Adsorption Behavior of Extracellular Polymeric ... - ACS Publications

Jun 27, 2016 - Two protein-like and two humic-like components were identified from EPS by EEM-PARAFAC. Adsorption of protein-like components was ...
0 downloads 5 Views 1MB Size
Subscriber access provided by Purdue University Libraries

Article

Adsorption Behavior of Extracellular Polymeric Substances on Graphene Materials Explored by Fluorescence Spectroscopy and TwoDimensional Fourier Transform Infrared Correlation Spectroscopy Bo-Mi Lee, and Jin Hur Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b01286 • Publication Date (Web): 27 Jun 2016 Downloaded from http://pubs.acs.org on June 28, 2016

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

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

Page 1 of 27

Environmental Science & Technology

1 2 3 4

Adsorption Behavior of Extracellular Polymeric Substances on Graphene

5

Materials Explored by Fluorescence Spectroscopy and Two-Dimensional Fourier

6

Transform Infrared Correlation Spectroscopy

7 8

Bo-Mi Lee and Jin Hur*

9 10 11

Department of Environment and Energy, Sejong University, Seoul, 143-747, South Korea

12 13 14 15

Revised and Re-submitted to Environmental Science & Technology, June, 2016

16 17 18 19 20 21 22 23 24 25 26

* Corresponding author: Tel. +82-2-3408-3826. E-mail: [email protected]

Fax +82-2-3408-4320.

ACS Paragon Plus Environment

Environmental Science & Technology

27 28

Abstract Adsorption isotherms of extracellular polymeric substances (EPS) on graphene oxide (GO)

29

and reduced GO (rGO) were studied using fluorescence excitation-emission matrix - parallel

30

factor analysis (EEM-PARAFAC) and two dimensional correlation spectroscopy (2D-COS)

31

combined with Fourier transform infrared spectroscopy (FTIR). Chemical reduction of GO

32

resulted in a greater extent of carbon adsorption with a higher degree of isotherm nonlinearity,

33

suggesting that heterogeneous adsorption sites were additionally created by GO reduction.

34

Two protein-like and two humic-like components were identified from EPS by EEM-

35

PARAFAC. Adsorption of protein-like components was greater than that of humic-like

36

components, and the preferential adsorption was more pronounced for GO versus rGO.

37

Adsorption of protein-like components was more governed by site-limiting mechanisms than

38

humic-like components as shown by the higher isotherm nonlinearity. 2D-COS provided

39

further information on the adsorption of secondary protein structures. Adsorption of the EPS

40

structures related to amide I and aromatic C-C bands was greater for rGO versus GO. Protein

41

structures of EPS were more favorable for adsorption in the order of α-helix → amide II → β-

42

sheet structures with increasing site limitation. Our results revealed successful applicability of

43

EEM-PARAFAC and 2D-COS in examining the adsorption behavior of heterogeneous

44

biological materials on graphene materials.

45 46 47 48

1. Introduction Graphene, a mono-layered and sp2 structured carbon material, has been used for a variety

49

of applications including electrodes, semiconductors, sensors, and hydrogen storage, owing to

50

its robust structures and excellent conductivity for heat and electricity.1-5 A recent report has 1

ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27

Environmental Science & Technology

51

projected that graphene will lead a billion-dollar industry over the next ten years.6 While a

52

number of studies focused on producing cost-effective and efficient graphene materials to

53

extend the applicability to many fields, public interest has rapidly grown over the

54

environmental concerns with respect to their fate and transport in natural and engineered

55

systems, which requires in-depth studies on related topics.4

56

Graphene oxide (GO), a common form of graphene materials, can be produced in massive

57

amounts through the so-called Hummer’s method, which renders GO to retain partially

58

defected sp2 structures and various functional groups. Often, GO is intentionally modified

59

into a reduced form to strengthen the unique physical/chemical functions, causing GO to have

60

more defected sp2 structures and less abundance of oxygen-containing functional groups.

61

Such modifications can even occur in natural systems via sulfur-containing reductants and/or

62

bacterial activities,5 leading to the existence of various forms of graphene materials in aquatic

63

environments.

64

It is known that adsorption of organic chemicals to GO materials are driven by several

65

adsorption mechanisms such as π bonding interaction, hydrophobic interaction, electrostatic

66

interaction, and hydrogen bonding, which may cause competitive adsorption among different

67

types of organic compounds.7 For example, π- π interaction is an important mechanism for

68

aromatic compounds and macromolecules, while the adsorption of oxygen-containing

69

compounds onto GO may be governed by hydrogen bonding.7 The adsorption behavior of

70

organic matter is affected by its structure as well as co-existing species. It was reported that

71

aromatic fractions in organic matter showed a higher affinity to adsorb on GO surfaces than

72

non-aromatic fraction.8 Such a preferential adsorption can occur for humic acids with

73

relatively heterogeneous characteristics with respect to molecular size. Lee et al.9 have shown

74

that large sized molecules of a humic acid exhibited a greater adsorption affinity than smaller 2

ACS Paragon Plus Environment

Environmental Science & Technology

75 76

sized molecules. Recently, the interplay between graphene materials and microbial products has received a

77

great deal of attention concerning biological responses (e.g., toxicity) of microbes after

78

contact with graphene materials and/or the production of bio-functionalized graphene.10-12

79

Many studies reported that graphene materials could absorb many types of biological

80

products.10,13-15 Ahmed and Rodrigues (2013) using a scanning electron microscope (SEM)

81

confirmed the interaction of graphene materials with activated sludge in wastewater.6 Proteins,

82

peptides, and lipids are known to adsorb graphene surfaces via hydrophobic interactions, π-π

83

interaction, electrostatic interactions, and/or hydrogen bonding.10,16 Polyanionic structures of

84

nucleic acids may create electrostatic repulsive forces between microbial products and

85

graphene surfaces.16 It was previously reported that positively charged side chains and

86

aromatic structures of amino acids may promote their adsorption onto graphene materials.10

87

The related adsorption behavior and the involved mechanisms may be affected by the types

88

of graphene materials (i.e., surface characteristics). For example, Zhang et al (2013) reported

89

that reduced forms of GO (rGO) were more effective in adsorbing model enzymes (e.g.,

90

horseradish peroxidase and oxalate oxidase) than untreated GO due to a greater contribution

91

of hydrophobic interactions.13

92

Extracellular polymeric substances (EPS), which are placed outside of microbial cells and

93

in the interior of microbial aggregates, relate to many biological properties of wastewater

94

such as settling of sludge, cell protection, and membrane fouling.17-20 EPS consists of a

95

variety of organic matter including polysaccharides, proteins, uronic acids, humic substances,

96

lipids, and DNA.21 The heterogeneous composition of EPS can make its adsorption behavior

97

much complicated. For example, Omoike and Chorover (2004, 2006) found that certain EPS

98

constituents might be preferentially adsorbed onto mineral surfaces.22,23 They observed that 3

ACS Paragon Plus Environment

Page 4 of 27

Page 5 of 27

Environmental Science & Technology

99 100 101

the infrared spectra of EPS were red-shifted with the decrease of amide Ι/amide ΙΙ peak area ratio after adsorption. Considering the environmental importance of EPS in biological treatment systems and the

102

growing demand of graphene materials for industry, it is surprising that there was no prior

103

study on the interactions between graphene materials and EPS. Although it is easily

104

postulated that the heterogeneous EPS structures make their adsorption behavior onto

105

graphene materials very complicated, the complication is possibly resolved by employing

106

advanced EPS characterization methods. Fluorescence excitation emission matrix (EEM)-

107

parallel factor analysis (PARAFAC) can separate dissimilar fluorescence components from

108

bulk dissolved organic matter (DOM), helping to trace different behaviors/responses of the

109

individual components upon a given process.24-26 Our previous study has demonstrated the

110

successful application of EEM-PARAFAC in evidencing adsorptive fractionation

111

phenomenon of humic acids on the GO surface.9 Two-dimensional correlation spectroscopy

112

(2D-COS) is another valuable mathematical tool for identifying the subtle responses of a

113

heterogeneous mixture (e.g., EPS) to changing conditions, in which the variations of the

114

correlations among different spectral variables are visualized in a two-dimensional space.27-29

115

2D-COS could also provide the extent and the order of the sequence in the variable changes

116

upon external factors such as solution pH and surface coverage by adsorption.30 Although

117

such spectroscopic characterization has some limitations, there is a high hope that applying

118

EEM-PARAFAC and 2D-COS could provide valuable information to untangle the complex

119

adsorption behavior of EPS on graphene materials and to explain the associated mechanisms,

120

which have not yet been explored.

121 122

The objectives of this study were (1) to track the compositional changes of EPS upon their adsorption onto differently reduced graphene materials, and (2) to compare the 4

ACS Paragon Plus Environment

Environmental Science & Technology

123

individual adsorption isotherm behavior of different EPS constitutes via the two advanced

124

mathematical tools - fluorescence EEM-PARAFAC and 2D-COS based on Fourier transform

125

infrared spectroscopy (FTIR).

126 127

2. Materials and Methods

128

2.1. Sludge collection and EPS extraction

129

Aerobic sludge was sampled from an aerobic tank in a municipal wastewater treatment

130

plant, located in the city of Seoul (Jungrang-gu), Korea. The facility has a treatment capacity

131

of 1,590,000 ton/day, and it is operated by two advanced biological treatment processes,

132

namely, A2/O (Anaerobic-anoxic/aerobic) and MLE (Modified Ludzack Ettinger). The

133

collected sludge was stored at 4˚C in the dark. Total suspended solids (TSS) and volatile

134

suspended solids (VSS) were measured immediately in a laboratory after sampling.

135

EPS of the collected sludge were extracted based on a modified formaldehyde/NaOH

136

method suggested in Domíguez et al.31, which is known to be beneficial for the protection of

137

cell lysis. A volume (700 mL) of sludge was centrifuged at 5000 rpm for 15 min at 4˚C

138

before the sludge pellet was re-suspended in 0.05% NaCl solution (350 mL). An aliquot (2.1

139

mL) of formaldehyde (37%) was then injected to the NaCl solution containing sludge, and

140

the mixture was stirred for 1 hour at 900 rpm at 4˚C. In the next step, 120 mL of 1N NaOH

141

was added and stirred for 3 hours under the same conditions. The extracted EPS was

142

separated by centrifugation (5000 rpm, 10 minutes) followed by filtration using a 0.2 µm

143

pre-washed cellulose acetate membrane filter (Advantec). The EPS was further purified by

144

using regenerated cellulose tubular dialysis membrane (3500 Da, Membrane Filtration

145

Products, Inc.) for 24 hours at 4 ˚C in Milli-Q water to remove inorganic species,

146

formaldehyde, and small organic molecules. A portion of the final EPS solutions was freeze5

ACS Paragon Plus Environment

Page 6 of 27

Page 7 of 27

147

Environmental Science & Technology

dried for FTIR analysis.

148 149 150

2.2. Preparation of GO and rGO GO was prepared based on a modification of Hummer’s method, generally following the

151

procedure reported in Marcano et al (2010).32 For pre-oxidation, 12 g of graphite flakes

152

(Sigma Aldrich) were mixed with 50 mL of concentrated sulfuric acid containing potassium

153

peroxydisulfate (10 g) and phosphorous pentoxide (10 g) at 80˚C. After 5 hours, the slurry

154

was added to 2 L of Milli-Q water and kept overnight in ambient temperature. The pre-

155

oxidized graphite flakes were then washed in Milli-Q water, filtered through 5 µm pore sized

156

filter paper to remove residuals, and dried overnight in ambient temperature. For the next

157

step, 10 g of the pre-oxidized graphite flakes were stirred in 230 mL of concentrated sulfuric

158

acid containing sodium nitrate (1 g) at 0˚C. Potassium permanganate (30 g) was very slowly

159

added to the mixture with the temperature maintained at under 10˚C. The mixture was

160

reacted at 35˚C for 2 hours before 2 L of Milli-Q water was added under 50˚C. It was then

161

stored overnight after adding 40 mL of hydrogen peroxide (30%). Precipitated GO was

162

finally washed with 6 L of hydrochloric acid (4 N) to remove the remaining manganese. The

163

prepared GO solution was sonicated for 30 minutes in a bath sonicator to separate un-

164

exfoliated graphene sheets. Single-sheet GO was finally obtained by centrifugation at 5000

165

rpm for 10 minutes after washing with Milli-Q water.

166

Hydrazine monohydrate (1 µL per 15 mg of GO) was added to a GO solution to induce

167

the chemical reduction of GO. The solution was then stirred at 80˚C for 2 hours (rGO-2h)

168

and 8 hours (rGO-8h) to obtain partially reduced GO. The obtained rGO was washed

169

repetitively (five times) with Milli-Q water followed by centrifugation at 8000 rpm for 15

170

minutes. The final graphene materials were characterized by using attenuated total reflection 6

ACS Paragon Plus Environment

Environmental Science & Technology

171

(ATR)-FTIR (Perkin-Elmer spectrum 100), X-ray diffraction (XRD, D/MAX-2500/PC,

172

Rigaku), Raman spectroscopy (Renisshaw 633 nm), X-ray photoelectron spectroscopy (XPS,

173

K-alpha, Thermo VG). The specific surface area (TriStar II 3020) and the zeta potentials

174

(NICOMP 380 ZLS) were also measured. All the characteristics are described in the

175

supporting information (Table S1, Figures S1 and S2).

176 177 178

2.3. Adsorption isotherm experiments The obtained EPS solution was diluted in a range of dissolved organic carbon (DOC)

179

concentrations from 10 mg C/L to 50 mg C/L for adsorption isotherm experiments at pH 5.0

180

± 0.02. The net surface charges of the GO materials were negative at pH 5.0 (Figure S1d).

181

The ionic strength was fixed at 0.1 M NaCl. Three types of graphene stock solutions (GO,

182

rGO-2h, and rGO-8h) were added to the EPS solutions to achieve the final adsorbent’s

183

concentrations of 300 mg/L, 300 mg/L, and 200 mg/L, respectively. The equilibrium time

184

was set at 24 hours based on a preliminary kinetic adsorption test (Figure S5). No change in

185

EPS (e.g., biodegradation) was confirmed for a control test without the graphene materials

186

during the equilibration (Figure S6). The residual EPS after adsorption was separated from

187

the adsorbents (i.e., graphene materials) by centrifugation at 8000 rpm followed by filtration

188

through a 0.2 µm cellulose acetate membrane (Advantec).

189 190

Adsorption parameters were estimated based on the following Freundlich model, a commonly used isotherm model: 1

Qe =k F Cen 191

where Qe (mg C/g) and Ce (mg C/L) are the adsorbed EPS amounts per graphene and the

192

EPS concentrations in solution at equilibrium, respectively. The model parameters of kF and

193

1/n refer to the model capacity factor and the isotherm nonlinearity, respectively. 7

ACS Paragon Plus Environment

Page 8 of 27

Page 9 of 27

Environmental Science & Technology

194 195 196

2.4. EPS characterization DOC concentrations of EPS samples were determined using a TOC analyzer (Shimadzu

197

V-series, TOC-V CPH). Fluorescence EEMs of EPS were measured by a luminescence

198

spectrometry (LS-55, Perkin-Elmer) with the excitation and emission wavelengths set at 220

199

nm to 500 nm with 5 nm increments and at 280 nm to 550 nm with a 0.5 nm increment,

200

respectively. ATR-FTIR (Perkin-Elmer spectrum 100) was used to determine EPS structures

201

and the distributions of the functional groups. The original EPS was fractionated into

202

hydrophobic acids (HPO-A), hydrophobic neutrals (HPO-N), transphilic acids (TPI-A),

203

transphilic neutrals (TPI-N), and hydrophilic acids (HPI) based on the resins of Supelite

204

DAX-8 (Supelco) and Amberite XAD-4 (Sigma).33 The compositions of the resin-fractioned

205

EPS were quantified by the measurements of DOC and fluorescence EEM (Table S3 and

206

Figure S4). The ATR-FTIR spectra of the freeze dried GO and rGO-8h were measured

207

before and after adsorption at a wavenumber range of 2000 to 450 cm-1.

208 209 210

2.5. EEM-PARAFAC modeling and 2D FTIR correlation spectroscopy Following Stedmon and Bro (2008) tutorial,26 EEM-PARAFAC modeling was conducted

211

for the EEM data set (n = 43) using MATLAB7.1 and the free-downloaded DOMFluor

212

toolbox (www.models.life.uk.dk). The number of the PARAFAC components was validated

213

by split-half analysis and a core consistency test. The maximum fluorescence intensities (Fmax)

214

of the identified fluorescent components were used to represent their concentrations.

215

Synchronous and asynchronous maps of 2D FT-IR correlation spectroscopy were generated

216

using 2Dshige software version 1.3 (https://sites.google.com/site/Shigemorita/home/2dshige).

217

In a synchronous map, auto peaks located at the diagonal line represent the variation of the 8

ACS Paragon Plus Environment

Environmental Science & Technology

218

changes in the spectral intensities corresponding to the locations upon an external

219

perturbation. Cross peaks, which are located off the diagonal line, refer to the simultaneous

220

changes of the spectral variables observed at two different locations (i.e., x- and y-axes).

221

Positive cross peaks indicate the two variables changing in the same direction, while negative

222

cross peaks imply the opposite trend in the changes. Meanwhile, cross peaks in an

223

asynchronous map represent the sequence of the events (or variable changes) by given

224

perturbations. If the signs of cross peaks are the same for synchronous and asynchronous

225

maps, the spectral change of one variable (e.g., x-axis) precedes that of the other (e.g., y-axis).

226

In contrast, if the signs are different, the orders of the changes are reversed. The general

227

interpretation of the event’s sequence is based on Noda’s rules.28

228 229

3. Results and Discussion

230

3.1. EPS characterization and adsorption of EPS

231

The TSS and VSS concentrtions of the sludge were 5991 and 3912 mg/L, respectively

232

(Table S2). The EPS yield per VSS was 53.2 mg/g (Table S2), which was lower than the

233

ranges previously reported (79-129 mg/g VSS).31,34 The ATR-FTIR showed the presence of

234

proteins (amide I and amide II bands), polysaccharides, and nucleic acids as the major

235

structures of EPS, which was based on the assignments following Omoike and Chorover 22

236

(Figure 1a). Meanwhile, the most abundant resin fraction of EPS was HPI, followed by HPO-

237

N → HPO-A → TPI-A → TPI-N (Figure 1b).

238

Irrespective of graphene type, all adsorption behaviors of EPS were well described by the

239

Freundlich model (R2 > 0.94), and the corresponding parameters are presented in Table 1.

240

Overall, an increasing and a decreasing trend were observed for the KF (i.e., adsorption

241

capacities) and the 1/n values (i.e., isotherm nonlinearity), respectively, on the order of GO, 9

ACS Paragon Plus Environment

Page 10 of 27

Page 11 of 27

Environmental Science & Technology

242

rGO-2h, and rGO-8h. The chemical reduction of GO seems to create a greater number of

243

adsorption sites available to EPS as shown by the increased specific surface area (Table S1).

244

The higher isotherm nonlinearity suggests that the adsorbing EPS molecules could compete

245

more strongly for the additional sites and/or for the modified rGO surface. The higher

246

adsorption capacity of rGO versus GO agreed well with a recent report of Yan et al. 35, in

247

which oxidized GO exhibited a lower extent of adsorption for aromatic organic compounds

248

(i.e., aniline, nitrobenzene, and chlorobenzene). In addition, oxidizing GO surface (i.e., more

249

oxygen presence) can make the adsorption of water molecules more favorable than EPS

250

adsorption and/or it may localize the π electrons, reducing the adsorption of EPS through π-π

251

interaction. 36 Considering that a lower 1/n value typically reflects more heterogeneous

252

surfaces with respect to adsorption sites,37 the higher nonlinear isotherms for rGO versus GO

253

observed for this study may be associated with a greater degree of the defected sp2 structures

254

(Figure S1c) and the reduced presence of C-OH (or more equal abundance of C-OH to those

255

of other carbon components) for rGO (Figures S1a and S2). This is because the defected sp2

256

structures and the equal abundances of different oxygen-containing carbon components can

257

represent more heterogeneous properties of rGO with respect to its carbon structure and

258

functional groups, respectively.

259 260 261

3.2. PARAFAC components and their adsorption behaviors Four different fluorescent components, as presented in Figure 2, were identified based on

262

a core consistency test (80.7%) and split-half validation. Component 1 (C1), assigned as

263

microbial humic-like fluorophores, had two peaks at the excitation/emission wavelengths

264

(Ex/Em) of 220 nm/430 nm and 310 nm/430 nm. Component 2 (C2), which can be assigned

265

to an aromatic protein-like component, exhibited two peaks at 220 nm/355 nm and 280 10

ACS Paragon Plus Environment

Environmental Science & Technology

266

nm/355 nm (Ex/Em). Component 3 (C3), denoted as a humic-like component, presented its

267

peaks at longer wavelengths (Em) than other components (Ex/Em = 250 nm/460 nm, 360

268

nm/460 nm). Component 4 (C4) appears to be associated with protein-like substances. All the

269

assignments of the PARAFAC components were based on the EEM peak locations

270

previously reported for EPS.38,39

271

For this study, compositional changes in EPS upon adsorption were tracked by using two

272

PARAFAC ratios (i.e., C2/C1 and C4/C3) in Fmax values of the residual EPS after adsorption,

273

which were plotted against percent adsorption in Figure 3. For all the graphene materials, the

274

two ratios of the residual EPS were much lower than the original values before adsorption

275

(12.7 for C2/C1 and 3.5 for C4/C3), ranging from 0.5 to 5.0 and 0.0 to 1.5 for C2/C1 and

276

C4/C3, respectively. The lower ratios of the residual EPS after adsorption indicate

277

preferential adsorption of protein-like components (C2 and C4) onto graphene surfaces over

278

humic-like components (C1 and C3) in EPS. Such a preferential adsorption behavior was also

279

confirmed by comparing the Freundlich model parameters of different PARAFAC

280

components individually fitted to the isotherms (Table 2 and Figure S7), in which the

281

adsorption affinities (i.e., KF values) were much higher for the protein-like components than

282

for the humic-like components regardless of graphene type.

283

Interestingly, the C2/C1 and C4/C3 ratios showed a decreasing trend with increasing

284

percent of adsorption (Figure 3), suggesting that the degree of the preferential adsorption

285

became greater for the situation in which the adsorption sites are more available and thus less

286

competition is expected among different EPS molecules. The deviations of the two ratios

287

from the original values were even more pronounced for GO versus rGO as shown by the

288

lower ratios for GO at the same percent adsorption rates (Figure 3). The difference in the

289

degree of the preferential adsorption between GO and rGO may be attributed to relatively 11

ACS Paragon Plus Environment

Page 12 of 27

Page 13 of 27

Environmental Science & Technology

290

more hydrophilic nature of GO surfaces, which favors the adsorption of protein-like EPS

291

substances (e.g., C2 and C4 components for this study) rather than humic-like components.

292

Our results can also be explained by electrostatic interaction of nitrogen-containing

293

functional groups with oxygen functional group of GO. 35 Importance of electrostatic

294

interaction between proteins (i.e., amide groups) and GO surfaces is well described in a

295

review article.40 It is notable from our study that the preferential adsorption of protein-like

296

versus humic-like components was more obvious for GO, which has relatively more

297

hydrophilic surface and more acidic functional groups, than for rGO.

298

For the same graphene materials, a higher isotherm nonlinearity (or a lower 1/n) was

299

exhibited for protein-like components (C2 and C4) than for humic-like components (C1 and

300

C3) (Table 2), suggesting that the adsorption of protein-related EPS molecules may be driven

301

by site-limiting mechanisms (e.g., electrostatic interaction) to a greater extent compared to

302

humic substances. It is notable that the presence of simple aromatic amino acids is a major

303

contributor to the protein-like fluorophores, with their peaks at shorter wavelengths (Em).

304 305

3.3. Changes in the ATR-FTIR spectra of graphene materials after EPS adsorption

306

The ATR-FTIR spectra of graphene materials (GO and rGO-8h) were compared before

307

and after EPS adsorption in Figure 4. The EPS-adsorbed graphene showed additional peaks

308

of amide I (near 1640 cm-1) and amide II (near 1545 cm-1), which originated from the EPS

309

only (Figure 1 and Table S5). Although it is easily inferred that adsorption of aromatic carbon

310

structures of EPS predominantly occurred, the related peaks were not directly confirmed from

311

the ATR-FTIR spectra probably because of the overwhelming presence of sp2 structures in

312

GO. The bands at 1380 cm-1, 1236 cm-1, 1081 cm-1 and 980 cm-1 were also found for both

313

GO and rGO after EPS adsorption. However, the origins were not clear although the 12

ACS Paragon Plus Environment

Environmental Science & Technology

314

possibility cannot be ruled out that the bands of EPS could be shifted upon adsorption to the

315

GO and rGO-8h. It is notable that those bands were not observed for the FTIR spectrum of

316

the EPS only.

317

In contrast to the EPS-adsorbed GO, the amide I band of EPS-adsorbed rGO-8h was

318

shifted toward a longer wavenumber from 1635 cm-1 to 1655 cm-1 (Figure 4b), implying

319

strong binding between the corresponding EPS structures and rGO surfaces, which was also

320

reported in Xu and Jiang.41 Considering that the amide I band originates from the peptide

321

backbone C=O stretching vibration, the difference between GO and rGO indicates that the

322

interaction between graphene surfaces and peptide backbone structures can be intensified by

323

chemical reduction of GO. Enhancement of both amide II and aromatic C-C bands (1450 cm-

324

1

325

(i.e., restoration of sp2 structures) could lead to greater adsorption of some aromatic and/or

326

nitrogenous EPS molecules. This is consistent with our previous observation of a higher

327

adsorption affinity of aromatic protein-like component (i.e., C2) on rGO-8h versus GO

328

(Table 2). Our results are also supported by Chen and Chen 37, who demonstrated that the

329

adsorption sites of GO through π-π interaction could be expanded by the restoration of sp2

330

structures upon chemical reduction.

) with no shifts was also found for the EPS-adsorbed rGO-8h, suggesting that GO reduction

331 332 333

3.4. 2D correlation ATR-FTIR spectroscopy To obtain more detailed information on the structural changes of EPS upon adsorption,

334

2D-COS was conducted using a series of the FTIR spectra of graphene materials with

335

increasing EPS adsorption amounts (Figure 5 and Figure S8), which were based on the

336

previous isotherm data. From the synchronous maps, two main auto peaks were observed at

337

the amide I (1600-1700 cm-1) and amide II (1500-1560 cm-1) bands for both GO and rGO-8h 13

ACS Paragon Plus Environment

Page 14 of 27

Page 15 of 27

Environmental Science & Technology

338

(Figures 5a and 5c). The band corresponding to the aromatic C-C band (1450 cm-1) seemed to

339

overlap with the amide II band in the synchronous maps of both graphene materials. There

340

was a notable difference in the synchronous maps between GO and rGO-8h for this study.

341

For GO, the amide I band was split into two separated peaks, while only one broad peak was

342

observed at the same location for rGO-8h. In general, the amide I band of FTIR spectra can

343

be further separated into two different bands of 1650-1700 cm-1 and 1600-1630 cm-1, each of

344

which is known to be turn and α-helix structure and β-sheet structure of peptides,

345

respectively.42 No polysaccharide-related peak (e.g., 1040 cm-1) was observed in the

346

synchronous maps although carbohydrate was one major constitute of EPS (Table S5). This

347

suggests that adsorption of polysaccharide in EPS may not be significant compared to those

348

of other EPS structures.

349

The intensities of the auto peaks can represent the extent of adsorption of the

350

corresponding structures with increasing adsorption amounts. For this study, higher

351

intensities of amide I and aromatic C-C auto peaks were found for rGO-8h than for GO,

352

while those of amide II band were similar for both graphene materials (Table S6). Our results

353

confirmed the previous observation (Figure 4) that the adsorption of the EPS molecules

354

relating to amide I and aromatic C-C bands might be greater for rGO-8h versus GO.

355

For the cross peaks (i.e., the peaks located off the diagonal line) of the synchronous map

356

of GO (Figure 5a), it was found that the two separated amide I bands at x-axis/y-axis of 1672

357

cm-1 / 1615 cm-1, each of which corresponds to α-helix and β-sheet structures, had positive

358

and weak negative signs, respectively. This suggests that the adsorption behavior might be

359

different between α-helix and β-sheet structures in responding to the site limitation of GO

360

(i.e., more surface coverage). The positive sign may reflect more favorable adsorption upon

361

the site limitation. For rGO-8h, unfortunately, only a broad positive peak appeared at the 14

ACS Paragon Plus Environment

Environmental Science & Technology

362

same location of the synchronous map, which makes the interpretation of the adsorption

363

behavior of the two secondary protein structures difficult. The other cross peaks at 1380 cm-1,

364

1236 cm-1, 1081 cm-1 and 980 cm-1 along the y-axis shared the locations of the x-axis

365

corresponding to the band of either amide I or amide II (Figures 5a and 5c), suggesting that

366

the associated structures were also involved in the adsorption of protein EPS structures.

367

Notably, such a concurrent adsorption of different EPS structures could be only confirmed by

368

2D-COS, not by the original FTIR spectra.

369

Asynchronous maps are useful in explaining the sequence of variable changes (or

370

different responses to an external perturbation) when the locations and signs of cross peaks

371

were compared between synchronous and asynchronous maps.43 For the asynchronous map

372

of GO, α-helix/β-sheet and α-helix/amide II (x-axis/y-axis) peaks displayed a negative and a

373

positive signs, respectively (Figure 5b and Table S7), which were the same as the signs in the

374

synchronous map. However, β-sheet/amide II peak showed a weak positive sign, which was

375

opposite to the sign in the synchronous map (Figure 5b and Table S7). For rGO-8h, the two

376

cross peaks of α-helix/β-sheet and α-helix/amide II in the asynchronous map showed the

377

same signs as those of the synchronous maps, while β-sheet/amide II peak had the opposite

378

signs between the two maps. Interestingly, secondary protein structures (i.e., α-helix and β-

379

sheet structures) within the amide I band, which were not identified from the previous

380

synchronous map of rGO-8h, could be distinguished by the asynchronous map (Figure 5).

381

This suggests that asynchronous maps could capture subtle responses of the associated EPS

382

structures in adsorption to increasing adsorption amount (or greater site limitation).

383

Following Noda’s rule, the adsorption of such protein-related structures seems to occur

384

on the order of α-helix → amide II → β-sheet structures for both GO and rGO-8h with

385

increasing adsorption amounts. Although no direct evidence was available in the literature to 15

ACS Paragon Plus Environment

Page 16 of 27

Page 17 of 27

Environmental Science & Technology

386

support our findings, recent studies have demonstrated the possibilities of the sequential

387

adsorption of secondary protein structures onto graphene materials. For example, Katoch et al.

388

(2012) found using FTIR spectra that α-helix structures of peptides bound to graphene

389

surfaces might undergo a transition to a complex reticular form by the interaction with the

390

surface.42 Park et al. (2014) demonstrated different adsorption affinities between single- and

391

double-stranded DNA for GO surface.16

392 393

Supporting Information

394

17 pages including text, 9 figures, 7 tables, and a reference list. This material is available free

395

of charge via the Internet at http://pubs.acs.org.

396 397

Acknowledgements

398

This work was supported by a National Research Foundation of Korea (NRF) grant funded

399

by the Korean government (MSIP) (No.2014R1A2A2A09049496) and also by Korea

400

Ministry of Environment (MOE) through Waste to energy recycling Human resource

401

development Project

402 403

References

404

(1)

405 406 407 408 409 410 411 412 413

Ramesha, G.K.; Kumara, A.V.; Muralidhara, H.B.; Sampath, S. Graphene and graphene oxide as effective adsorbents toward anionic and cationic dyes. J. Colloid and Interf. Sci., 2011. 361(1), 270-277. (2) Singh, A.P.; Mishra, M.; Chandra, A.; Dhawan, S.K. Graphene oxide/ferrofluid/cement composites for electromagnetic interference shielding application. Nanotechnology, 2011. 22(465701), 1-9. (3) Wang, L.; Lee, K.; Sun, Y.Y.; Lucking, M.; Chen, Z.F.; Zhao, J.J.; Zhang, S.B.B. Graphene oxide as an ideal substrate for hydrogen storage. ACS Nano, 2009. 3(10), 29953000. (4) Wang, F.; Haftka, J.J.H.; Sinnige, T.L.; Hermens, J.L.M.; Chen, W. Adsorption of 16

ACS Paragon Plus Environment

Environmental Science & Technology

414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455

polar, nonpolar, and substituted aromatics to colloidal graphene oxide nanoparticles. Environ. Pollut., 2014. 186, 226-233. (5) Chowdhury, I.; Mansukhani, N.D.; Guiney, L.M.; Hersam, M.C.; Bouchard, D.B. Aggregation and stability of reduced graphene oxide: complex roles of divalent cations, pH, and natural organic matter. Environ. Sci. Technol., 2015. 49, 9995-10009. (6) Ahmed, F.; Rodrigues, D.F. Investigation of acute effects of graphene oxide on wastewater microbial community: A case study. J. Hazard. Mater. 2013, 256-257, 33-39. (7) Zhao, J.; Wang, Z.; White, J. C.; Xing, B. Graphene in the Aquatic Environment: Adsorption, Dispersion, Toxicity and Transformation. Environ. Sci. Technol., 2014. 48 (18), 10886-10893. (8) Cai, N.; Peak, D.; Larese-Casanova, P. Factors influencing natural organic matter sorption onto commercial graphene oxides. Chemical Engineering Journal 2015, 273, 568-579. (9) Lee, B.M.; Seo, Y.S.; Hur, J. Investigation of adsorptive fractionation of humic acid on graphene oxide using fluorescence EEM-PARAFAC. Water Res., 2015. 73, 242-251. (10) Sanchez, V.C.; Jachak, A.; Hurt, R.H.; Kane, A.B. Biological interactions of graphene-family nanomaterials- An interdisciplinary review. Chem. Res. Toxicol., 2012. 25(1), 15-34. (11) Sharma, P.; Tuteja, S.K.; Bhalla, V.; Shekhawat, G.; Dravid, V.P. Bio-functionalized graphene–graphene oxide nanocomposite based electrochemical immunosensing. Biosensors and Bioelectronics, 2013. 39, 99-105. (12) Gravagnuolo, A.M.; Morales-Narváez, E.; Longobardi,S.; Silva, E.T.D.; Giardina, P.; Merkoçi, A. In situ production of biofunctionalized few-layer defect-free microsheets of graphene. Adv. Funct. Mater., 2015. 25, 2771-2779. (13) Zhang, C.L.; Wu, L.; Cai, D.Q.; Zhang, C.Y.; Wang, N.; Zhang, J.; Wu, Z.Y. Adsorption of polycyclic aromatic hydrocarbons (fluoranthene and anthracenemethanol) by functional graphene oxide and removal by pH and temperature-sensitive coagulation. ACS appl. Mater. Inter., 2013. 5(11), 4783-4790. (14) Saifuddin, N.; Raziah, A. Z.; Junizah, A. R. Carbon nanotubes: A review on structure and their interaction with proteins. J. Chem., 2012. ID 676815, 1-18. (15) Kamiya, Y.; Yamazaki, K.; Ogino, T. Protein adsorption to graphene surfaces controlled by chemical modification of the substrate surfaces. J. Colloid Interf. Sci., 2014. 431, 77-81. (16) Park, J.S.; Goo, N.I,; Kim, D.E. Mechanism of DNA adsorption and desorption on graphene oxide. Langmuir, 2014. 30, 12587-12595. (17) Ding, Z.; Bourven, I.; Guibaud, G.; Hullebusch, E.D.V.; Panico, A.; Pirozzi, F.; Esposito, G. Role of extracellular polymeric substances (EPS) production in bioaggregation: application to wastewater treatment. Appl. Microbiol. Biot., 2015, 99(23), 9883-905. (18) Ahmed, Z.; Cho, J.; Lim, B. R.; Song, K. G.; Ahn, K. H. Effects of sludge retention time on membrane fouling and microbial community structure in a membrane bioreactor. J. Membrane Sci. 2007, 287(2), 211-218. 17

ACS Paragon Plus Environment

Page 18 of 27

Page 19 of 27

456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496

Environmental Science & Technology

(19) More, T.T.; Yadav, J.S.S.; Yan, S.; Tyagi, R.D.; Surampalli, R.Y. Extracellular polymeric substances of bacteria and their potential environmental applications. J. Environ. Manage., 2014. 144, 1-25. (20) Herzberg, M.; Kang, S.; Elimelech, M. Role of extracellular polymeric substances (EPS) in biofouling of reverse osmosis membranes. Environ. Sci. Technol, 2009. 43, 4393-4398. (21) Sheng, G.P.; Yu, H.Q.; Li, X.Y. Extracellular polymeric substances (EPS) of microbial aggregates in biological wastewater treatment system: A review, Biotechnol. Adv., 2010. 28, 882-894. (22) Omoike, A.; Chorover, J. Spectroscopic study of extracellular polymeric substances from Bacillus subtilis: Aqueous chemistry and adsorption effects. Biomacromolecules, 2004. 5, 1219-1230. (23) Omoike, A.; Chorover, J. Adsorption to goethite of extracellular polymeric substances from Bacillus subtilis. Geochim. Cosmochim Ac., 2006. 70, 827-838. (24) Cuss, C.W.; Guéguen, C. Determination of relative molecular weights of fluorescent components in dissolved organic matter using asymmetrical flow field-flow fractionation and parallel factor analysis. Anal. Chim. Acta., 2012, 733 (6), 98-102. (25) Li, W.T.; Chen, S.Y.; Xu, Z.X.; Li, Y.; Shuang, C.D.; Li, A.M. Characterization of dissolved organic matter in municipal wastewater using fluorescence PARAFAC analysis and chromatography multi-excitation/emission scan: A comparative study. Environ. Sci. Technol, 2014. 48(5), 2603-2609. (26) Stedmon, C.A.; Bro, R. Characterizing dissolved organic matter fluorescence with parallel factor analysis: a tutorial. Limnol. Oceanog.-Meth., 2008. 6, 572-579. (27) Abdulla, H.A.N.; Sleighter, R.L.; Hatcher, P.G. Two dimensional correlation analysis of Fourier transform ion cyclotron resonance mass spectra of dissolved organic matter: A new graphical analysis of trends. Anal. Chem. 2013, 85, 2895-3902. (28) Hur, J.; Lee, B.M. Comparing the heterogeneity of copper-binding characteristics for two different-sized soil humic acid fractions using fluorescence quenching combined with 2D-COS. The Scientific World Journal, 2011. 11, 1865-1876. (29) Yan, W.; Zhang, J.; Jing, C. Adsorption of Enrofloxacin on montmorillonite: Twodimensional correlation ATR/FTIR spectroscopy study. J. Colloid Interf. Sci., 2013. 390, 196-203. (30) Chen, W.; Qian, C.; Liu, X.Y.; Yu, H.Q. Two-dimensional correlation spectroscopic analysis on the interaction between humic acids and TiO2 nanoparticles. Environ. Sci. Technol. 2014, 48 (19), 11119-11126. (31) Domínguez, L.; Rodríguez, M.; Prats, D. Effect of different extraction methods on bound EPS from MBR sludges Part II: Influence of extraction methods over molecular weight distribution, Desalination, 2010. 262, 106-109. (32) Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved Synthesis of Graphene Oxide. ACS Nano, 2010. 4(8), 4806-4814. 18

ACS Paragon Plus Environment

Environmental Science & Technology

497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526

(33) Kim, H.C.; Dempsey, B.A. Comparison of two fractionation strategies for characterization of wastewater effluent organic matter and diagnosis of membrane fouling. Water res., 2012. 46(11), 3714-3722. (34) Liu, Y.; Fang, H. Extraction of extracellular polymeric substances (EPS) of sludges, J. Biotechnol., 2002. 95, 249-256. (35) Yan, H.; Wu, H.; Li, K.; Wang, Y.; Tao, X.; Yang, H.; Li A.; Cheng, R. Influence of the surface structure of graphene oxide on the adsorption of aromatic organic compounds from water. ACS appl. Mater. Inter., 2015. 7(12), 6690-6697. (36) Apul, O.G.; Karanfil, T. Adsorption of synthetic organic contaminants by carbon nanotubes: A critical review. Water Res. 2015. 68, 34-55. (37) Chen, X.; Chen, B. Macroscopic and spectroscopic investigations of the adsorption of nitroaromatic compounds on graphene oxide, reduced graphene oxide, and graphene nanosheets. Environ. Sci. Technol., 2015, 49 (10), 6181-6189. (38) Xu, H.; Cai, H.; Yu, G.; Jiang, H. Insights into extracellular polymeric substances of cyanobacterium Microcystis aeruginosa using fractionation procedure and parallel factor analysis. Water Res., 2013. 47, 2005-2014. (39) Qu, F.; Liang, H.; Wang, Z.; Wang, H.; Yu, H.; Li, G. Ultrafiltration membrane fouling by extracellular organic matters (EOM) of Microcystis aeruginosa in stationary phase: Influences of interfacial characteristics of foulants and fouling mechanisms. Water Res., 2012. 46(5), 1490-1500. (40) Zhang, Y.; Wu, C.; Guo, S.; Zhang, J. Interactions of graphene oxide with proteins and peptides. Nanotechnol. Rev., 2013. 2(1), 27-45. (41) Xu, H.; Jiang, H. Effects of cyanobacterial extracellular polymeric substances on the stability of ZnO nanoparticles in eutrophic shallow lakes. Environ. Pollut., 2015. 197, 231-239. (42) Vedantham, G.; Sparkes, H.G.; Sane, S.U.; Tzannis, S.; Przybycien, T.M. A holistic approach for protein secondary structure estimation from infrared spectra in H2O solutions. Anal. Biochem., 2000. 285, 22-49. (43) Park, Y.; Noda, I.; Jung, Y.M. Two-dimensional correlation spectroscopy in polymer study. Frontiers in Chemistry, 2015. 3, 1-16.

527 528 529 530 531 532 533 534 535 536 537 538 19

ACS Paragon Plus Environment

Page 20 of 27

Page 21 of 27

Environmental Science & Technology

539 540 541 542

Tables

543 544

Table 1. Freundlich Isotherm model parameters of EPS adsorption onto graphene materials. Adsorbent type

K F*

1/n**

R2

GO

3.49 ± 0.34

0.77 ± 0.03

0.97

rGO-2h

6.59 ± 0.74

0.71 ± 0.04

0.95

rGO-8h (mg C/g)/(mg C/L)(1/n) ** Unitless

25.09 ± 1.64

0.37 ± 0.02

0.94

545 546 547 548 549 550

*

551

C2, C3 and C4) for EPS adsorption onto graphene materials.

Table 2. Freundlich isotherm model parameters of the individual PARAFAC components (C1,

Components C1

C2

C3

C4

Adsorbent type GO

KF (QSE/mg)/(QSE)(1/n) 4.0×10-5 ± 2.10-5

1/n (Unitless) 2.10 ± 0.09

0.98

rGO-2h

0.03 ± 0.01

0.83 ± 0.09

0.86

rGO-8h

0.08 ± 0.01

0.66 ± 0.04

0.95

GO

0.11 ± 0.01

0.79 ± 0.03

0.99

rGO-2h

0.43 ± 0.12

0.62 ± 0.06

0.91

rGO-8h

2.61 ± 0.14

0.35 ± 0.01

0.99

GO

0.01 ± 0.002

0.80 ± 0.03

0.99

rGO-2h

0.05 ± 0.003

0.50 ± 0.03

0.96

rGO-8h

0.12 ± 0.002

0.39 ± 0.01

0.98

GO

0.26 ± 0.05

0.38 ± 0.08

0.69

rGO-2h

0.29 ± 0.03

0.36 ± 0.04

0.87

rGO-8h

0.72 ± 0.04

0.25 ± 0.02

0.95

552 553 554 555 20

ACS Paragon Plus Environment

R2

Environmental Science & Technology

Page 22 of 27

556 557 558 559 560

Figures

(a) Absorbance (arb. unit)

1635 1540

1000 1040

1400 1454 1236

2000

1750

561

1500 1250 1000 -1 Wavenumber (cm )

750

500

45

Proportion of fractions (%)

40

(b)

35 30 25 20 15 10 5 0

562 563

HPO-A

HPO-N

TPI-A

TPI-N

HPI

Figure 1. Characteristics of EPS (a) FTIR spectra, and (b) proportion of resin fractions. 21

ACS Paragon Plus Environment

Page 23 of 27

564 565 566 567 568 569 570 571 572

Environmental Science & Technology

Figure 2. Fluorescent components identified by EEM-PARAFAC (a) C1, (b) C2, (c) C3, and (d) C4.

573

22

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 27

6 GO

(a)

rGO-2h

5

rGO-8h EPS initial = 12.7

C2/C1

4 3 2

1 Increasing adsorption amount (More surface coverage) 0 20

30

574

40 50 Percent adsorption (%)

60

70

1.6 Increasing adsorption amount (More surface coverage)

(b)

1.4 1.2

C4/C3

1 GO 0.8

rGO-2h rGO-8h

0.6

EPS initial = 3.50 0.4 0.2 0 20 575 576 577 578 579 580

30

40 50 Percent adsorption (%)

60

70

Figure 3. Changes in the ratios of protein-like to humic-like PARAFAC components for the residual EPS after adsorption with increasing percent adsorption or with less surface coverage. (a) C2/C1 and (b) C4/C3.

23

ACS Paragon Plus Environment

Page 25 of 27

Environmental Science & Technology

Amide I

(a)

Amide II Polysaccharides ester PO2-

Absorbance (arb. unit)

Aromatic C-C Carboxylate uronic acid

EPS

epoxy C-O

Aromatic C-C

Carboxy COO

C=O 2000

1700

(b)

GO

Alkoxy C-O

1400 1100 Wavenumber (cm-1)

581

EPS-adsorbed GO

800

500

Amide I Amide II

Absorbance (arb. unit)

Polysaccharides Aromatic C-C ester PO2Carboxylate uronic acid

EPS

epoxy C-O

EPS-adsorbed rGO-8h

Aromatic C-C Carboxy COO

C=O 2000 582 583 584

1700

rGO-8h

Alkoxy C-O

1400 1100 Wavenumber (cm-1)

800

500

Figure 4. Comparison of FTIR spectra of GO (a) and rGO-8h (b) with and without adsorption of EPS in the band regions from 750 to 1860 cm-1.

24

ACS Paragon Plus Environment

Environmental Science & Technology

585 586 587 588

Figure 5. Synchronous and asynchronous maps of 2D FTIR correlation spectra of GO (a, b) and rGO-8h (c, d) with different EPS amounts in the region of 750-1860 cm-1.

589 590 591 592 593 594 595

25

ACS Paragon Plus Environment

Page 26 of 27

Page 27 of 27

Environmental Science & Technology

ACS Paragon Plus Environment