into polymeric membrane support layer have the effect of increasing

95

the tensile strength while exhibiting increased porosity, larger pore sizes and enhanced water permeability

96

when compared to other materials.32

97

In this study, we present the synthesis and characterization of metallized nanocellulose (MNC)

98

composites as candidate materials to incorporate in the support layer of a forward osmosis membrane to

99

improve its performance. The nanocellulose-based composites combine the advantages of both the guest

100

nanomaterial and the nanocellulose substrate that are known to exhibit synergetic properties.29 As

101

mentioned previously, the inclusion of metallic nanoparticles in the structural framework of the

102

membrane may help enhance the membrane properties and performance. Therefore in this work we

103

studied two types of metal nanoparticles (silver and platinum) as they both exhibit particular and

104

interesting properties for further applications. The silver nanoparticles (AgNP) have been widely studied

105

as antimicrobial agent33, and the platinum nanoparticles (PtNP) are well-known for its catalytic

ACS Paragon Plus Environment

Environmental Science & Technology

106

properties.34 The nanoparticles were immobilized in the support layer as a way to decrease the risk of

107

leaching.27, 33

108

By linking the advantages of the different materials we present the composites characterization and

109

respective membrane performance. The combination of NC with the advantages of metallic nanoparticles

110

could present a way to generate outstanding support layers for water purification membranes with

111

possible capabilities beyond reclaiming water.

112

Materials and Methods

113

2.1 Materials

114

Nanocellulose (NC, 11.8%) aqueous solution was purchased from University of Maine Process and

115

Development Center (Onoro, USA). 3-aminopropyltriethoxysilane (APTES, 97%), silver nitrate (AgNO3,

116

99%), potassium tetrachloroplatinate (K2PtCl4, 99.99%), sodium borohydride (NaBH4, 99%), polysulfone

117

(PSF, average Mn ~22,000), N-methyl-2-pyrrolidone (NMP, 99%), m-phenylenediamine (MPD, 99%), 1,

118

3, 5-benzenetricarbonyl trichloride (TMC, 98%), hexane (anhydrous 95%), sodium chloride (NaCl, ACS

119

reagent 99.0%) and urea (ACS reagent, 99.0-100.5%) were all purchased from Sigma-Aldrich. All

120

chemicals and solvents were used as received without further purification. Nanopure water (18.2

121

MΩ·cm2, MilliQ Direct 16) was used at all times.

122

2.2 Preparation of amino-modified nanocellulose (NC-NH2)

123

The surface of the NC was modified to incorporate amino silane functionalities following published

124

procedures with minor modifications.35 Amino modified nanocellulose (NC-NH2) were prepared by

125

reacting 15 mL of a 3% w/v aqueous solution of NC and 15 mL of APTES 0.1M in 80% EtOH. First, the

126

NC 3% solution was placed on a hot plate at 60ºC under constant stirring while APTES solution was

127

slowly added to the reaction vessel. The reaction was carried out under these conditions for 2 hours, after

128

completing the reaction, the NC-NH2 composite was allowed to reach ambient temperature.

ACS Paragon Plus Environment

Page 6 of 31

Page 7 of 31

Environmental Science & Technology

129

2.3 Deposition of silver and platinum nanoparticles to NC-NH2

130

The deposition of the metal nanoparticles to the NC-NH2 was performed by an in-situ chemical reduction

131

of the precursor materials using sodium borohydride as a reducing agent. Briefly, 10 mL of a 0.05M

132

AgNO3 solution was thoroughly mixed with 30 mL of the NC-NH2. Thereafter, 15 mL of a 0.1 M NaBH4

133

solution was used to titrate the mixture promoting the reduction of the silver ions in solution to silver

134

metal nanoparticles (NC-NH2Ag). This reaction was conducted on an ice bath under constant stirring. The

135

mixture was stirred for 15 minutes after reaction. The composites were washed in a vacuum filtration

136

system using nanopure water and a 50% v/v ethanol solution. The same procedure was employed to

137

generate the platinum NC-based composite (NC-NH2Pt) where 10 mL of a 0.05 M K2PtCl4 solution was

138

thoroughly mixed with 30mL of the NC-NH2. The mixture was titrated with 15 mL of the reducing agent,

139

the solution turned black after addition of reducing agent and left reacting for 15 minutes and, then

140

filtered. All composites were lyophilized and pulverized for further characterization.

141

2.4 Physical and chemical characterization of nanocellulose based composites

142

Freeze-dried samples were analyzed, the infrared spectra were recorded on a Thermo Scientific

143

TM Nicolet TM Continuµm TM infrared (IR) microscope using transmittance mode. X-ray diffraction

144

measurements were conducted over 10 to 90º 2θ range using a Rigaku SmartLab diffractometer at 40 kV

145

and 44 mA equipped with monochromatic CuKα (1.54 Å) X-ray source. X-Ray Photoelectron

146

spectroscopy (XPS) binding energy were obtained using a PHI 5600 spectrometer equipped with an Al

147

Kα mono and polychromatic X-ray source operating at 15 kV, 350 W and pass energy of 58.70 eV. A

148

Perkin-Elmer STA 6000 simultaneous thermal analyzer was used to measure the changes in weight and

149

heat flow as a function of temperature. X-Ray Absorption spectroscopy (XAS) was performed at Cornell

150

High Energy Synchrotron Source (CHESS) using line F3. The Ag K edge (25,514 eV) and Pt L3 edge

151

(11,563 eV) were measured for the NC-NH2Ag and NC-NH2Pt samples in transmission mode. The

152

antimicrobial properties were tested using the Kirby-Bauer diffusion method and the electrochemical

ACS Paragon Plus Environment

Environmental Science & Technology

153

profile was evaluated by cyclic voltammetry. Detailed information of these measurements is included at

154

the Supporting Information.

155

2.5 Fabrication of NC-TFC mesh-embedded supports

156

The NC-PSF membranes were prepared using the non-solvent induced phase separation (NIPS) process

157

as published elsewhere.36 The dope solution was prepared with 12% w/v PSF and 0.5% w/v NC-

158

composites in NMP. To dissolve the materials, the solutions were stirred at room temperature for 24

159

hours. The polyester (PE) mesh was attached to a clean glass plate, and a dust magnet was employed to

160

remove the particles over the surface. The solution was casted over the PE using a casting knife adjusted

161

to 150 µm. The film was immediately immersed in a nanopure water precipitation bath for 10 minutes

162

before transferring the membrane to a nanopure water bath for storage. Thereafter, the NC-TFC

163

membranes were fabricated by fixing the NC-PSF membranes over a glass plate to allow coating of only

164

the top surface. A 2% w/v MPD aqueous solution was poured to the membranes until it was completely

165

soaked for 2 minutes and the excess was removed using an air knife. Next, a 0.1% w/v TMC aqueous

166

solution in hexane was poured to the membranes for 1 minute and the excess was removed by using an air

167

knife. Finally, the membrane was soaked in 0.2% w/v sodium carbonate solution for 5 minutes, rinsed

168

and stored in nanopure water.

169

2.6 Characterization of MNC-TFC membranes

170

Scanning electron microscopy images were performed in a high-resolution field emission JEOL JSM-

171

7500F SEM to analyze the surface morphology of the NC-TFC membranes. Samples were lyophilized,

172

placed in sample holders and sputtered with a thin gold film (ca. 20 nm thick). The TFC membranes

173

performance was tested using an in-house FO system (Figure 2) with nanopure water (18 MΩ·cm2), urea

174

solution (1.3% w/v) and wastewater samples as the feed. These membranes were screened against

175

commercially available membranes Hydration Technology Innovation (HTI-CTA). The exposed area of

176

the membrane was 4.25 cm2, the active layer was facing the feed solution and the results were collected at

ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

Environmental Science & Technology

177

constant pressure (1 bar). To evaluate the performance of the membrane, water flux rate from the feed to

178

the draw solution side is determined using the following equation:

‫= ݓܬ‬

߂ܸ ‫ܲ߂ݐ ܯܣ‬

179

(1)

180

where, Jw is the water flux rate in LMH/bar (L·m-2·h-1/bar), ∆V is the volume increment in the osmotic

181

solution in L, AM is active membrane area, t is time or duration of test in hours and ∆P is the

182

transmembrane pressure in bar.

183

The concentration of salt in the feed was determined by ion chromatography, using a cation column:

184

CS12 and anion column: AS4A with a conductivity detector and preparing a calibration curve with a

185

standard solution of ions. The reverse salt flux from the draw solution to the feed side was obtained with

186

the following equation:

187

Js =

188

where Js is the reverse salt flux in GMH (g·m-2·h-1), Ct and Vt are the salt concentration and the feed

189

volume at the end of the FO tests, respectively. Total organic carbon content was also determined using a

190

TOC from Shimadzu model TOC-LCCH.

∆(CtVt ) AM t

(2)

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 31

Figure 2. Experimental setup diagram for the FO performance test. The 1-hour experiments were performed by recirculating the feed and draw solutions continuously with the aid of a peristaltic pump through the custom-made cell containing the membrane. The arrows represent the flow of each solution and the conductivity was monitored at each solution using the conductometers. 191

Results and discussion

192

3.1 Metalized nanocellulose-based composites characterization

193

Considering the abovementioned properties attributed to NC, the use of this biopolymer as an additive for

194

functional materials presents an interesting approach. In this study, the surface of the nanocellulose was

195

modified and functionalized through the addition of amino-silane groups. Presumably, the incorporation

196

of -SiO and -NH functional groups provide more nucleation sites where metallic nanoparticles can be

197

localized at the expense of a slight decrease in NC hydrophilicity.37-38 In order to assure chemical integrity

198

of modified-NC, FT-IR and XPS were employed as characterization tools. Figure S1 shows the FTIR

199

spectrum of NC and the amino modified-NC where absorption signals are constant suggesting that NC

200

structural functional groups remain unaltered after the synthesis process. An obvious absorption band is

ACS Paragon Plus Environment

Page 11 of 31

Environmental Science & Technology

201

observed at ca. 1560 cm-1 attributed to the bending vibration of primary amines (N-H).39 The

202

characteristic two bands of the N-H stretching vibration are observed in the region of 3500-3200 cm-1

203

while presence of silanes (νSi-O) can be observed at absorption signals around 820 cm-1.

204

Moreover, XPS was performed and survey results confirm the presence of carbon, oxygen, nitrogen and

205

silicon in the NC-NH2 composite (see supporting information, Figure S2a). Also, to further evaluate the

206

integrity of the modified material a high-energy photoelectron spectrum deconvolution procedure was

207

performed, see Figure S2b. FTIR and XPS results confirm successful incorporation of the amino-silane

208

group to the NC surface.

209

The deposition of metallic nanoparticles within the NC-NH2 network was achieved via a wet chemical

210

approach and using a reducing agent. We performed a thermogravimetric analysis of the composites to

211

study the thermal stability of the materials. Changes in the thermal decomposition transitions can account

212

to verify the presence of the metal nanoparticles (see Figure S3). The TG curve suggests a metal loading

213

of 12% Ag and 20% Pt. More important, is the XRD analysis for all the materials shown in Figure S4

214

where it can be observed the characteristic diffraction signals for NC (type I and II) at low angles for all

215

composites,30 suggesting the presence of NC with a crystalline structure, which provides an indication

216

that the crystallinity of the NC was not perturbed by the chemical reaction at the surface. Moreover, the

217

presence of metallic nanoparticles was verified, and crystalline phases for Ag and Pt are in agreement

218

with the literature.40-41

219

Additional XPS analyses were also utilized in this study to account for the presence of all the elements

220

and to obtain information of the main interactions between elements in each composite. Survey results

221

confirm the presence of carbon, oxygen, nitrogen and silicon in all composites whereas silver and

222

platinum were also detected in the respective composites. Moreover, Figure S5 (a) and (b) presents the

223

high-resolution photoelectron spectrum deconvolution of the composites and suggests the presence of the

224

metallic nanoparticles in both the metal and oxide forms. To further elucidate the electronic structure and

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 31

225

provide qualitative information of the electronic environment within the composites, X-Ray Absorption

226

Near Edge Spectroscopy (XANES) was performed. The XANES spectra of Ag K-edge and Pt LIII edge,

227

respectively, suggest the presence of the nanoparticles mainly in their reduced form, which is in

228

agreement with the XPS results (see Figure S5 c and d). We evaluate the electronic environment of the

229

composites containing the metal nanoparticles because differences at the nanoparticle surface could

230

generate significant changes in the support layer. These changes may include hydrophilicity, which will

231

directly affect the water permeability and also the membrane selectivity.27

232

Physical characterizations confirmed the successful modification of the nanocellulose with the amino-

233

silane group and the deposition of metal nanoparticles. The principal objective of incorporating AgNP

234

and PtNP was due to the inherent properties of each material, antimicrobial and catalytic respectively. We

235

evaluated the NC-NH2Ag biocidal potential using a Kirby-diffusion test (Figure S6) suggesting

236

antimicrobial properties for AgNP. The NC-NH2Pt electrochemical properties were studied by cyclic

237

voltammetry (Figure S7), which indicate that the PtNP are electrochemically active. Even though this

238

work focused on the effect of these metal nanoparticles at the supporting layer, these types of materials

239

could also be incorporated at the membrane’s surface in the future to generate reactive FO membranes.

240

3.2 Nanocellulose-based thin film composite (NC-TFC) membranes

241

The fabricated mesh-reinforced thin film composite membranes for FO were prepared using the phase

242

inversion method, illustrated in Figure 3. A polyester (PE) mesh was employed as mechanical reinforcing

243

material followed by a porous mid-layer made of polysulfone (PSF) and the NC-composites. Afterwards,

244

the membrane’s surface was modified with a thin polyamide layer through the interfacial polymerization

245

of MPD and TMC to improve the permeability, the selectivity and the water fluxes of the membranes.42

246

Recent investigations have explored the advantages and the importance of the rational design and

247

development of PSF supports for FO applications.11,

248

formation is one of the most important factors to evaluate. Herein, we used the phase inversion method,

36, 43

When fabricating a FO membrane, pore

ACS Paragon Plus Environment

Page 13 of 31

Environmental Science & Technology

249

which is governed by the solvent exchange rate. Depending on this solvent exchange rate, also known as

250

the precipitation rate, the pores morphology and size may vary. The polymer precipitation rates affect the

251

pore morphology as rapid precipitation creates finger-like pores and slow precipitation produces sponge-

252

like pores. Membranes with finger-like pores present higher water fluxes while sponge-like pores present

253

higher salt rejection. NMP was used as the solvent because it has been known that induces a high

254

precipitation rate where the non-solvent influx (i.e. water) directs the net flux during the phase inversion

255

process.44

Figure 3. Two-step fabrication of NC embedded TFC membranes. 256

ACS Paragon Plus Environment

Environmental Science & Technology

a) a)

f)

b)

g) g)

c)

h)

d) d)

i)

e)

j)j)

Figure 4. SEM micrographs of non-polymerized support layer (a-e) active side and (f-j) back side of the membrane, where P represents polysulfone (PSF). 257

ACS Paragon Plus Environment

Page 14 of 31

Page 15 of 31

Environmental Science & Technology

258

Figure 4 shows the top (a)-(e) and bottom (f)-(j) SEM micrographs of the non-polymerized support layers

259

used for the TFC-membranes. The membrane’s surface porosity plays a key role in the interfacial

260

polymerization reaction because it essentially determines the amount of MPD molecules that can

261

penetrate the support layer. As can be observed in the figure, the surface of the non-polymerized supports

262

present homogeneous pores sizes in the range of ca. PtNP

339

suggesting that the particle size affects the FO performance of the membrane. The reverse salt flux was

340

calculated and similarly, all membranes showed a low reverse salt flux. The selectivity of the membrane

341

can be assessed from the total organic carbon (TOC) rejection. The TOC was measured in both solutions

342

(feed and draw solution) before and after each run. All the membranes present high TOC rejection values

343

although the commercial membrane still has the higher performance (96.9%) followed by the NC (94.6%)

344

and PSF (90.3%) membranes. The membrane with NC-NH2 (90.3%) presented the lowest rejection, which

345

also suggests that the urea molecules in the draw solution could contribute to higher gradient in osmotic

346

pressure. The incorporation of metal nanoparticles to the membranes, also seem to have an opposite trend

347

in the TOC rejection, AgNP (92.4%) < PtNP (94.0%).

348

Another parameter that helps to determine membrane selectivity is the specific reverse salt flux, Js/Jw.

349

The specific reverse salt flux takes into consideration the amount of draw solute leakage per unit of water

350

permeated across the membrane. As shown in Figure S10, all membranes have a low Js/Jw value (less

351

than 0.22 g/L) with both feed solutions (DI water and urea) meaning that the leakage of salt is minimum

352

compared to the permeated water through the membrane.

353

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 31

354

Figure 8. Puerto Nuevo Regional Wastewater Treatment Plant: municipalities45 and facilities46, and a schematic illustration of primary treatment (from wastewater to primary effluent) and the experimental setup employed for the FO membrane system. 355

As a proof-of-concept and in an effort to assess the effectiveness of the fabricated membranes we tested

356

the membranes containing the composites with the metal nanoparticles using wastewater from the Puerto

357

Nuevo Regional Wastewater Treatment Plant (RWWTP), a primary water treatment plant in Puerto Rico.

358

The Puerto Nuevo RWWTP is located in San Juan, Puerto Rico and serves the municipalities of the

359

metropolitan area (Figure 8). The facility’s layout includes a pumping station, mechanical bar screen, grit

360

removal mechanism, primary clarifiers, sludge handling facilities and a disinfection area. The treated

361

effluent from the Puerto Nuevo RWWTP is then discharged approximately 7,365 ft (2,246 m) from the

362

shoreline into the Atlantic Ocean.47 Figure 8 also shows a schematic representation of the overall

363

treatment that samples from RWWTP received, starting at the wastewater treatment plant through the FO

364

system employed in this work.

ACS Paragon Plus Environment

Page 23 of 31

Environmental Science & Technology

a)

b)

Figure 9. Results after wastewater treatment: a) SEM micrographs, and b) table summarizing the bacteria per membrane area and the obtained water fluxes with each membrane 365

We analyze the membrane surface using SEM after the FO performance evaluation with raw wastewater.

366

The micrographs are shown in Figure 9a, and when compared to the membranes before the FO process

367

(Figure 5) significant differences can be noticed. After the performance test structures consistent with the

368

basic shape and size of bacteria48 appears, suggesting the attachment of microorganisms. A semi-

369

quantitative analysis of the SEM micrographs (0.12 mm x 0.09 mm) was performed in order to compare

370

the number of microorganisms per area of membrane. The results are summarized in Figure 9b, and it

371

suggests that fewer microorganisms were able to attach onto the fabricated membranes. Specifically, the

372

P:NC-NH2Ag membrane reduced in 87% and 67% the amount of microorganisms on the membrane’s

373

surface when compared to the HTI-CTA and P:NC-NH2Pt, respectively. Nguyen, T. et al.49 have

374

previously identified the surface roughness, surface hydrophilicity, surface charge and membrane material

375

as the main factors that affect the attachment of microorganisms onto a membrane. Based on these

376

factors, all of the fabricated membranes are similar in all of the surface properties (roughness,

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 31

377

hydrophilicity and charge) due to the polyamide layer, but for the membrane material. Therefore,

378

differences (i.e. microorganisms per area) between the MNC-TFC membranes are due to the nature of the

379

nanoparticles.

380

We measured the efficiency of the membranes by the water fluxes and contaminant rejection using GC-

381

MS. The water fluxes obtained are shown in Figure 9b and it can be observed that the performance of the

382

fabricated membranes was higher than the HTI-CTA. As in our previous trend, the membrane containing

383

the AgNP exhibited higher water fluxes suggesting that the pore distribution also enhanced its

384

performance using a real wastewater sample. The FO membrane performance with various feed solutions

385

were compared and we recorded an increment in the water flux for both MNC-TFC membranes when raw

386

wastewater was used as the feed solution: P:NC-NH2Ag increased 15% and P:NC-NH2Pt 33%. It was also

387

of interest to determine the identity of the compounds that cross over the membrane. In order to

388

accomplish this we utilized a GC-MS to account for the contaminants in the draw solution after the

389

experiment, see Figure S11. These results suggest that the fabricated membranes can be considered as

390

promising materials as to their high rejection of contaminants. Unexpectedly, both membranes containing

391

the metallic nanoparticles have higher contaminant rejection than the HTI-CTA. We can notice that the

392

membranes P:NC-NH2Ag and P:NC-NH2Pt only allow the passage across the membrane of linear alkanes

393

and ester-based contaminants while the HTI-CTA also allows other contaminants: aromatics, cyclic and a

394

wide variety of molecules with other functional groups. Therefore, a significant difference can be

395

appreciated when comparing the results from the MNC-TFC with the HTI-CTA membranes. The MNC-

396

TFC allows for higher water fluxes and also higher contaminant rejection.

397

Closer examination of the SEM micrographs (Figure 9a) allows to further identify the organized

398

structures at the membrane’s surface. These organized structures are more apparent in the P:NC-NH2Pt

399

surface, by its particular arrangement and similarity to reported results48, 50 we can speculate that a biofilm

400

forms at the membrane surface. As previously reported,14, 49, 51 studies utilizing raw wastewater promotes

401

membrane fouling and this is considered as one of the major problems in membrane-based processes23, 51.

ACS Paragon Plus Environment

Page 25 of 31

Environmental Science & Technology

402

However it could also play a favorable role in the water extraction and solute rejection of the

403

membranes.15 By virtue of the polyamide layer properties (higher roughness, more hydrophobicity and

404

higher negative charge), the MNC-TFC membranes might suffer from biofilm formation. In regards to the

405

P:NC-NH2Ag membrane it is apparent that a biofilm did not formed, yet an improvement in water flux

406

and contaminant rejection was obtained. Overall, the results with wastewater samples suggest a better

407

performance for the fabricated MNC-TFC membranes than the HTI-CTA.

408

In summary, we successfully synthesized promising MNC-based composites that exhibit electrochemical

409

and antimicrobial activity and certainly could be utilized for FO membrane fabrication. These composites

410

were well characterized by several techniques, which verified the crystallinity of NC after surface

411

modifications and confirmed the presence of each metal nanoparticle (Ag and Pt). We fabricated the TFC

412

membranes using the NIPS technique and the support layer of the membranes was modified by the

413

incorporation of the NC-based composites. The fabricated membranes exhibited finger-like pore

414

morphologies and varying pore sizes. These results directly affect the FO performance using different

415

feed solutions: DI water, urea aqueous solution and raw wastewater samples. Interestingly, higher water

416

fluxes and solute rejection were obtained with the MNC-TFC membranes when using wastewater sample

417

due to the formation of a biofilm. This approach provides an effort to produce membranes with high water

418

fluxes, high selectivity and low reverse salt fluxes that could further degrade contaminants and prevent

419

bacterial growth. In future studies we will address the electrochemical and bactericidal properties of the

420

MNC-TFC membranes, to further improve their properties under operating conditions.

421

BRIEFS

422

This work presents the synthesis, characterization and application of metalized nanocellulose composites

423

for the fabrication of forward osmosis membranes.

424 425

SYNOPSIS (TOC)*

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 31

426

427

AKNOWLEDGEMENTS

428

This work was supported by the NASA Experimental Program to Stimulate Competitive Research

429

(EPSCoR) under grant #NNX14AN18A and the NASA Advanced STEM Training and Research

430

(ASTAR) Fellowship under grant #NNX15AU27H.

431

The authors acknowledge the UPR Materials Characterization Center (MCC) for the provided support

432

during the attainment of this work. We would also like to thank Institute for Functional Nanomaterials

433

Nanoscopy Facility and the CHESS facilities for their help supported by the NSF & NIH/NIGMS via

434

NSF award DMR-1332208. Also we want to thank Dina Bracho, Valerie Ortíz and Dr. Carlos González

435

for their help with the microbiological assays. Also, we would like to thank Luis Betancourt for the X-ray

436

absorption analysis and Dr. Craig Priest, from the University of South Adelaide for the contact angle

437

measurements. Finally, special thanks are due to Arnulfo Rojas for helping with the GC-MS analysis.

438

AUTHORS INFORMATION

439

Corresponding Author, * Prof. Eduardo Nicolau, [email protected]

440

Perla Cruz-Tato, [email protected]

441

Edwin Ortiz-Quiles, [email protected]

442

Karlene Vega-Figueroa, [email protected]

443

Liz Santiago Martoral, [email protected]

ACS Paragon Plus Environment

Page 27 of 31

Environmental Science & Technology

444

Michael Flynn, [email protected]

445

Liz M. Díaz-Vázquez, [email protected]

446

Present Addresses

447

(E.N.) Department of Chemistry, University of Puerto Rico, Rio Piedras Campus, PO Box 23346, San

448

Juan, Puerto Rico USA 00931-3346// Molecular Science Research Center, University of Puerto Rico,

449

1390 Ponce De Leon Ave, Suite 2, San Juan Puerto Rico USA 00931-3346

450

Author Contributions

451

The manuscript was written through contributions of all authors. All authors have given approval to the

452

final version of the manuscript. PCT is the main author of this work.

453

Notes

454

The authors declare no competing financial interest.

455

Supporting Information

456 457 458

Supporting information contains: FTIR, XRD, XPS, XANES, microbial and electrochemical characterizations of the MNC-composites, AFM and SEM images, roughness, contact angle and GC-MS results of the prepared membranes. The material is complementary to the main work.

459

References

460 461 462 463 464 465 466 467 468 469 470 471 472 473 474

1. WHO/UNICEF, Progress on sanitation and drinking water. 2015; p 90. 2. Ma, H.; Burger, C.; Hsiao, B. S.; Chu, B., Nanofibrous microfiltration membrane based on cellulose nanowhiskers. Biomacromolecules 2012, 13 (1), 180-186. 3. Okamura, D.; Mori, Y.; Hashimoto, T.; Hori, K., Effects of microbial degradation of biofoulants on microfiltration membrane performance in a membrane bioreactor. Environ Sci Technol 2010, 44 (22), 8644-8648. 4. Su, Y.; Mu, C.; Li, C.; Jiang, Z., Antifouling Property of a Weak Polyelectrolyte Membrane Based on Poly(acrylonitrile) during Protein Ultrafiltration. Industrial & Engineering Chemistry Research 2009, 48 (6), 3136-3141. 5. Zhao, S.; Wang, Z.; Wei, X.; Zhao, B.; Wang, J.; Yang, S.; Wang, S., Performance Improvement of Polysulfone Ultrafiltration Membrane Using Well-Dispersed Polyaniline–Poly(vinylpyrrolidone) Nanocomposite as the Additive. Industrial & Engineering Chemistry Research 2012, 51 (12), 4661-4672. 6. Zhao, F. Y.; Ji, Y. L.; Weng, X. D.; Mi, Y. F.; Ye, C. C.; An, Q. F.; Gao, C. J., High-Flux Positively Charged Nanocomposite Nanofiltration Membranes Filled with Poly(dopamine) Modified Multiwall Carbon Nanotubes. ACS Appl Mater Interfaces 2016, 8 (10), 6693-6700.

ACS Paragon Plus Environment

Environmental Science & Technology

475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525

Page 28 of 31

7. Kim, J. H.; Choi, D. C.; Yeon, K. M.; Kim, S. R.; Lee, C. H., Enzyme-immobilized nanofiltration membrane to mitigate biofouling based on quorum quenching. Environ Sci Technol 2011, 45 (4), 16011607. 8. Ray, J. R.; Tadepalli, S.; Nergiz, S. Z.; Liu, K. K.; You, L.; Tang, Y.; Singamaneni, S.; Jun, Y. S., Hydrophilic, bactericidal nanoheater-enabled reverse osmosis membranes to improve fouling resistance. ACS Appl Mater Interfaces 2015, 7 (21), 11117-11126. 9. Huang, X.; Marsh, K. L.; McVerry, B. T.; Hoek, E. M.; Kaner, R. B., Low-Fouling Antibacterial Reverse Osmosis Membranes via Surface Grafting of Graphene Oxide. ACS Appl Mater Interfaces 2016, 8 (23), 14334-8. 10. Le, N. L.; Nunes, S. P., Materials and membrane technologies for water and energy sustainability. Sustainable Materials and Technologies 2016, 7, 1-28. 11. Yip, N. Y.; Tiraferri, A.; Phillip, W. A.; Schiffman, J. D.; Elimelech, M., High performance thinfilm composite forward osmosis membrane. Environ Sci Technol 2010, 44 (10), 3812-3818. 12. Saren, Q.; Qiu, C. Q.; Tang, C. Y., Synthesis and characterization of novel forward osmosis membranes based on layer-by-layer assembly. Environ Sci Technol 2011, 45 (12), 5201-5208. 13. Raval, H. D.; Koradiya, P., Direct fertigation with brackish water by a forward osmosis system converting domestic reverse osmosis module into forward osmosis membrane element. Desalination and Water Treatment 2015, 57 (34), 15740-15747. 14. Mafirad, S.; Mehrnia, M. R.; Azami, H.; Sarrafzadeh, M. H., Effects of biofilm formation on membrane performance in submerged membrane bioreactors. Biofouling 2011, 27 (5), 477-85. 15. Valladares Linares, R.; Yangali-Quintanilla, V.; Li, Z.; Amy, G., Rejection of micropollutants by clean and fouled forward osmosis membrane. Water Res 2011, 45 (20), 6737-44. 16. Werber, J. R.; Osuji, C. O.; Elimelech, M., Materials for next-generation desalination and water purification membranes. Nature Reviews Materials 2016, 1 (5), 16018-16033. 17. Huang, L.; McCutcheon, J. R., Impact of support layer pore size on performance of thin film composite membranes for forward osmosis. Journal of Membrane Science 2015, 483, 25-33. 18. Mural, P. K. S.; Kumar, B.; Madras, G.; Bose, S., Chitosan Immobilized Porous Polyolefin As Sustainable and Efficient Antibacterial Membranes. ACS Sustainable Chemistry & Engineering 2016, 4 (3), 862-870. 19. Bui, N. N.; McCutcheon, J. R., Hydrophilic nanofibers as new supports for thin film composite membranes for engineered osmosis. Environ Sci Technol 2013, 47 (3), 1761-9. 20. Hilke, R.; Pradeep, N.; Madhavan, P.; Vainio, U.; Behzad, A. R.; Sougrat, R.; Nunes, S. P.; Peinemann, K. V., Block copolymer hollow fiber membranes with catalytic activity and pH-response. ACS Appl Mater Interfaces 2013, 5 (15), 7001-6. 21. Qiu, X.; Yu, H.; Karunakaran, M.; Pradeep, N.; Nunes, S. P.; Peinemann, K. V., Selective separation of similarly sized proteins with tunable nanoporous block copolymer membranes. ACS Nano 2013, 7 (1), 768-76. 22. 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 2014, 48 (1), 384-93. 23. Liu, C.; Faria, A. F.; Ma, J.; Elimelech, M., Mitigation of Biofilm Development on Thin-Film Composite Membranes Functionalized with Zwitterionic Polymers and Silver Nanoparticles. Environ Sci Technol 2017, 51 (1), 182-191. 24. Raval, H. D.; Trivedi, J. J.; Joshi, S. V.; Devmurari, C. V., Flux enhancement of thin film composite RO membrane by controlled chlorine treatment. Desalination 2010, 250 (3), 945-949. 25. Romanos, G. E.; Athanasekou, C. P.; Likodimos, V.; Aloupogiannis, P.; Falaras, P., Hybrid Ultrafiltration/Photocatalytic Membranes for Efficient Water Treatment. Industrial & Engineering Chemistry Research 2013, 52 (39), 13938-13947. 26. Zhang, Y.; Pinoy, L.; Meesschaert, B.; Van der Bruggen, B., A natural driven membrane process for brackish and wastewater treatment: photovoltaic powered ED and FO hybrid system. Environ Sci Technol 2013, 47 (18), 10548-55.

ACS Paragon Plus Environment

Page 29 of 31

526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575

Environmental Science & Technology

27. Advanced Materials for Membrane Preparation. M.G. Buonomenna, G. G., Ed. Bentham Science: 2012; p. 295. 28. Astruc, D., Transition-metal Nanoparticles in Catalysis: From Historical Background to the Stateof-the Art. 2007, 1-48. 29. Wei, H.; Rodriguez, K.; Renneckar, S.; Vikesland, P. J., Environmental science and engineering applications of nanocellulose-based nanocomposites. Environ. Sci.: Nano 2014, 1 (4), 302-316. 30. Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J., Cellulose nanomaterials review: structure, properties and nanocomposites. Chem Soc Rev 2011, 40 (7), 3941-3994. 31. Kaushik, M.; Moores, A., Review: nanocelluloses as versatile supports for metal nanoparticles and their applications in catalysis. Green Chem. 2016, 18 (3), 622-637. 32. Carpenter, A. W. d. L., C.-F.; Wiesner, M. R., Cellulose Nanomaterials in Water Treatment Technologies. Environmental Science & Technology 2015, 49 (9), 5277-5287. 33. Park, S. Y.; Chung, J. W.; Chae, Y. K.; Kwak, S. Y., Amphiphilic thiol functional linker mediated sustainable anti-biofouling ultrafiltration nanocomposite comprising a silver nanoparticles and poly(vinylidene fluoride) membrane. ACS Appl Mater Interfaces 2013, 5 (21), 10705-10714. 34. Jang, J.-H.; Kim, J.; Lee, Y.-H.; Kim, I. Y.; Park, M.-H.; Yang, C.-W.; Hwang, S.-J.; Kwon, Y.U., One-pot synthesis of core–shell-like Pt3Co nanoparticle electrocatalyst with Pt-enriched surface for oxygen reduction reaction in fuel cells. Energy & Environmental Science 2011, 4 (12), 4947-4956. 35. Koga, H.; Kitaoka, T.; Isogai, A., In situ modification of cellulose paper with amino groups for catalytic applications. Journal of Materials Chemistry 2011, 21 (25), 9356-9361. 36. Stillman, D.; Krupp, L.; La, Y.-H., Mesh-reinforced thin film composite membranes for forward osmosis applications: The structure–performance relationship. Journal of Membrane Science 2014, 468, 308-316. 37. Lee, L.; Leroux, Y. R.; Hapiot, P.; Downard, A. J., Amine-terminated monolayers on carbon: preparation, characterization, and coupling reactions. Langmuir 2015, 31 (18), 5071-5077. 38. Taokaew, S. P., M.; Newby, B.-m. Z., Modification of bacterial cellulose with organosilanes to improve attachment and spreading of human fibroblasts. Cellulose 2015, 22 (4), 2311-2324. 39. Ahire, J. H.; Wang, Q.; Coxon, P. R.; Malhotra, G.; Brydson, R.; Chen, R.; Chao, Y., Highly luminescent and nontoxic amine-capped nanoparticles from porous silicon: synthesis and their use in biomedical imaging. ACS Appl Mater Interfaces 2012, 4 (6), 3285-3292. 40. Smetana, A. B.; Klabunde, K. J.; Marchin, G. R.; Sorensen, C. M., Biocidal activity of nanocrystalline silver powders and particles. Langmuir 2008, 24 (14), 7457-7464. 41. Xia, H.; Yang, G., Facile synthesis of inorganic nanoparticles by a precipitation method in molten ε-caprolactam solvent. Journal of Materials Chemistry 2012, 22 (35), 18664-18670. 42. Xie, W.; Geise, G. M.; Freeman, B. D.; Lee, H.-S.; Byun, G.; McGrath, J. E., Polyamide interfacial composite membranes prepared from m-phenylene diamine, trimesoyl chloride and a new disulfonated diamine. Journal of Membrane Science 2012, 403-404, 152-161. 43. Ghosh, A. K.; Hoek, E. M. V., Impacts of support membrane structure and chemistry on polyamide–polysulfone interfacial composite membranes. Journal of Membrane Science 2009, 336 (1-2), 140-148. 44. Strathmann, H.; Kock, K., The formation mechanism of phase inversion membranes. Desalination 1977, 21 (3), 241-255. 45. USDA Census of Agriculture - 2012 Census Publications - Island and Municipio Profiles - Puerto Rico.http://www.agcensus.usda.gov/Publications/2012/Online_Resources/County_Profiles/Puerto_Rico / (accessed November 2016). 46. CSA-Group Puerto Rico Water & Wastewater Infrastructure Program. http://www.csagroup.com/project.php?msid=1&pid=134 (accessed November 2016). 47. EPA FACT SHEET FOR DRAFT NATIONAL POLLUTANT DISCHARGE ELIMINATION SYSTEM (NPDES) PERMIT TO DISCHARGE INTO THE WATERS OF THE UNITED STATES: NPDES Permit No. PR0021555; Environmental Protection Agency: 2010.

ACS Paragon Plus Environment

Environmental Science & Technology

576 577 578 579 580 581 582 583 584

Page 30 of 31

48. O'Toole, G.; Kaplan, H. B.; Kolter, R., Biofilm formation as microbial development. Annual review of microbiology 2000, 54, 49-79. 49. Nguyen, T. R., F. A.; Fan, L., Biofouling of Water Treatment Membranes: A Review of the Underlying Causes, Monitoring Techniques and Control Measures. Membranes 2012, 2, 804-840. 50. Pang, C. M.; Hong, P.; Guo, H.; Liu, W.-T., Biofilm Formation Characteristics of Bacterial Isolates Retrieved from a Reverse Osmosis Membrane. Environmental Science & Technology 2005, 39 (19), 7541-7550. 51. Rana, D.; Matsuura, T., Surface modifications for antifouling membranes. Chem Rev 2010, 110 (4), 2448-2471.

585

ACS Paragon Plus Environment

Page 31 of 31

Environmental Science & Technology

SYNOPSIS (TOC)

 

ACS Paragon Plus Environment