Catalytic Membrane Reactor Immobilized with Alloy Nanoparticle

Sep 13, 2016 - †Tianjin Key Laboratory of Indoor Air Environmental Quality Control, School of Environmental Science and Engineering, ‡State Key La...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Massachusetts Amherst Libraries

Article

Catalytic Membrane Reactor Immobilized with Alloy NanoparticlesLoaded Protein Fibrils for Continuous Reduction of 4-Nitrophenol Renliang Huang, Hongxiu Zhu, Rongxin Su, Wei Qi, and Zhimin He Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b03431 • Publication Date (Web): 13 Sep 2016 Downloaded from http://pubs.acs.org on September 18, 2016

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

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

1

2

Catalytic Membrane Reactor Immobilized with Alloy

3

Nanoparticles-Loaded Protein Fibrils for Continuous Reduction of

4

4-Nitrophenol

5

Renliang Huang,†, ‖ Hongxiu Zhu,‡, ‖ Rongxin Su,*,‡,§,# Wei Qi,‡,§,# and Zhimin He‡

6



School of Environmental Science and Engineering, Tianjin University, Tianjin 300072, PR China

7



State Key Laboratory of Chemical Engineering, School of Chemical Engineering and Technology,

8

Tianjin University, Tianjin 300072, PR China

9

§

Collaborative Innovation Center of Chemical Science and Engineering (Tianjin),Tianjin 300072,

10

PR China

11

#

12

Tianjin 300072, PR China

Tianjin Key Laboratory of Membrane Science and Desalination Technology, Tianjin University,

13 14 15



16

* Correspondence concerning this article should be addressed to R. Su at [email protected]

R. Huang and H. Zhu contributed equally to this work.

17

1

18

ABSTRACT

19

A catalytic membrane reactor, which contains a membrane matrix and a catalytic film of alloy

20

nanoparticle-loaded β-lactoglobulin fibrils (NPs@β-LGF), was developed for the continuous-flow

21

reduction of 4-nitrophenol (4-NP). The Cu-Ag and Cu-Ag-Au alloy NPs were synthesized using

22

β-LGF as a scaffold and stabilizing agent. In this process, the Cu nanoclusters were formed in the

23

initial stage and were able to promote the synthesis of Ag0, which acts as a reducing agent for the

24

rapid formation of Au0. Furthermore, a catalytic membrane reactor was constructed by depositing

25

the NPs@β-LGFs on a membrane matrix. The catalytic activity of the Cu-Ag-Au alloy NPs was

26

higher than that of the Cu-Ag alloy NPs, using the reduction of 4-NP to 4-AP as a model reaction.

27

The observed rate constant in the continuous-flow system is also higher than that in the batch

28

system. In addition, these catalytic membrane reactors had good operating stability and antibacterial

29

activity.

30

KEYWORDS: catalytic membrane reactor, continuous-flow, alloy nanoparticles, β-lactoglobulin

31

2

32



INTRODUCTION

33

Membrane reactors (MRs), in which both the catalytic reaction and separation are performed in a

34

single unit, have attracted considerable attention for their application in the fields of environmental

35

engineering,1-3 energy production, 4, 5 and chemical engineering.6, 7 Generally, MRs can be classified

36

as inert membrane reactors (IMRs) and catalytic membrane reactors (CMRs). The membrane in the

37

IMRs is adjacent to the catalytic reaction zone containing the suspended or packed catalysts, while

38

the membrane in the CMRs is either embedded with the catalysts or possesses catalytic activity

39

itself.6 Regarding the combination of the reaction and separation, MRs are widely used in

40

continuous-flow systems, which provide high operability, easy reuse of the catalysts, easy control of

41

the reaction, automation, and scale up. However, the systems that are equipped with IMRs often

42

suffer from the limited mass transport because the membrane pores are blocked by the suspended

43

catalysts (e.g., enzymes, nanoparticles) or dead-end pores are present in the packed catalyst beads.

44

Because the catalysts are embedded in the membranes, CMRs are enable to reduce this mass

45

transport limitation and are attracting an increased amount of research interest for their industrial

46

applications. It is worth nothing that the catalyst loading will inevitably influence the membranes’

47

properties in many cases; therefore, it is essential to determine how to construct the catalytic

48

membrane reactor to simultaneously achieve good catalytic and separation performance.

49

Metal nanoparticles (NPs) are widely used as catalysts in environmental engineering due to their

50

excellent catalytic performance, such as Au nanoparticles,8, 9 zerovalent iron,10-14 and Pt

51

nanoparticles.15, 16 However, these nanocatalysts aggregate easily, reducing their activity, and thus

52

are undesirable for industrial applications.17 Strategies to address this problem by immobilizing

53

these nanoparticles on carriers, such as active carbon,18 metallic oxide,19, 20 protein crystals,21, 22 and 3

54

mineral chabazite,23 have met with some success. The column reactors packed with these

55

immobilized catalysts are generally applied to the continuous-flow systems, but they have a

56

disadvantage in terms of the mass transport limitation mentioned above. Alternatively, a promising

57

strategy is to construct CMRs via the integration of the metal nanoparticles within the membranes.

58

For instance, Seto et al 24 recently reported a new catalytic membrane reactor immobilized with a

59

palladium (Pd)-loaded polymer nanogel for continuous-flow Suzuki coupling reaction. The Pd

60

nanoparticles were synthesized in situ on the membrane and exhibited high catalytic activity and

61

long-term stability. In addition to the in situ growth, embedding the prepared metal NPs is also a

62

good choice because it is easy to control the synthesis of the NPs. In this study, we attempt to

63

develop an efficient approach for the synthesis of metal NPs and construct an NP-loaded catalytic

64

membrane reactor.

65

Currently, the fabrication of metal NPs is dominated by solution-based techniques using

66

dissoluble metal salts. Inspired by the mineralization in biology, one of the most attractive

67

directions for this research is the use of proteins and peptides as reducing and/or stabilizing agents

68

to direct the synthesis of metal NPs with controlled shape, size, uniformity, and stability.25 Using

69

the Au NPs as an example, various proteins, such as bovine serum albumin (BSA), α-amylase,

70

trypsin, and lysozyme, have been utilized as templates, because they contain a large number of

71

specific amino acid residues (e.g., cysteine, tyrosine, or arginine).25-28 However, the spherical

72

structure, small size, and low mechanical strength of these proteins cannot match the demand of the

73

CMRs. On the contrary, protein fibrils, which are regarded as pathogenic proteins in many

74

neurodegenerative disorders,29 have very long contour lengths (>5 μm) and possess exceptional

75

mechanical properties (Young’s modulus: 0.2-14 GPa).30 In previous studies, some protein fibrils

4

76

were made from affordable protein sources in vitro and used to fabricate hybrid composite

77

films,31-33 hybrid aerogels,34 etc. For example, Li et al 31 demonstrated that graphene and

78

β-lactoglobulin (β-LG) fibrils can be combined to create a new class of biodegradable composite

79

film with shape-memory and enzyme-sensing properties. The Young's modulus of the β-LG fibrils

80

(β-LGFs) is in the range of 2-4 GPa and is comparable to the value of silk.35 Moreover, β-LG fibril

81

is also a good template for the synthesis of inorganic nanomaterials. For instance, Mezzenga and

82

co-workers synthesized the metal (gold, silver and palladium) nanoparticles,36 gold single

83

crystals,37-39 TiO2 nanowires,40 and Fe3O4/β-LG fibrils (or nanoclusters) hybrids 41 for diverse

84

applications. Due to its high strength and easy availability,42 the β-LG fibril is a promising

85

candidate to serve as the catalyst scaffold for the construction of the metal NP-loaded CMRs.

86

Recently, Bolisetty et al 43 reported a catalytic membrane containing gold or palladium

87

nanoparticles immobilized on amyloid hybrid fibrils for continuous flow reaction. The metal NPs

88

were synthesized on the surface of β-LGFs using NaBH4 as reducing agent. Here, we confirm these

89

earlier findings on the efficiency of metal-β-LGF hybrids membrane for continuous flow catalysis

90

and we further expand to metal alloys-β-LGF hybrids membranes. So far, fabrication of bimetallic

91

NPs in membranes was reported by some groups; for example, Su and co-workers developed the in

92

situ synthesis of Au-based bimetallic NPs in polyelectrolyte multilayers.44-47 In this study, we report

93

a new approach for the rapid synthesis of Ag/Au-based alloy NPs on β-LG fibrils. The synthesis of

94

metal alloys was carried out in one pot by the adsorption of metal ions (Cu2+, Ag+, AuCl4-) and in

95

situ reduction (without using any additional reducing agent), in which Cu0 was preferentially

96

formed and facilitated the synthesis of Ag0, that in turn promoted the production of Au0, leading to

97

the formation of Cu-Ag and Cu-Ag-Au alloy NP-loaded β-LG fibrils (denoted as Cu-Ag@β-LGF

5

98

and Cu-Ag-Au@β-LGF, respectively). Furthermore, a catalytic membrane reactor, in which a

99

catalytic layer of alloy NP-loaded β-LG fibrils was attached to nylon membrane matrix, was

100

developed for a continuous-flow system. Morphological and structural characterizations were

101

performed to confirm the successful formation of the alloy NPs on the β-LG fibrils using several

102

microscopy and spectroscopy techniques. The catalytic performance of the membrane reactor was

103

evaluated using the reduction of 4-nitrophenol (4-NP) into 4-aminophenol (4-AP) as a model

104

reaction.48-50 The flow rate dependence of the reactant solution, operating stability, and antibacterial

105

activity of the membrane reactor were also investigated.

106

 EXPERIMENTAL SECTION

107

Materials. β-lactoglobulin (>90%, β-LG) and Thioflavin T (>98%, ThT) were purchased from

108

Sigma-Aldrich (Beijing, China). Hydrochloroauric acid (HAuCl4), 4-nitrophenol (99%, 4-NP) and

109

sodium borohydride (NaBH4) were obtained from Aladdin Reagent Company (Shanghai, China).

110

Sodium hydroxide, copper sulfate, silver nitrate, and other chemicals were analytical grade and

111

obtained from commercial sources. Nylon membranes with an average pore size of 0.22 μm and the

112

ultrafiltration centrifuge tubes with a molecular weight cut off of 3,000 Da were obtained from

113

Millipore Corp. (Bedford, MA, USA).

114

Preparation of the β-LG Fibrils. To purify β-LG, 50 mg of the β-LG powder was dissolved in 1

115

mL of ultrapure water, and the pH of the resulting solution was adjusted to 9.0 with 1 M NaOH.

116

After stirring at 4 °C for 1 h, the precipitate containing the insoluble protein was removed by

117

centrifugation (15,000 rpm, 15 min). The supernatant was then washed three times with 0.1 M

118

NaOH solution in a 3,000 Da ultrafiltration centrifuge tube (4 oC, 8,000 rpm, 15 min) to remove the

119

metal ions. The final concentration of β-LG was adjusted to 40 mg mL-1, as determined by the 6

120

absorption of the solution at 278 nm using the specific molar absorption coefficient 0.8272 L g-1

121

cm-1. In a typical experiment, the pH of the purified β-LG solution was adjusted to 12 with 1 M

122

NaOH and then incubated at 90 °C for 20 h to yield the β-LG fibrils.

123

Synthesis of the Metal NP-Loaded β-LG Fibrils. In a typical experiment, 1.8 mL of CuSO4

124

solution (20 mM), 2.4 mL of AgNO3 (10 mM), and 500 μL of the β-LG fibril solution were added to

125

40.3 mL of ultrapure water. The pH of the resulting solution was adjusted to 12 with 2 M NaOH.

126

After being purged with N2 for 10 min, the solution was incubated in a 90 °C water bath for 5 min.

127

The formed Cu-Ag@β-LGF was collected and washed five times with ultrapure water in a

128

ultrafiltration centrifuge tube (4 °C, 8,000 rpm, 15 min).

129

Similarly, Cu-Ag-Au@β-LGF was also synthesized starting from a 45 mL precursor solution,

130

which was prepared by adding 1.8 mL of a CuSO4 solution (20 mM), 2.4 mL of AgNO3 (10 mM),

131

1.1 mL (or 2.5 mL) of HAuCl4 (20 mM), and 500 μL of the β-LGF solution into 39.2 mL (or 37.8

132

mL) of ultrapure water. The other conditions were the same as those used for the synthesis of

133

Cu-Ag@β-LGF.

134

Fabrication of the NP-Loaded Membrane Reactors. A nylon membrane (d=1.3 cm, area=0.78

135

cm2) with an average pore size of 0.22 μm was used as the membrane matrix for the construction of

136

the membrane reactor (Figure S1). Two mL of the metal NP-loaded β-LG fibril (denoted as

137

NPs@β-LGF) solution was filtered through the membrane at a flow rate of 200 μL min-1. The flow

138

rate was controlled by a syringe pump (LSP01-2A, Longer Precision Pump Co., China) equipped

139

with a 10 mL syringe. Subsequently, 7 mL of ultrapure water was flowed through the membrane at a

140

flow rate of 300 μL min-1 to remove the free molecules. The NP-loaded membrane reactor was

141

dried overnight at room temperature for further use.

7

142

Assessment of the Catalytic Performance of the NP-Loaded Membranes. The reactant

143

solution was prepared by adding 750 μL of the 4-NP stock solution (3 mM) into 7 mL of ultrapure

144

water. N2 gas was then purged through the solution for 10 min to remove the dissolved O2.

145

Subsequently, 1 mL of a freshly prepared NaBH4 solution (0.3 M) was injected under stirring. Five

146

mL of the mixture was passed through the membrane reactor at 300 K and various flow rates

147

(60-540 μL min-1) using a syringe pump (LSP01-2A, Longer Precision Pump Co., China). The

148

effluent was continuously collected, and the concentrations of 4-NP in the effluent were determined

149

using UV-vis absorption spectroscopy. The amount of product (4-AP) in the effluent was also

150

analyzed with an Agilent 1200 HPLC system using an Agilent Eclipse XDB-C18 column (150

151

mm×4.6 mm, 5 μm particle size). Each reaction was performed five times using five freshly

152

prepared membranes.

153

Given that the concentration of NaBH4 significantly exceeds that of 4-NP in the reaction system,

154

the catalytic kinetics can be considered as pseudo-first-order, with respect to 4-NP. In these cases,

155

the residence time (τ) and the kinetics of the reaction were defined by

156

τ=πr2Lε/V

(1)

157

ln(1-x)=-kobsτ

(2)

158 159 160

where r, L, ε, V, x, and kobs are the membrane radius, thickness, membrane porosity, flow rate, conversion rate, and the observed rate constant, respectively. For comparison, the control experiments were also performed in a batch system using the

161

NPs@β-LGF composites as catalysts. The reactant solution was prepared as described above and

162

190 μL of the reaction solution was added to 96-well plates. Then, 10 μL of a dispersion of the

163

NPs@β-LGF composites were added to initiate the reaction. In these cases, the conversion of 4-NP

8

164

is given by

165 166 167

–rt = –dCt/dt = kCt

(3)

where -rt is the conversion rate of 4-NP at time t, Ct is the concentration of 4-NP at time t, and k is the first-order rate constant.

168

Antimicrobial Assays. A colony of ampicillin-resistant Escherichia coli (E. coli) (No. ATCC

169

43886) was chosen as a model bacterium. The E. coli were cultured on solid LB (lysogeny broth)

170

medium for 24 h at 37 °C. Subsequently, a fraction of the E. coli solution was diluted with 1 mL of

171

water and 10 μL of a dispersion of β-LGF or NPs@β-LGF were added. The mixture was then

172

incubated at 37 °C for 24 h. The OD600 values were used to represent the concentrations of E. coli.

173

Characterizations. The morphology of the metal NPs, β-LG fibrils, and membranes were

174

observed with a field emission scanning electron microscope (FESEM, S-4800, Hitachi

175

High-Technologies CO., Japan) operated at 3 keV and a high resolution transmission electron

176

microscope (HRTEM, Tecnai G2 F20, FEI, Holland) equipped with an EDS analyzer at an

177

acceleration voltage of 200 kV. The UV-vis spectra were monitored on a TU-1810

178

spectrophotometer (Persee Instruments Ltd., China) at wavelengths ranging from 200-600 nm.

179

Fourier transform infrared spectroscopy (FTIR) spectra of β-LGFs and NP-loaded β-LGFs were

180

recorded on a Nicolet-560 FTIR spectrometer (Nicolet Co., USA) with a KBr pellet method across

181

the range of 400−4000 cm−1. The fluorescence emission spectra and 3D-fluorescence spectra were

182

recorded on a fluorescence spectrophotometer (Varian Cary Eclipse, Agilent, USA). The ThT

183

fluorescence analysis was performed to investigate the kinetics of fibrils formation with an

184

excitation wavelength of 450 nm and emission wavelength of 480 nm. The Fourier-transform

185

infrared spectroscopy (FTIR) spectra were recorded on a Nicolet-560 FTIR spectrometer (Nicolet

9

186

Co., USA) from 400−4000 cm−1 using a KBr pellet method. A total of 16 scans were accumulated,

187

with a resolution of 4 cm−1 for each spectrum. The crystal structure of NPs@β-LGF was analyzed

188

by X-ray diffraction (XRD, D/max 2500, Rigaku, Japan) with a CuKα radiation source (λ=0.154056

189

nm). The data were analyzed using the Jade 5 software package, which included a JCPDS powder

190

diffraction database. The elemental analysis was performed using X-ray photoelectron spectroscopy

191

(XPS, Axis Ultra DLD, Kratos Analytical, UK) with a monochromatic Al Kα source radiation at

192

1486.69 eV and an acquisition time of 0.15 s. The amount of Cu, Ag, and Au loaded in the

193

membranes was measured by inductively coupled plasma mass spectrometry (ICP-MS, IRIS

194

Advantage, Thermo, USA) after the membranes were dissolved in nitrohydrochloric acid at 60 °C.

195

 RESULTS AND DISCUSSION

196

Scheme 1 illustrates the strategy for the construction of the metal NP-loaded catalytic membrane

197

reactor. In the proposed method, the β-LG fibrils were synthesized via the self-aggregation of the

198

β-LG molecules. Subsequently, the dissoluble metal salts were added to the β-LGF solutions for the

199

in situ growth of the metal NPs. The NPs@β-LGFs were then deposited onto a microfiltration

200

membrane, and thus a catalytic membrane reactor was constructed for the continuous-flow reaction

201

(Figure S1).

202

203

204

205

10

206

Scheme 1. Schematic illustration of the construction of the metal NP-loaded catalytic

207

membrane reactor.

208 209

Preparation and Characterization of the Alloy Nanocatalysts. To prepare the β-LG fibrils, the

210

β-LG solution was incubated at 90 °C for 20 h. As shown in Figure 1a, uniform β-LGFs measuring

211

approximately 15 nm in diameter and at least 5 μm in length were obtained in this study. After the

212

addition of CuSO4 into the β-LGF dispersion and incubation at 90 °C for 4 h, the color of the

213

solutions visibly changed from clear to light yellow, suggesting that Cu0 had successfully formed on

214

the β-LGFs (denoted as Cu@β-LGF, Figure S2). The HRTEM image shows that the Cu

215

nanoclusters have an approximate diameter of 2.5 nm and a lattice plane distance (LPD) of 2.08 Å,

216

which is indicative of the Cu (111) plane (Figure S3a). We further investigated the optical

217

characteristics of Cu@β-LGF using UV-vis and photoluminescence spectroscopy. As shown in

218

Figure 2a, an aqueous dispersion of Cu@β-LGF has an absorption peak at approximately 330 nm,

219

without any accompanying SPR peak near 570 nm (the λmax of Cu), which is the characteristic

220

absorption peak of Cu nanoparticles (Figure 2a). Meanwhile, the Cu@β-LGF dispersion exhibits

221

purple luminescence (λem=405 nm) under an excitation wavelength of 330 nm. These results

222

demonstrated that β-LGF has the capability to reduce Cu2+ into Cu0 nanoclusters. 11

223 224

Figure 1. a-b) TEM images of (a) β-LGF and (b) Cu-Ag-Au1@β-LGF. c-d) HRTEM images of (c)

225

the Cu-Ag alloy nanoparticles and (d) Cu-Ag-Au1 alloy nanoparticles.

226

We further attempted to synthesize the Ag NPs with the assistance of β-LGFs; however, the yield

227

of the Ag NPs was low, even after a 4 h incubation, as shown by the weak SPR peaks in the UV-vis

228

spectra (Figure 2b, blue and black curves). This may be attributed to the weak binding and/or

229

reducing ability of the amino acid residues for Ag+ ions, leading to the slow nucleation and growth

230

of Ag0 crystals, particularly in the absence of chemical reductants like NaBH4 and sodium citrate.

231

Compared with Ag+, Cu2+ has a stronger affinity for the amino acid residues and can be more easily

232

reduced to Cu0 nanoclusters, although it has a lower standard electrode potential (Eo(Cu2+/Cu)

233

=0.34V vs (Eo(Ag+/Ag)=0.79V). Therefore, we speculate that the addition of Cu2+ is able to

234

promote the growth of the Ag NPs due to the reducing effect of Cu nanoclusters. As expected, when 12

235

Cu2+ and Ag+ were added into the dispersion of β-LGFs, the color of the mixture changed from

236

clear to light brown within approximately 5 min (Figure S2). The corresponding UV-vis spectrum

237

exhibits a strong SPR peak at approximately 410 nm (Figure 2b, olive curve), suggesting the that

238

Cu-Ag alloy NPs had formed. When the incubation time was increased to 4 h, a small red shift in

239

the SPR peak from 410 to 412 nm was observed and the intensity was also slightly increased

240

(Figure 2b, orange curve). The results indicate that the growth of the Cu-Ag NPs proceeds quickly

241

in the initial stage and is nearly complete at 5 min. Additionally, the HRTEM image shows that the

242

LPD values range from 2.11-2.22 Å (Figure 2c), likely corresponding to the Cu-Ag alloy (111)

243

plane.

244 245

Figure 2. UV-vis absorption spectra of aqueous suspensions of (a) β-LGF and Cu@β-LGF, (b)

246

Ag@β-LGF and Cu-Ag@β-LGF, and (c-d) Cu-Ag-Au@β-LGF. The blue curve in (a) is the

247

fluorescence emission spectrum of an aqueous suspension of Cu@β-LGF.

13

248

Using the same procedure, the Cu-Ag-Au alloy NPs were synthesized with the assistance of

249

β-LGFs and the addition of Cu2+, Ag+ and AuCl4- ions. As shown in Figure 1b, the NPs grew on the

250

surface of β-LGF, which provides evidence to support the hypothesis that β-LGF acts as stabilizing

251

agent and prevents the ions from aggregating. In this study, two alloy NPs were prepared by varying

252

the molar ratio of the Ag and Au salts, one of which exhibited a single SPR peak at 465 nm and the

253

other exhibited a peak at 512 nm (denoted as Cu-Ag-Au1@β-LGF and Cu-Ag-Au2@β-LGF,

254

respectively, Figure 2c-d). After a 5 min incubation, the intensities of the SPR peaks were

255

approximately 0.5, indicating that the rate of Cu-Ag-Au alloy NP growth is also very high in the

256

initial stage. It is worth noting that the growth of the Cu-Au alloy NPs is slow in the absence of Ag+

257

(Figure 2d, red curve). We expect that the Ag0, which is formed by the reduction of Ag+ by Cu0, can

258

serve as a reducing agent and thus promotes the formation of Au0 (Eo(AuCl4-/Au )= 1.002 V).

259

Figure 1d and Figure S3b show the HRTEM images of Cu-Ag-Au1@β-LGF and

260

Cu-Ag-Au2@β-LGF, respectively. The LPD values for the Cu-Ag-Au alloy NPs ranged from

261

2.2-2.3 Å, which is different from the classical values for the Cu (111) (2.08 Å), Ag (111) or Au

262

(111) planes (2.35 Å). Additionally, an LPD of 1.3 Å was also observed in a HRTEM image (Figure

263

1d), indicating the presence of the Cu (220) plane. As shown in Figure 1d, a notable icosahedron

264

containing five twin structures was observed, which is one of the most popular planar defects in

265

face center cubic (FCC) structures and results from the change in the volume of the Cu phases

266

during the formation of the alloy NPs. In addition to the alloy NPs, the HRTEM images revealed the

267

presence of Cu nanoclusters, with a size of less than 3 nm, while the Cu-Ag-Au alloy NPs have a

268

size distribution of 3-13 nm (Figure S4).

14

269 270

Figure 3. The fitted core-level XPS spectra of (a) Cu 2p, (b) Ag 3d and (c) Au 4f in the

271

NPs@β-LGF.

272

To confirm that the alloy formed, the XPS spectra of Cu@β-LGF, Cu-Ag@β-LGF, and

273

Cu-Ag-Au@β-LGF were obtained (Figure 3). In the Cu 2p spectra, two peaks were observed at

274

approximately 931.6 eV and 951.5 eV, which correspond to the Cu 2p3/2 band and Cu 2p1/2 band,

275

respectively, and indicate of the presence of Cu0 in all the samples. The Ag 3d core-level spectra

276

show peaks at 367.9 and 373.2 eV, which is in good agreement with the values of Ag0, suggesting

277

that there was no Ag+ (for example, in the form of Ag2O) on the particles’ surfaces. In the Au 4f

278

spectrum, two peaks centered at binding energies of 84.1 eV (Au 4f7/2) and 87.5 eV (Au 4f5/2)

279

appeared, confirming the presence of Au0. Additionally, the TEM elemental line scan mapping

280

(Figure S5) also demonstrated that Cu, Ag and Cu, Ag, Au were homogeneously distributed in the

281

alloy NPs. The results of elemental analysis (Figure S5-S6) indicate the content of Ag in

282

Cu-Ag-Au@β-LGF decreased with increasing incubation time from 5 min to 4h due to its

283

consumption in the reduction of AuCl4-. Additionally, it is worth noting that few Ag or Au

284

nanoparticles may be formed when the metal ions were bound to the free space of fibrils surface.

15

285 286

287

Figure 4. The XRD patterns of the Cu-Ag@β-LGF and Cu-Ag-Au1@β-LGF composites.

The crystal structure of the NPs@β-LGF was further characterized by XRD. As shown in Figure

288

4, Bragg reflections of the Cu-Ag phase are clearly observed at 2θ=38.8°, 44.4°, 64.6°, and 77.4°,

289

which are between the values of the individual Cu crystal (JCPDS No. 65-9026) and Ag crystal

290

(JCPDS No. 65-2871) and correspond to the (111), (200), (220), and (311) planes of the Cu-Ag

291

crystals, respectively. For the Cu-Ag-Au alloy NPs, Bragg reflection peaks appeared at 2θ=38.7°

292

and 45.4°, which are between the values of the Ag-Au alloy (JCPDS No. 65-8424, 2θ=38.2° and

293

44.4°) and Ag-Cu alloy (JCPDS No. 65-8608, 2θ=40.3° and 46.9°) and correspond to the (111) and

294

(200) planes of the ternary alloy. The XRD results are consistent with the LPD values from the

295

TEM analysis.

296

Mechanistic Analysis of the Formation of NPs@β-LGF. According to a previous study, β-LG

297

contains two tryptophan (Trp) and four tyrosine (Tyr) residues and has an emission wavelength (λem)

298

at 332 nm following excitation at 280 nm, while the λem of the β-LG fibrils formed in alkaline

299

conditions was 350 nm (Figure 5a). This distinct red shift should be attributed to a change in the

300

tertiary structure of β-LG, which inhibits the non-radiative energy transfer between Tyr and Trp51.

301

The results indicate that the Trp and Tyr residues in the β-LGF are exposed to water from the 16

302

hydrophobic core, and thus are easily able to capture the metal ions. Moreover, the 3D-fluorescence

303

(λex=280 nm, λem=350 nm) of β-LGF was significantly quenched after the addition of the Cu2+ ion

304

(800 μM) (Figure 5a-b), which provides powerful evidence in support of the hypothesis that β-LGF

305

chelated Cu2+. Moreover, the FTIR spectra (Figure S7) show that the metal NP-loaded β-LGFs have

306 307

an increase in amide II absorption band at ∼1560 cm−1 (the N−H bending vibration and C−N

308

from C−O and C−N stretching vibrations, respectively, also appeared in the spectra of NP-loaded

309

β-LGFs. These results suggested that metal NPs probably be bound to the amino acid residues

310

containing N−H, C−O, and/or C−N groups, such as Trp and Tyr. Such observation is consistent with

311

that in the fluorescence analysis as described before.

stretching vibration) compared to β-LGFs. Two new peaks at 1110 cm-1 and 1190 cm-1, resulted

312 313

Figure 5. 3-D fluorescence spectra of β-LGF (a) before and (b) after the addition of 800 μM of Cu2+.

314

(c) Plots of lg((F0-F)/(F-Fsat)) versus lg([M]) for the fluorescence response to the different metal

17

315

316

ions. (d) The change in ThT fluorescence intensity proceeds with incubation at 90 °C.

We calculated the binding constant (Kb) and the numbers of binding sites (n) by monitoring the

317

fluorescence quenching of β-LGF at 350 nm in the presence of metal ions (Figure 5c). The

318

parameters were calculated using Eq (4) 27.

319

lg((F0-F)/(F-Fsat))=nlg[M]+nlgKb

(4)

320

where F0 and F represent the Tyr fluorescence of β-LGF before and after the addition of metal

321

ions, respectively; Fsat represents the Tyr fluorescence of β-LGF saturated with metal ions, and [M]

322

is the concentration of the metal ions.

323

As shown in Table 1, the Kb of Cu2+ to β-LGF is 3 orders of magnitude higher than that of Ag+.

324

Therefore, the growth of the Ag NPs in the absence of Cu2+ is very slow, as discussed above.

325

Additionally, the value of n for Cu2+ is also larger than that for Ag+. The high binding affinity and

326

binding sites for Cu2+ ions allow them to preferentially bind β-LGF and be further reduced to Cu0

327

nanoclusters. The AuCl4- ions have a high electrode potential; however, the low concentration of

328

AuCl4- ions around β-LGF derived from the electrostatic repulsion results in the slow growth rate of

329

Au0 in the absence of the Cu2+ and Ag+ ions.

330

Table 1. Numbers of binding sites (n), binding constant (Kb) of β-LGF with different metal ions,

331

and their standard potential. Metal ions

n

Kb

Standard potential

Cu2+

2.53

1487.3

Eo(Cu2+/Cu) = 0.34 V

Ag+

0.688

1.28

Eo(Ag+/Ag) = 0.79 V

AuCl4-

1.25

3565.8

Eo(AuCl4-/Au )= 1.002 V

332

18

333

According to previous results, a proposed mechanism is shown in Figure S8. At pH 12, which is

334

much higher than the pI value (~5.2) of β-LG, a large number of amino acid residues in β-LGF

335

possesses negative static charge. Upon the addition of the metal ions, the Cu2+ ions were

336

preferentially bound to β-LGF and the Cu0 cluster was formed due to the high binding affinity.

337

Subsequently, Cu0 can reduce the Ag+ ions into Ag0 on the surface of the Cu0 cluster, while the Ag0

338

can more easily reduce the AuCl4- into Au0 in the presence of Cl- ions compared to Cu0 due to the

339

bigger potential difference (Eo(AgCl/Ag)=0.22V, Eo(Cu2+/Cu)=0.34V, Eo(AuCl4-/Au)=1.002V),

340

leading to a new kind of Cu-Ag-Au ternary nucleus. The reaction was shown below. Additionally,

341

the addition of Ag+ ions probably increases the concentration of AuCl4- ions around β-LGF because

342

of the screening of negative charge. Therefore, the high reducing ability of Ag in the presence of Cl-

343

ions and the increased AuCl4- concentration around β-LGF facilitate the synthesis of

344

Cu-Ag-Au@β-LGF.

345

β-LGFre+ Cu2+

346

Cu0+2Ag+

347

3Ag0+AuCl4-

348

β-LGFox+ Cu0

Cu2++2Ag0,

(5)

ΔEo=0.45

3AgCl+Cl-+Au0,

ΔEo=0.78

(7)

This Cu0-promoted synthesis of the alloy NPs was also demonstrated in other protein systems. As

349

shown in Figure S9, the yield of the Cu-Ag NPs is much higher than that of the Ag NPs in the

350

absence of Cu2+ ions in the different protein solutions, including lysozyme, BSA, insulin, and

351

ovalbumin.

352

(6)

In addition, the results of the ThT fluorescence assay indicate that the metal NPs can also

353

promote the growth of β-LGFs (Figure 5d). In the control experiment, there was no obvious change

354

in the fluorescence intensity of β-LGFs as the incubation time increased. However, the fluorescence 19

355

intensity of the Cu-Ag NPs was significantly increased, suggesting that the growth of β-LGFs

356

occurred by aggregation of the β-LG molecules. Similarly, the other NPs can also promote this type

357

of fibrillation (Figure 5d).

358

Catalytic Reduction of 4-NP in the Membrane Reactor. Figure 6 shows the SEM images of

359

the nylon membrane matrix and the catalytic membranes. After the attachment of the NPs@β-LGF

360

catalytic layer with a thickness of ~8 μm (Figure S10), the pore size of the membrane surface

361

decreased from ~0.22 μm to 0.04-0.12 μm. The contents of the metal elements in the catalytic

362

membranes were determined by ICP-MS. As summarized in Table 2, the Cu-Ag@β-LGF membrane

363

contains 45.31 μg mg-1 Cu and 46.16 μg mg-1 Ag, while the Cu-Ag-Au membranes are mainly

364

composed of Cu and Au elements.

365

Table 2 The contents of the Cu, Ag and Au elements (μg mg-1) in the Cu-Ag@β-LGF, Cu-Ag-Au1@

366

β-LGF and Cu-Ag-Au2@β-LGF membranes, respectively. Cu-Ag@β-LGF

Cu-Ag-Au1@β-LGF

Cu-Ag-Au2@β-LGF

Elements

Membrane

Membrane

Membrane

Cu

45.31

34.7

33.23

Ag

46.16

0.48

0.51

Au

-

112.66

200.74

20

367 368

Figure 6. SEM images of the (a) nylon membrane matrix, (b) Cu-Ag@β-LGF membrane, (c) the

369

Cu-Ag-Au1@β-LGF membrane, and (d) the Cu-Ag-Au2@β-LGF membrane.

370

4-NP is a common pollutant that is generated from the production of pesticides, herbicides,

371

insecticides, etc. Nevertheless, its corresponding derivative, 4-AP, is in great demand in a large

372

number of industrial applications. In this study, the reduction of 4-NP to 4-AP in the presence of

373

NaBH4 was selected as a target reaction to quantitatively evaluate the catalytic performance of the

374

NPs@β-LGF composites and membrane reactors. As shown in Figure 7a, the intensity of the

375

UV-vis absorption peak at 400 nm quickly decreased due to the consumption of the substrate (4-NP)

376

during the reaction process. Figure 7b shows the plots of ln(C0/Ct) versus the reaction time for the

377

reduction of 4-NP. The C0/Ct values were obtained from the relative intensity of the absorption at

378

400 nm, because it is proportional to the concentration of 4-NP. The linear relationship between 21

379

ln(C0/Ct) and reaction time (t) confirms the pseudo-first-order kinetics. The rate constants (k) of the

380

catalytic reaction were 0.29 min−1, 1.56 min−1, and 2.30 min−1 for the Cu-Ag@β-LGF,

381

Cu-Ag-Au1@β-LGF, Cu-Ag-Au2@β-LGF composites, respectively. The catalytic activity is much

382

higher than that of Ag NPs (0.175 min-1) and Au NPs (0.89 min-1) alone, as demonstrated in our

383

previous studies21, 48.

384

We investigated the effect of flow rates on the conversion of 4-NP in the catalytic membrane

385

reactors. As shown in Figure 8, with increasing flow rate, the conversion of 4-NP decreased, due to

386

the decreased residence time. We further test the Cu-Ag-Au2@β-LGF catalytic membrane reactor

387

with the lake water (from a lake located in Tianjin University) containing 4-NP (note: 4-NP was

388

added into lake water in lab). At a flow rate of 60 μL min-1, the conversion of 4-NP in lake water is

389

approximately 86.3%, which is very close to the value (~89%) in the ultrapure water (Figure 8).

390

Additionally, compared with the Cu-Ag@β-LGF CMR, the Cu-Ag-Au@β-LGF CMR has better

391

catalytic performance, indicating a significant effect of the concentration of Au on the reaction rate.

392

The enhanced catalytic performance when Au is added to the catalytic system should be attributed

393

to the high adsorption constant of 4-nitrophenol on Au surface.52 The results are also consistent with

394

those from the NPs@β-LGF composites, as mentioned above. According to Eq.(1) and (2), the

395

observed rate constants (kobs) of Cu-Ag@β-LGF, Cu-Ag-Au1@β-LGF and Cu-Ag-Au2@β-LGF

396

CMRs were calculated as 9.33 min-1, 17.1 min-1, and 20.83 min-1, respectively, as shown in Figure

397

7c, which are significantly higher than that of the NPs@β-LGF composites in batch systems.

22

398 399

Figure 7. (a) Time-dependent UV-vis absorption at 400 nm in the presence of the NPs@β-LGF

400

composites. (b) The rate constant (k) for the conversion of 4-NP using 10 μL of the NPs@β-LGF

401

composites as catalysts. c) The observed rate constant (kobs) for the conversion of 4-NP by flowing

402

the reactant solution through the catalytic membranes.

403

23

404 405

Figure 8. The effect of flow rates on the conversion of 4-NP in the catalytic membrane reactors.

406

The operating stability of the NPs@β-LGF membrane reactor was also investigated by the

407

continuous conversion of 4-NP for 8 cycles (5 mL per cycle). The change in the conversion ratio of

408

4-NP in the continuous-flow system is shown in Figure 9. The membrane reactor retained its

409

catalytic activity over eight cycles, which is indicative of good operating stability. The cycle

410

performance of these catalytic membranes (>97% productivity after eight cycles) is comparable to

411

or even better than those of metal NPs immobilized on the other carriers, such as eggshell

412

membrane (92% productivity after eight cycles),48 and magnetite-polymer hybrid (~80%

413

productivity after nine cycles).53 After the operation of the catalytic membrane, the presence of

414

alloy NPs on the membrane was confirmed by SEM (Figure 6b-d, inset figures). In addition, the

415

fibrous structures of the NPs@β-LGF composite film were maintained. According to the ICP-MS

416

results, leakage of the alloy NPs was not observed.

24

417 418

Figure 9. The operating stability in the conversion of 4-NP using the NPs@β-LGF membrane

419

reactors at a flow rate of 180 μL/min.

420

Antimicrobial Assays. Membrane fouling from microorganisms is a ubiquitous and problematic

421

phenomenon that can reduce the lifetime of the membrane reactors. In the present study, the

422

antimicrobial activity of the NPs@β-LGF composite was evaluated. As shown in Figure 10, the

423

Cu-Ag@β-LGF, which contains a high content of Ag, exhibited higher antimicrobial activity

424

towards E. coli than the other NPs@β-LG composites. The two samples of Cu-Ag-Au@β-LGF have

425

a similar content of Ag, but the antimicrobial activity of Cu-Ag-Au1@β-LGF is higher than the

426

other composite, likely due to its higher content of Ag on the outer surface of the NPs.

25

427 428

Figure 10. The content of E. coli in the presence of β-LGF or the different NPs@β-LGF composites

429

after incubation at 37°C for 24 h. The OD600 value was used to represent the concentrations of E.

430

coli. The E. coli content in the β-LGF membrane (used as a control) was normalized to 100%.

431

In this work, we have successfully developed a facile and one-step synthesis of Cu-Ag and

432

Cu-Ag-Au alloy nanoparticles using β-LGFs as reducing and stabilizing agents. The addition of

433

Cu2+ ions into a dispersion of β-LGFs produced Cu0 nanoclusters on the fibrils’ surfaces, which

434

promoted the synthesis of Ag0 due to its reducing ability. Meanwhile, the formed Ag0 is also able to

435

speed up the formation of Au0, leading to the rapid synthesis of the Cu-Ag-Au alloy NPs.

436

Furthermore, a NPs@β-LGF composite film was fabricated on a nylon membrane matrix and

437

applied to a catalytic membrane reactor for the reduction of 4-NP to 4-AP. The observed rate

438

constant in the continuous-flow system is higher than that in the batch system. In the

439

continuous-flow reaction, the catalytic activity of Cu-Ag-Au alloy NP-loaded membrane reactor

440

was also higher than that of the Cu-Ag alloy NP-loaded membrane. In addition, the catalytic

441

membrane reactor exhibited good operating stability over eight cycles. During this process, it was

26

442

not necessary to recover the catalyst from the reactants/products solution, and the alloy NP-loaded

443

membrane was reusable. The results agree with earlier report by Bolisetty et al.43 on the feasibility

444

of metal-betalactoglobulin hybrids membranes for continuous flow catalysis. Additionally, the

445

NPs@β-LGF composite also possessed antibacterial activity toward E. coli, likely due to the

446

presence of Ag0 in the alloy nanoparticles. This strategy is expected to be applicable to the

447

construction of various catalytic membrane reactors. The development of this type catalytic

448

membrane reactor for continuous-flow reactions also has industrial significance.

449

 AUTHOR INFORMATION

450

Corresponding Author

451

*E-mail: [email protected] (R. S.)

452

Tel: +86 22 27407799. Fax: +86 22 27407599.

453

Notes

454

The authors declare no competing financial interest.

455

 ACKNOWLEDGMENTS

456

This work was supported by the Ministry of Science and Technology of China (No.

457

2012YQ090194), the Natural Science Foundation of China (Nos. 51473115, 31071509 and

458

21306134).

459

 ASSOCIATED CONTENT

460

Supporting Information

461

Photos of nylon membrane matrix, NPs@β-LGF loaded catalytic membrane, and continuous-flow

462

catalytic membrane reactor system, photos of the dispersions of β-LGF and NPs@β-LGF before

463

and after incubation, HRTEM images of Cu@β-LGF and Cu-Ag-Au2@β-LGF, size distribution of

27

464

Cu-Ag-Au1@β-LGF, element line scan mapping for Cu-Ag@β-LGF and Cu-Ag-Au1@β-LGF, EDX

465

spectra of Cu-Ag@β-LGF, Cu-Ag-Au1@β-LGF, and Cu-Ag-Au2@β-LGF, FTIR spectra of the

466

β-LGF, Cu@β-LGF, Cu-Ag@β-LGF, and Cu-Ag-Au1@β-LGF, the proposed mechanism for the

467

formation of Cu-Ag-Au@β-LGF, the intensity of UV-vis absorption of different NPs@protein

468

systems, the thickness of (a) Cu-Ag@β-LGF and (b) Cu-Ag-Au1@β-LGF composite film. This

469

material is available free of charge via the Internet at http://pubs.acs.org/.

470

 REFERENCES

471

(1)

Feng, Q.; Zhao, Y.; Wei, A. F.; Li, C. L.; Wei, Q. F.; Fong, H., Immobilization of catalase

472

on electrospun PVA/PA6-Cu(II) nanofibrous membrane for the development of efficient and

473

reusable enzyme membrane reactor. Environ. Sci. Technol. 2014, 48 (17), 10390-10397.

474

(2)

Lloret, L.; Eibes, G.; Moreira, M. T.; Feijoo, G.; Lema, J. M., Removal of estrogenic

475

compounds from filtered secondary wastewater effluent in a continuous enzymatic membrane

476

reactor. identification of biotransformation products. Environ. Sci. Technol. 2013, 47 (9), 4536-4543.

477

(3)

Shi, Y.; Hu, S. H.; Lou, J. Q.; Lu, P. L.; Keller, J.; Yuan, Z. G., Nitrogen removal from

478

wastewater by coupling anammox and methane-dependent denitrification in a membrane biofilm

479

reactor. Environ. Sci. Technol. 2013, 47 (20), 11577-11583.

480 481 482 483 484 485

(4)

Gallucci, F.; Fernandez, E.; Corengia, P.; Annaland, M. v. S., Recent advances on

membranes and membrane reactors for hydrogen production. Chem. Eng. Sci. 2013, 92, 40-66. (5)

Meng, L.; Tsuru, T., Microporous membrane reactors for hydrogen production. Curr. Opin.

Chem. Eng. 2015, 8, 83-88. (6)

Diban, N.; Aguayo, A. T.; Bilbao, J.; Urtiaga, A.; Ortiz, I., Membrane reactors for in situ

water removal: A review of applications. Ind Eng Chem Res 2013, 52 (31), 10342-10354.

28

486

(7)

Kertalli, E.; Ferrandez, D. M. P.; Schouten, J. C.; Nijhuis, T. A., Direct synthesis of

487

propene oxide from propene, hydrogen and oxygen in a catalytic membrane reactor. Ind Eng Chem

488

Res 2014, 53 (42), 16275-16284.

489

(8)

Nalbandian, M. J.; Greenstein, K. E.; Shuai, D. M.; Zhang, M. L.; Choa, Y. H.; Parkin, G.

490

F.; Myung, N. V.; Cwiertny, D. M., Tailored synthesis of photoactive TiO2 nanofibers and Au/TiO2

491

nanofiber composites: structure and reactivity optimization for water treatment Applications.

492

Environ. Sci. Technol. 2015, 49 (3), 1654-1663.

493

(9)

Xu, Q. L.; Lei, W. Y.; Li, X. Y.; Qi, X. Y.; Yu, J. G.; Liu, G.; Wang, J. L.; Zhang, P. Y.,

494

Efficient removal of formaldehyde by nanosized gold on well-defined CeO2 nanorods at room

495

temperature. Environ. Sci. Technol. 2014, 48 (16), 9702-9708.

496 497

(10) Bae, S.; Hanna, K., Reactivity of nanoscale zero-valent iron in unbuffered systems: effect of pH and Fe(II) dissolution. Environ. Sci. Technol. 2015, 49 (17), 10536-10543.

498

(11) Bhattachatjee, S.; Basnet, M.; Tufenkji, N.; Ghoshal, S., Effects of rhamnolipid and

499

carboxymethylcellulose coatings on reactivity of palladium-doped nanoscale zerovalent iron

500

particles. Environ. Sci. Technol. 2016, 50 (4), 1812-1820.

501

(12) Li, J. X.; Shi, Z.; Ma, B.; Zhang, P. P.; Jiang, X.; Xiao, Z. J.; Guan, X. H., Improving the

502

reactivity of zerovalent iron by taking advantage of its magnetic memory: implications for arsenite

503

removal. Environ. Sci. Technol. 2015, 49 (17), 10581-10588.

504

(13) Xu, C. H.; Zhang, B. L.; Zhu, L. J.; Lin, S.; Sun, X. P.; Jiang, Z.; Tratnyek, P. G.,

505

Sequestration of antimonite by zerovalent iron: using weak magnetic field effects to enhance

506

performance and characterize reaction mechanisms. Environ. Sci. Technol. 2016, 50 (3), 1483-1491.

507

(14) Tang, H.; Zhu, D. Q.; Li, T. L.; Kong, H. N.; Chen, W., Reductive dechlorination of

29

508

activated carbon-adsorbed trichloroethylene by zero-valent iron: carbon as electron shuttle. J.

509

Environ. Qual. 2011, 40 (6), 1878-1885.

510

(15) Cheng, J.; Zhang, M.; Wu, G.; Wang, X.; Zhou, J. H.; Cen, K. F., Photoelectrocatalytic

511

reduction of CO2 into chemicals using Pt-modified reduced graphene oxide combined with

512

Pt-modified TiO2 nanotubes. Environ. Sci. Technol. 2014, 48 (12), 7076-7084.

513

(16) Yan, Z. X.; Xu, Z. H.; Yu, J. G.; Jaroniec, M., Highly active mesoporous ferrihydrite

514

supported Pt catalyst for formaldehyde removal at room temperature. Environ. Sci. Technol. 2015,

515

49 (11), 6637-6644.

516 517

(17) Min, Y.; Akbulut, M.; Kristiansen, K.; Golan, Y.; Israelachvili, J., The role of interparticle and external forces in nanoparticle assembly. Nat. Mater. 2008, 7 (7), 527-538.

518

(18) Yang, X.; Tian, P.-F.; Zhang, C.; Deng, Y.-q.; Xu, J.; Gong, J.; Han, Y.-F., Au/carbon as

519

Fenton-like catalysts for the oxidative degradation of bisphenol A. Appl. Catal. B-Environ. 2013,

520

134–135, 145-152.

521

(19) Menegazzo, F.; Signoretto, M.; Marchese, D.; Pinna, F.; Manzoli, M., Structure-activity

522

relationships of Au/ZrO2 catalysts for 5-hydroxymethylfurfural oxidative esterification: effects of

523

zirconia sulphation on gold dispersion, position and shape. J. Catal. 2015, 326, 1-8.

524

(20) Han, W.; Deng, J. G.; Xie, S. H.; Yang, H. G.; Dai, H. X.; Au, C. T., Gold supported on iron

525

oxide nanodisk as efficient catalyst for the removal of toluene. Ind Eng Chem Res 2014, 53 (9),

526

3486-3494.

527

(21) Liang, M.; Wang, L.; Liu, X.; Qi, W.; Su, R.; Huang, R.; Yu, Y.; He, Z., Cross-linked

528

lysozyme crystal templated synthesis of Au nanoparticles as high-performance recyclable catalysts.

529

Nanotechnology 2013, 24 (24), 245601.

30

530

(22) Liang, M.; Wang, L.; Su, R.; Qi, W.; Wang, M.; Yu, Y.; He, Z., Synthesis of silver

531

nanoparticles within cross-linked lysozyme crystals as recyclable catalysts for 4-nitrophenol

532

reduction. Catal. Sci. Technol. 2013, 3 (8), 1910.

533

(23) Liu, Y.; Chen, F.; Kuznicki, S. M.; Wasylishen, R. E.; Xu, Z., A novel method to control the

534

size of silver nanoparticles formed on chabazite. J. Nanosci. Nanotechnol. 2009, 9 (4), 2768-2771.

535

(24) Seto, H.; Yoneda, T.; Morii, T.; Hoshino, Y.; Miura, Y.; Murakami, T., Membrane reactor

536

immobilized with palladium-loaded polymer nanogel for continuous-flow Suzuki coupling reaction.

537

AIChE J. 2015, 61 (2), 582-589.

538 539 540 541 542

(25) Dickerson, M. B.; Sandhage, K. H.; Naik, R. R., Protein- and peptide-directed Syntheses of inorganic materials. Chem. Rev. 2008, 108 (11), 4935-4978. (26) Luo, Z.; Zheng, K.; Xie, J., Engineering ultrasmall water-soluble gold and silver nanoclusters for biomedical applications. Chem. Commun. 2014, 50 (40), 5143-5155. (27) Zou, L.; Qi, W.; Huang, R.; Su, R.; Wang, M.; He, Z., Green synthesis of a gold

543

nanoparticle–nanocluster composite nanostructures using trypsin as linking and reducing agents.

544

ACS Sustain. Chem. Eng. 2013, 1 (11), 1398-1404.

545

(28) Wei, H.; Wang, Z.; Zhang, J.; House, S.; Gao, Y.-G.; Yang, L.; Robinson, H.; Tan, L. H.;

546

Xing, H.; Hou, C.; Robertson, I. M.; Zuo, J.-M.; Lu, Y., Time-dependent, protein-directed growth of

547

gold nanoparticles within a single crystal of lysozyme. Nat. Nanotechnol. 2011, 6 (2), 93-97.

548

(29) Riek, R., Cell biology: infectious Alzheimer's disease? Nature 2006, 444 (7118), 429-431.

549

(30) Knowles, T. P. J.; Buehler, M. J., Nanomechanics of functional and pathological amyloid

550 551

materials. Nat. Nanotechnol. 2011, 6 (8), 469-479. (31) Li, C.; Adamcik, J.; Mezzenga, R., Biodegradable nanocomposites of amyloid fibrils and

31

552

graphene with shape-memory and enzyme-sensing properties. Nat. Nanotechnol. 2012, 7 (7),

553

421-427.

554

(32) Li, C.; Born, A.-K. K.; Schweizer, T.; Zenobi-Wong, M.; Cerruti, M.; Mezzenga, R.,

555

Amyloid-hydroxyapatite bone biomimetic composites. Adv. Mater. 2014, 26 (20), 3207-3212.

556 557 558 559 560

(33) Bolisetty, S.; Mezzenga, R., Amyloid–carbon hybrid membranes for universal water purification. Nat. Nano. 2016, 11 (4), 365-371. (34) Nystroem, G.; Fernandez-Ronco, M. P.; Bolisetty, S.; Mazzotti, M.; Mezzenga, R., Amyloid templated gold aerogels. Adv. Mater. 2016, 28 (3), 472-+. (35) Adamcik, J.; Lara, C.; Usov, I.; Jeong, J. S.; Ruggeri, F. S.; Dietler, G.; Lashuel, H. A.;

561

Hamley, I. W.; Mezzenga, R., Measurement of intrinsic properties of amyloid fibrils by the peak

562

force QNM method. Nanoscale 2012, 4 (15), 4426-9.

563

(36) Bolisetty, S.; Boddupalli, C. S.; Handschin, S.; Chaitanya, K.; Adamcik, J.; Saito, Y.; Manz,

564

M. G.; Mezzenga, R., Amyloid fibrils enhance transport of metal nanoparticles in living cells and

565

induced cytotoxicity. Biomacromolecules 2014, 15 (7), 2793-2799.

566

(37) Bolisetty, S.; Vallooran, J. J.; Adamcik, J.; Handschin, S.; Gramm, F.; Mezzenga, R.,

567

Amyloid-mediated synthesis of giant, fluorescent, gold single crystals and their hybrid sandwiched

568

composites driven by liquid crystalline interactions. J. Colloid. Interface Sci. 2011, 361 (1), 90-96.

569

(38) Li, C.; Bolisetty, S.; Mezzenga, R., Hybrid nanocomposites of gold single-crystal platelets

570

and amyloid fibrils with tunable fluorescence, conductivity, and sensing properties. Adv. Mater.

571

2013, 25 (27), 3694-3700.

572 573

(39) Zhou, J.; Saha, A.; Adamcik, J.; Hu, H.; Kong, Q.; Li, C.; Mezzenga, R., Macroscopic single-crystal gold microflakes and their devices. Adv. Mater. 2015, 27 (11), 1945-50.

32

574

(40) Bolisetty, S.; Adamcik, J.; Heier, J.; Mezzenga, R., Amyloid directed synthesis of titanium

575

dioxide nanowires and their applications in hybrid photovoltaic devices. Adv. Funct. Mater. 2012,

576

22 (16), 3424-3428.

577

(41) Bolisetty, S.; Vallooran, J. J.; Adamcik, J.; Mezzenga, R., Magnetic-responsive hybrids of

578

Fe3O4 nanoparticles with β-lactoglobulin amyloid fibrils and nanoclusters. ACS Nano 2013, 7 (7),

579

6146-6155.

580

(42) Gosal, W. S.; Clark, A. H.; Pudney, P. D. A.; Ross-Murphy, S. B., Novel amyloid fibrillar

581

networks derived from a globular protein: β-lactoglobulin. Langmuir 2002, 18 (19), 7174-7181.

582

(43) Bolisetty, S.; Arcari, M.; Adamcik, J.; Mezzenga, R., Hybrid amyloid membranes for

583 584 585 586 587 588

continuous flow catalysis. Langmuir 2015, 31 (51), 13867-13873. (44) Zhang, X.; Su, Z., Polyelectrolyte-multilayer-supported Au@Ag core-shell nanoparticles with high catalytic activity. Adv. Mater. 2012, 24 (33), 4574-4577. (45) Zhang, X.; Wang, H.; Su, Z., Fabrication of Au@Ag core-shell nanoparticles using polyelectrolyte multilayers as nanoreactors. Langmuir 2012, 28 (44), 15705-15712. (46) Chu, C.; Su, Z., Facile synthesis of AuPt alloy nanoparticles in polyelectrolyte multilayers

589

with enhanced catalytic activity for reduction of 4-nitrophenol. Langmuir 2014, 30 (50),

590

15345-15350.

591 592

(47) Chu, C.; Su, Z., Recent progress in the synthesis and catalytic application of polymer-supported nanomaterials. Chin. J. Appl. Chem. 2016, 33 (4), 379-390.

593

(48) Liang, M.; Su, R.; Huang, R.; Qi, W.; Yu, Y.; Wang, L.; He, Z., Facile in situ synthesis of

594

silver nanoparticles on procyanidin-grafted eggshell membrane and their catalytic properties. ACS

595

Appl. Mater. Interfaces 2014, 6 (7), 4638-49.

33

596

(49) Saha, A.; Adamcik, J.; Bolisetty, S.; Handschin, S.; Mezzenga, R., Fibrillar networks of

597

glycyrrhizic acid for hybrid nanomaterials with catalytic features. Angew. Chem. Int. Ed. 2015, 54

598

(18), 5408-5412.

599

(50) Gao, L.; Li, R.; Sui, X.; Li, R.; Chen, C.; Chen, Q., Conversion of chicken feather waste to

600

N-doped carbon nanotubes for the catalytic reduction of 4-nitrophenol. Environ. Sci. Technol. 2014,

601

48 (17), 10191-10197.

602 603

(51) Creamer, L. K., Effect of sodium dodecyl sulfate and palmitic acid on the equilibrium unfolding of bovine beta-lactoglobulin. Biochemistry 1995, 34 (21), 7170-7176.

604

(52) Wunder, S.; Polzer, F.; Lu, Y.; Mei, Y.; Ballauff, M., Kinetic analysis of catalytic reduction

605

of 4-nitrophenol by metallic nanoparticles immobilized in spherical polyelectrolyte brushes. J. Phys.

606

Chem. C 2010, 114 (19), 8814-8820.

607

(53) Marcelo, G.; Muñoz-Bonilla, A.; Fernández-García, M., Magnetite–polypeptide hybrid

608

materials decorated with gold nanoparticles: study of their catalytic activity in 4-nitrophenol

609

reduction. J. Phys. Chem. C 2012, 116 (46), 24717-24725.

610 611

34

612

Table of Contents/Abstract art/Synopsis

613

614 615 616

35