Nzvi-Catalyst Research Paper

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

View Article Online

Environmental
View Journal

Science
Water Research & Technology
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: A. Masud, N. G.
Chavez Soria, D. S. Aga and N. Aich, Environ. Sci.: Water Res. Technol., 2020, DOI: 10.1039/D0EW00140F.
Volume 4
This is an Accepted Manuscript, which has been through the
Environmental
Number 2
February 2018
Pages 101-338
Royal Society of Chemistry peer review process and has been
Science accepted for publication.
Water Research & Technology
rsc.li/es-water
Accepted Manuscripts are published online shortly after acceptance,
before technical editing, formatting and proof reading. Using this free
service, authors can make their results available to the community, in
citable form, before we publish the edited article. We will replace this
Accepted Manuscript with the edited and formatted Advance Article as
soon as it is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes to the
text and/or graphics, which may alter content. The journal’s standard
ISSN 2053-1400 Terms & Conditions and the Ethical guidelines still apply. In no event
PAPER
Jeyong Yoon et al.
Electrochemical lithium recovery and organic pollutant
shall the Royal Society of Chemistry be held responsible for any errors
removal from industrial wastewater of a battery recycling plant

or omissions in this Accepted Manuscript or any consequences arising


from the use of any information it contains.

rsc.li/es-water
Page 1 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

Water Impact Statement


Emerging pharmaceuticals and personal care products (PPCPs) including antibiotics,
antidepressants, or stimulants remain as a complex mixture in our wastewater or natural aquatic

Environmental Science: Water Research & Technology Accepted Manuscript


systems and have shown to pose adverse effects on fish and other aquatic organisms.
Understanding the competitive interactions of diverse PPCPs with the newly engineered
graphene-iron nanohybrids can provide with significant insight about their effective removal
from water through adsorption and/or advanced oxidation processes (AOP).
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.
Environmental Science: Water Research & Technology Page 2 of 40
View Article Online
DOI: 10.1039/D0EW00140F

1
2 Adsorption and Advanced Oxidation of Diverse Pharmaceuticals and

Environmental Science: Water Research & Technology Accepted Manuscript


3 Personal Care Products (PPCPs) from Water Using Highly Efficient
4 rGO-nZVI Nanohybrids

5
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

7 Arvid Masud,1 Nita G. Chavez Soria,2 Diana S. Aga,2 and Nirupam Aich1,*

8 1Department of Civil, Structural and Environmental Engineering, University at Buffalo, The State

9 University of New York, Buffalo, NY 14260

10 2Department of Chemistry, University at Buffalo, The State University of New York, Buffalo, NY 14260

11 Submitted to

12 Environmental Science: Water Research & Technology

13 06/22/2020

14

15

16

17

18

19

20

21

22

23

24

25

26 * Corresponding Author: Nirupam Aich, Phone: 716-645-0977, Email: nirupama@buffalo.edu

1
Page 3 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

27 Abstract:

28 Engineered nanomaterials, like graphene with tunable adsorption sites and nanoscale zero-valent

Environmental Science: Water Research & Technology Accepted Manuscript


29 iron (nZVI) with unique redox chemistry, offer great prospect for removing pharmaceutical and

30 personal care products (PPCPs) through adsorption and catalytic advanced oxidation process
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

31 (AOP) in comparison to bulk materials. Although PPCPs are found to be as a complex mixture in

32 wastewater and in the environment, most studies regarding nano-enabled PPCP removal reported

33 results with only one PPCP at a time and typically at high initial concentrations. In this study, we

34 utilized reduced graphene oxide (rGO) to support nZVI to synthesize rGO-nZVI nanohybrid (NH)

35 and used rGO-nZVI NH for the removal of a complex mixture of 12 diverse PPCPs that includes

36 antibiotic, anti-inflammatory, anti-seizure, and antidepressant pharmaceuticals, and are

37 recalcitrant in the environment. We also tested the removal of PPCPs at their individual

38 environmentally relevant concentrations at ppb level. The rGO-nZVI NH synergistically

39 performed as both an adsorbent and a heterogeneous Fenton catalyst (for AOP) in the presence of

40 H2O2, to remove ~95-99% of environmentally relevant concentrations (200 ppb) of the PPCPs

41 within 10 minutes. Even in the absence of H2O2, the hybridization resulted in better adsorptive

42 property (14-72% more removal) in the rGO-nZVI NH compared to the parent nanomaterials (rGO

43 or nZVI) for various PPCPs in the mixture, removing ~82-99% of the PPCPs, at the end of 30

44 minutes, with comparatively slower kinetics (~3-5.5 times) than in presence of H2O2.

45 Hydrophobic PPCPs were removed faster and more with both the adsorption and AOP. The

46 potential of utilizing rGO-nZVI NH in drinking/wastewater treatment system, or even in point-of-

47 use system where necessary, was substantiated by the enhanced and fast PPCP removal capacity

48 of the nanohybrid.

49

2
Environmental Science: Water Research & Technology Page 4 of 40
View Article Online
DOI: 10.1039/D0EW00140F

50 1. Introduction

51 Pharmaceuticals and personal care products (PPCPs) are a diverse set of compounds including

Environmental Science: Water Research & Technology Accepted Manuscript


52 antibiotics, analgesics, anti-seizure, nonsteroidal anti-inflammatory (NSAID), and antidepressants

53 used to treat diseases and improve quality of life, thus heavily used worldwide. Conventional
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

54 wastewater treatment plants (WWTPs) with primary (e.g., sedimentation) and secondary (e.g.,

55 activated sludge) treatment processes exhibit limited PPCP removal efficiencies.1-4 This causes

56 inevitable release of mixtures of PPCPs in natural aquatic environment,5 resulting in their presence

57 at ppb to ppt (µg/L to ng/L) levels in water bodies.6-8 Recently, trace level concentrations of

58 selective serotonin reuptake inhibitors (SSRIs), some of the most prescribed antidepressants, have

59 been detected in fish from the Great Lakes.9 Many of the PPCPs10-12 including SSRIs,4, 13-16 have

60 been reported to induce adverse physiological effects in aquatic organisms, e.g., mortality,

61 developmental and reproductive disorder, endocrine disruption.17, 18 Moreover, drinking water

62 quality may also be affected from the use of PPCP contaminated surface water that are used as

63 intake source for drinking water treatment plants (DWTPs).19-21 This warrants incorporation of

64 advanced treatment technologies i.e., adsorption and advanced oxidation process (AOP) into

65 WWTPs and DWTPs for the effective removal of these recalcitrant PPCPs including SSRIs.22-25

66 Engineered nanomaterials, having high surface area with tunable adsorption sites and

67 enhanced catalytic property, possess high potential for removing PPCPs through adsorption and

68 AOP, respectively, in comparison to conventional bulk materials.26-28 Graphene and its derivatives

69 namely graphene oxide (GO) and reduced graphene oxides (rGO) are two dimensional (2D)

70 carbonaceous nanomaterials that can favorably adsorb various PPCPs, mainly through

71 hydrophobic and electrostatic interaction.29-31 However, the adsorbed PPCP requires further

72 treatment for their degradation. On the other hand, nanoscale zero-valent iron (nZVI) particles has

3
Page 5 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

73 been explored as heterogeneous Fenton catalysts for AOP based PPCP degradation.32, 33 nZVI can

74 generate Fe2+ ions through slow dissolution at surface, which promotes formation of reactive

Environmental Science: Water Research & Technology Accepted Manuscript


75 oxygen species (ROS) in the presence of hydrogen peroxide (H2O2). These nascent ROS react

76 with PPCPs to degrade them. But, nZVI suffers from low surface reactivity due to aggregation
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

77 because of its high magnetic and van der Waals attraction forces.34, 35 To mediate these challenges,

78 a nanohybrid approach can be followed36 that will utilize planar rGO to support nZVI particles

79 resulting in rGO-nZVI nanohybrids (NHs). rGO-nZVI NHs can offer a synergistic approach for

80 enhanced PPCP removal where rGO can adsorb PPCPs on its surface, that can simultaneously be

81 degraded using AOP catalyzed by nZVI.29, 37 Moreover, rGO can improve nZVI’s catalytic

82 property by providing fast electron transfer38, 39 and resisting nZVI aggregation.36

83 Recently, a few studies have reported graphene supported iron nanoparticles (mostly iron

84 oxides and hydroxides, but not nZVI) for heterogeneous Fenton catalysis and exploited combined

85 adsorption and AOP for enhanced PPCP removal.40-43 For example, goethite (α-FeOOH)

86 nanoparticle deposited rGO hydrogel utilized the advantage of strong Fe-O-C bond in generation

87 of reactive oxygen species which enhanced antibiotic tetracycline degradation.43 However, in the

88 case of rGO-nZVI, the zero valent state of iron in nZVI does not rely on the Fe2+/Fe3+ conversion

89 cycle facilitated by Fe-O-C bonds and can keep donating electron to generate Fe2+ at the surface

90 to directly interact with H2O2, thereby, making rGO-nZVI more catalytic than other graphene-iron

91 materials44 and our choice of material for this study. Moreover, all the above-mentioned studies

92 only focused on removal/degradation of only one antibiotic, and not a mixture of PPCPs. Also,

93 all these studies (except one) used PPCPs at significantly higher concentrations (>1 mg/L)

94 compared to the environmentally relevant PPCP concentrations (i.e., sub-ppm and ppb level).28, 32,

95 45-51 Although Peng et al. tested graphene for adsorption of a mixture of 7 antibiotics at sub-ppm

4
Environmental Science: Water Research & Technology Page 6 of 40
View Article Online
DOI: 10.1039/D0EW00140F

96 level,52 realistically, actual wastewater contains a mixture of diverse PPCPs that include SSRIs,

97 SNRIs (Serotonin-Norepinephrine Reuptake Inhibitors), antibiotics, anti-seizure, and NSAID

Environmental Science: Water Research & Technology Accepted Manuscript


98 pharmaceuticals. These various PPCPs including SSRIs may interact with each other affecting

99 their overall removal by these nanomaterials.53, 54 Therefore, evaluating the performance of rGO-
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

100 nZVI NHs in removing a mixture of PPCPs rather than individual ones is more relevant and

101 potentially benefit the water treatment processes by delineating the competition between PPCPs

102 during adsorption and AOP.

103 The objectives of this study were, for the first time, to determine and compare the extent

104 of adsorptive and AOP-based removal of (i) a complex mixture of diverse PPCPs, including SSRIs,

105 antibiotics, anti-seizure, NSAIDs, and stimulants, (ii) at environmentally relevant ppb level

106 concentration using (iii) catalytic rGO-nZVI NHs. We also aimed to compare the PPCP removal

107 performance of rGO-nZVI NHs with that of its parent materials, i.e., rGO and nZVI, to delineate

108 the effect of hybridization on PPCP removal. To achieve these aims, rGO-nZVI NHs were

109 synthesized through chemical reduction process and comprehensively characterized with

110 transmission electron microscopy (TEM), X-ray diffraction spectrometry (XRD), Raman

111 spectroscopy, thermogravimetric analysis (TGA), and Fourier-transform infrared (FTIR)

112 spectroscopy. The performance of rGO-nZVI NHs for the removal of a mixture of 12 PPCPs at

113 environmentally relevant ppb level concentration was investigated using liquid chromatography

114 tandem mass spectrometry (LC-MS/MS). A series of PPCP removal tests using the nanohybrids

115 and their parent nanomaterials (i.e., rGO and nZVI) and their mixtures were conducted to

116 determine the synergistic effect of hybridization and the effect of PPCP mixture along with

117 determining the underlying PPCP removal mechanisms (i.e., adsorption and/or AOP).

118 2. Materials and methods

5
Page 7 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

119 2.1 Chemical and Reagents

120 Single layer graphene oxide (GO, >99.3 wt%, thickness 0.43-1.23 nm), sodium borohydride

Environmental Science: Water Research & Technology Accepted Manuscript


121 (NaBH4, >98%), and ferric chloride hexahydrate (FeCl3.6H2O, 98%) were purchased from US

122 Research Nanomaterials, Inc. (Houston, TX), Fisher Scientific (Fair Lawn, NJ), and Acros
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

123 Organics (Geel, Belgium), respectively. Standards for the selected 12 pharmaceuticals including

124 acetaminophen (ACT), caffeine (CAF), carbamazepine (CBZ), carbamazepine-d10, citalopram

125 (CIT), diclofenac (DIC), fluoxetine (FLX), ibuprofen (IBU), lamotrigine (LAM), naproxen (NPX),

126 paroxetine maleate (PRX), sulfamethoxazole (SMX), and venlafaxine (VEN) were purchased from

127 Cerilliant (Round Rock, Texas), Sigma Aldrich (St Louis, MO) and Cambridge Isotopes

128 (Tewsbury, MA). All the PPCPs had chromatographic purity over 98%. N,N′-diethyl-p-

129 phenylenediamine (DPD), ammonium molybdate, and potassium iodide were obtained from Hach

130 (Loveland, CO). Deionized (DI) water with conductivity of 18.2 Ω from Barnstead

131 NANOpureTM water system (Waltham, MA) was used for all the experiments. LC-MS grade

132 acetonitrile (>99.9%) was purchased from EMD Millipore Corporation (Billerica, MA). Formic

133 acid (88%) and n-butanol (>99.9%) was obtained from Fisher Chemical (Pittsburgh, PA) and

134 Sigma Aldrich (St Louis, MO) respectively.

135 2.2 Synthesis and characterization of nanohybrids

136 rGO-nZVI NHs were synthesized following a well-established wet chemistry method.55 Briefly,

137 100 mg GO was dispersed in 200 mL DI water (i.e., providing 0.5 mg/mL concentration) by

138 sonication for 2 hours (amplitude 50, pulse on 8 sec, pulse off 2 sec) with a microtip (1/4”) based

139 ultrasonic dismembrator (Q700, Qsonica Sonicators, Newtown, CT). Subsequently, 1 g

140 FeCl3.6H2O was added to make iron concentration 18.5 mM in the dispersion. A peristaltic pump

141 was used to add 20 mL of NaBH4 solution (1.59 M) with a flow rate of 0.5 mL/minute into the

6
Environmental Science: Water Research & Technology Page 8 of 40
View Article Online
DOI: 10.1039/D0EW00140F

142 mixture to co-reduce GO and iron salt. The reaction was carried out in an inert atmosphere with

143 constant N2 gas flow (0.3 L/minute) under continuous magnetic stirring (300 rpm) at room

Environmental Science: Water Research & Technology Accepted Manuscript


144 temperature. After adding the NaBH4, the suspension was kept stirring for another half an hour

145 under N2 gas. The suspension was filtered and was washed with ethanol and DI water for couple
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

146 of times through vacuum filtration (0.45 µm high volume low pressure (HVLP) filter) for removing

147 excess ions. The solid residues were dried overnight in vacuum oven at 60 oC before further

148 experiments. In case of parent nanomaterial synthesis, rGO and nZVI, similar synthesis protocol

149 was followed using their respective precursors individually. For rGO synthesis, no iron salt was

150 added and 20 mL 0.2 M NaBH4 was used dropwise to reduce GO. For nZVI synthesis, 1 g

151 FeCl3.6H2O was dissolved in 200 mL DI water without addition of any GO. For reduction of iron

152 salt, 20 mL 1.32 M NaBH4 was used under N2 gas flow.

153 Transmission electron microscopy (TEM, JEOL JEM 2010, JEOL USA, Inc.) was

154 performed at an accelerating voltage of 200 KV to image the physical morphology of the rGO,

155 nZVI, and rGO-nZVI NHs. 1 mg of sample was dispersed in 10 mL ethanol with ultrasonic

156 dismembrator for TEM sample preparation. One drop of the dispersion was placed on a carbon-

157 coated copper grid (100 mesh, Ted Pella, Inc., Redding, CA) using micro-pipette and was air-dried

158 for 20 minutes before analysis. The X-ray diffraction (XRD) analysis of the samples were

159 performed at room temperature by Rigaku Ultima IV (Rigaku Corporation) using the Kα emission

160 of a Cu X-ray source (λ = 1.5418 Å). The incident angle range was kept in the range from 5° to

161 90° with a scan rate of 2º/min and step size of 0.02º. The vibrational modes and lattice order of

162 the molecular structures were studied using Raman Microscope (Renishaw InVia, Renishaw plc.)

163 with incident laser wavelength of 514 nm. 1 mg of dried samples was placed as a thin layer on a

164 glass slide to carry out the Raman spectroscopy. Thermogravimetric analysis (TGA,

7
Page 9 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

165 Micromeritics Instrument Corp.) with the samples were carried out in air with a ramp rate of

166 10 °C/min to 1000 °C to analyze the thermal stability of the samples. A Fourier-transform infrared

Environmental Science: Water Research & Technology Accepted Manuscript


167 (FTIR) spectroscopy instrument (1760 FTIR, Perkin-Elmer) was used to obtain infrared adsorption

168 spectrum of the samples within the range of 500–4000 cm−1 with 4 cm−1 resolution to characterize
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

169 the functional groups present in the sample. Elemental carbon content in the dry samples were

170 measured with a carbon, hydrogen, and nitrogen (CHN) bulk elemental analyzer (2400 Series II,

171 Perkin-Elmer).

172 2.3 PPCP removal experiments

173 Figure S2 presents a schematic showing the experimental steps for evaluating the efficacy of rGO-

174 nZVI NHs for PPCP removal through both adsorption and AOP based adsorption. In general,

175 PPCP mixtures were dosed with nanomaterials (either rGO, nZVI, or rGO-nZVI) in individual

176 vials and subjected to either only adsorption (no H2O2) or AOP augmented adsorption (with H2O2)

177 while shaking. This was followed by sample collection at different time intervals, reaction

178 quenching, centrifugation (to remove nanomaterials), and LC-MS/MS analysis to obtain and

179 compare PPCP removal efficiency to evaluate the performance of catalyst/oxidant system.

180 Detailed description of the experiments with specific quantities and time is given below. Two

181 different aqueous stock solutions were prepared for PPCP batch removal tests. One of the stock

182 solutions consisted of all 12 PPCPs that are listed in Table 1 with their chemical structures shown

183 in Figure S3, each having a concentration of 200 µg/L (ppb). The other stock solution, used for

184 kinetic removal tests, consisted of 6 PPCPs (serial no. 4, 5, 6, 10, 11, and 12 from Table 1) also

185 having concentration of 200 µg/L. PPCP removal tests were conducted under three different

186 experimental arrangements: (i) only H2O2 (control), (ii) nanomaterials (i.e., rGO-nZVI NHs;

187 parent nanomaterials rGO or nZVI; mixture of parent nanomaterials rGO and nZVI), and (iii)

8
Environmental Science: Water Research & Technology Page 10 of 40
View Article Online
DOI: 10.1039/D0EW00140F

188 combination of nanomaterials and H2O2. For all of the removal tests, 15 mL PPCP solution was

189 placed in a glass vial covered with aluminum foil and kept under constant shaking at 150 rpm for

Environmental Science: Water Research & Technology Accepted Manuscript


190 30 minutes. The nanomaterial dosage were set at 8 mg in the 15 mL PPCP solution (i.e., 533 mg/L

191 nanomaterials) for all the removal tests. In the case of removal test with mixture of parent
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

192 nanomaterials, 1.9 mg of rGO (i.e., 127 mg/L rGO) and 6.1 mg (i.e., 407 mg/L nZVI) of nZVI

193 were used. The amounts of individual nanomaterials (i.e., rGO and nZVI) for the mixture was

194 same as their amounts in the synthesized rGO-nZVI NH. For removal tests with H2O2, 15.60 µL

195 of 30 wt.% H2O2 was added to the 15 mL PPCP solution to achieve the final H2O2 concentration

196 of 10 mM. The initial pH was adjusted to 3 with 1M HCl and 1 M NaOH in all cases, which has

197 been reported in literature as the most favorable operating pH condition for heterogeneous Fenton

198 catalysts.32, 56 For kinetic removal tests, 300 µL aliquots were collected at 0, 5, 10, 15, 20, 25, and

199 30 minutes. 10 mM n-Butanol was added to the aliquots instantly to scavenge all the ROS and

200 centrifuged at15000 rpm for 1 minute. The amount of n-Butanol used in this study was chosen

201 based on the range presented in previous literature for n-Butanol amount for complete quenching

202 of ROS produced during heterogeneous Fenton reactions.57, 58 200 µL of supernatant was collected

203 and consequently spiked with an internal standard (Carbamazepine-d10, final concentration 50

204 µg/L) prior to LC-MS/MS analysis.

205 Hach (method 10290) procedures were followed to quantify residual H2O2 concentration

206 in batch reactions through DPD colorimetric method.59

207 2.4 LC-MS/MS analysis of PPCPs

208 A previously established LC/MS/MS method was adapted to monitor the 12 PPCPs in this study

209 (Table 1).60 In summary, analysis of PPCPs was perfo1rmed using an Agilent 1100 LC system

210 coupled to a 6410 triple quadrupole mass spectrometer (Palo Alto, CA). Chromatographic

9
Page 11 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

211 separation was achieved with a Waters Cortecs™ C18+ column (Milford, MA) with an inner

212 diameter and length of 2.1 mm and 150 mm respectively, and a 2.7 µm particle size. The column

Environmental Science: Water Research & Technology Accepted Manuscript


213 was fitted with a 5.0 mm x 2.1 mm, 2.7 µm Cortecs™ guard cartridge (Milford, MA). A 39-

214 minute run time method was used with mobile phase A (0.3% formic acid in DI water obtained
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

215 from NANOpure™ water system, Waltham, MA) and B (100% acetonitrile). The gradient used

216 was as followed: 90% A for 2 minutes, ramp to 95% B over 23 minutes, followed by a four minute

217 hold time and brought back to starting conditions over 3 minutes. The column was equilibrated

218 for 7 minutes at the end of the gradient. The flow rate was set to 0.2 mL/min and a volume of 20

219 µL was injected. The mass spectrometer was set to positive electrospray ionization with 4 kV

220 spray voltage, 300 ◦C drying gas temperature, 15 psi nebulizer pressure (N2) and 6 L/min drying

221 gas flow (N2). Data collection and analysis were performed using Agilent Technologies Mass

222 Hunter Software Version B (Palo Alto, CA). Details of the multiple reaction monitoring (MRM)

223 transitions are listed in Table 1.

10
Environmental Science: Water Research & Technology Page 12 of 40
View Article Online
DOI: 10.1039/D0EW00140F

224 Table 1. Target Analytes and LC/MS/MS parameters.

Environmental Science: Water Research & Technology Accepted Manuscript


Serial Analyte Type Log Kow(§) Fragmentor MRM Collision
No. Voltage (eV) transitions Energy
(Q and q) (eV)
1 Acetaminophen Analgesic 0.5 100 152→65 29
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

(ACT) 152→110 14
2 Caffeine (CAF) Stimulant -0.1 126 195→42 33
195→138 18
3 Carbamazepine Anti-seizure 2.5 119 237→179 37
(CBZ) 237→194 18
4 Citalopram (CIT) Antidepressant 3.2 150 325→109 25
(SSRI) 325→262 15
5 Diclofenac (DIC) Nonsteroidal anti- 4.4 100 296→214 34
inflammatory 296→250 10
6 Fluoxetine Antidepressant 4.0 40 310→148 4
(SSRI)
7 Ibuprofen (IBU) Nonsteroidal anti- 3.5 60 207→119 20
inflammatory 207→161 10
8 Lamotrigine Anticonvulsant 1.4 140 256→109 60
256→211 20
9 Naproxen (NPX) Nonsteroidal anti- 3.3 100 231→115 60
inflammatory 231→185 10
(NSAID)
10 Paroxetine (PRX) Antidepressant 3.5 100 330→123 25
(SSRI) 330→192 20
11 Sulfamethoxazole Antibiotic 0.9 120 254→156 10
(SMX) 254→108 22
12 Venlfaxine (VEN) Antidepressant 2.9 114 278→121 30
(SNRI) 278→215 10
225 (§) https://pubchem.ncbi.nlm.nih.gov

11
Page 13 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

226 3. Results and Discussion

227 3.1 Physicochemical properties of rGO-nZVI NHs and parent nanomaterials

Environmental Science: Water Research & Technology Accepted Manuscript


228 The physical morphology of the rGO-nZVI NHs along with parent nanomaterials rGO and nZVI

229 were analyzed using TEM images (Figure 1). The GO nanosheets are quasi-transparent with
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

230 lateral dimensions between 1-5 µm on average (Figure 1a). In the case of bare nZVI (Figure 1b),

231 the nanoparticles aggregated to form a chainlike connected structure through iron oxide formation

232 in the outer shell. In the case of rGO-nZVI NH (Figure 1c), much smaller nZVI particles

233 (compared to bare nZVI) were segregated and distributed throughout the rGO sheets. Further

234 analysis of the TEM images helped to determine the particle size distribution of nZVI particles for

235 bare nZVI and rGO-nZVI NHs (Figure S1). The nZVI particle size was mostly in the range of

236 100-200 nm for bare nZVI, whereas it was in the range of 10-40 nm for the rGO-nZVI NH. Similar

237 decrease in size of the nZVI particles due to hybridization with rGO nanosheets has been well

238 documented in previous literatures.35, 61 This confirms that the hybridization led to the successful

239 prevention of nZVI aggregation by the rGO nanosheets.

12
Environmental Science: Water Research & Technology Page 14 of 40
View Article Online
DOI: 10.1039/D0EW00140F

Environmental Science: Water Research & Technology Accepted Manuscript


Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

240

241 Figure 1. TEM images of (a) rGO, (b) nZVI, (c) and (d) rGO-nZVI NH.

242 Figures 2(a, b, c, and d) present respectively the XRD, Raman, TGA, and FTIR spectra of

243 synthesized rGO-nZVI NHs and parent rGO and nZVI. The XRD spectrum of rGO shows the

244 characteristic carbon sharp peak for GO at ~10o and broadened peak for graphene at ~26o.35 The

245 XRD spectrum of bare nZVI has both Fe0 characteristic peaks at 45 and 65 along with Fe3O4

246 characteristics peaks at 30°, 35°, and 57°.35 The XRD spectrum of rGO-nZVI shows predominant

247 Fe0 peak at ~45° indicating that the iron nanoparticles on the graphene sheet were mostly present

248 in zero valence state. This suggests that graphene could resist oxidation of nZVI nanoparticle in

249 the rGO-nZVI NH.62 The Raman spectra shown in Figure 2(b) provide information regarding

250 surface defects for different nanomaterials studied here. The ratio of the defect D-band at (ID ∼

251 1350 cm–1) and graphitic G-band (IG ∼ 1590 cm–1) in the Raman spectrum, ID/IG is a measure of

13
Page 15 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

252 lattice defects in graphene lattice structure.61 The higher defect in graphitic structure for the NH

253 (ID/IG = 1.05) compared to rGO (ID/IG = 0.79) confirms the impregnation of nZVI particle on

Environmental Science: Water Research & Technology Accepted Manuscript


254 graphene surface.55 Additionally, the signal for iron nanoparticles in lower Raman shift range (100

255 – 500 cm-1) also suggests the hybridization of nZVI with graphene. Figure 2c shows the TGA
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

256 results for rGO, nZVI, and rGO-nZVI NH, and shows that rGO gradually undergoes combustion

257 in the presence of air between 250 oC and 600 oC and decreases to 10% of its original mass which

258 finally decreases to 7% of its original mass at 1000 oC.63 However, for nZVI and rGO-nZVI, we

259 observe an increase in the mass after 400-500 oC during TGA. In case of nZVI the mass remains

260 same until 400 oC and then increases sharply to 139% at 550 oC. For rGO-nZVI, the mass steadily

261 decreases (due to rGO combustion) to 82% of its original mass until 500 oC and then sharply

262 increases to 96%. For both nZVI and rGO-nZVI, the reason of mass increase after 400-500 oC

263 can be attributed to the iron oxidation in the presence of air.64, 65 This suggests that the iron present

264 in nZVI, both in the rGO-nZVI NH and parent nanomaterial, are mostly in reduced form, and gets

265 oxidized during TGA. The FTIR spectrum of rGO suggests most of the functional groups were

266 removed during reduction from GO, as the peak for stretching vibration of C=O (at 1720 cm-1) and

267 C-O (at 1040 cm-1) bonds were significantly weakened in case of rGO compared to GO.66 In case

268 of FTIR spectrum of nZVI, the peaks at 547, 840, and 1630 cm-1 correspond to the Fe-O stretch.36

269 The weakened peak of the Fe-O band for rGO-nZVI NH spectrum can be attributed to lesser

270 oxidation and smaller size of nZVI in the nanohybrid which results in more amorphous

271 structures.67 The stretching vibration band for the functional groups are also absent in case of

272 rGO-nZVI NH. The shift of the C=C band to 1560 cm-1 (compared to 1605 cm-1) can be attributed

273 to chemical bond between nZVI and the graphitic skeleton of rGO.68 The CHN analysis suggests

274 the elemental carbon content in rGO and rGO-nZVI NH is 62.35% and 14.91% respectively. This

14
Environmental Science: Water Research & Technology Page 16 of 40
View Article Online
DOI: 10.1039/D0EW00140F

14.91
275 indicates the rGO content in the NH to be 23.93% (= 0.623 %). If the rest of the rGO-nZVI

NH is considered to be nZVI, the ratio of rGO and nZVI turns out to be ~1:3.2, which conforms

Environmental Science: Water Research & Technology Accepted Manuscript


276

277 to our previous study.61


Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

278

279 Figure 2. (a) XRD, (b) Raman, (c) TGA, and (d) FTIR spectra of rGO-nZVI NH and parent

280 nanomaterials rGO and nZVI.

281 3.2 PPCP removal efficiencies for rGO-nZVI NHs and parent nanomaterials

282 3.2.1. Adsorptive removal of PPCPs

283 Aqueous solution combining a diverse matrix of 12 PPCPs including 4 antidepressants (1 SNRI +

284 3 SSRIs), 3 NSAIDs, 1 analgestic, 2 anti-seizure, 1 antibiotic, and 1 stimulant drug, each at an

15
Page 17 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

285 environmentally relevant concentration (200 ppb), was used to evaluate and compare the PPCP

286 removal performance of the rGO-nZVI NH along with the parent nanomaterials under the same

Environmental Science: Water Research & Technology Accepted Manuscript


287 experimental conditions. Figure 3a and Table S1 present the results for PPCP removal efficiency

288 by nanomaterial based adsorption using rGO-nZVI NH or rGO or nZVI for 30 min. One of the
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

289 parent nanomaterials, nZVI, had minor adsorptive removal capacity, removing less than 15% of

290 the initial concentaration of all the PPCPs after 30 minutes. The other parent nanomaterial rGO,

291 removed significantly higher amount of PPCPs than nZVI; however, still had removal efficiency

292 within the range of 15-51% except for paroxetine (SSRI) with a removal efficiency of 71%. In the

293 case of rGO-nZVI NH, the PPCP removal efficiencies ranged from 74-100% after 30 minutes. The

294 NH showed more than 95% removal performance for six of the PPCPs including all 3 SSRIs

295 (citalopram, paroxetine, and fluoxetine), 2 NSAIDs (diclofenac and naproxen), and 1 anti-seizure

296 (lamotrigine). The lowest removal performance (74%) by the NH was observed for

297 sulfamethoxazole (antibiotic), but this removal performance by NH was still higher than removal

298 performance by the parent nanomaterials for any of the PPCPs including sulfamethoxazole.

299 Therefore, a significant increase (28% - 70%) in the adsorptive PPCP removal percentage was

300 observed for rGO-nZVI NHs compared to the parent nanomaterials rGO and nZVI. A Previous

301 study, which worked with bare graphene to remove a mixture of antibiotics at environmentally

302 relevant concentrations, showed similar removal efficiency (>89%) but with much higher dose of

303 the nano-adsorbent (1 mg/mL) i.e., double the dose used in our study and only afer a long

304 equilibrium time of upto 24 hours.52

305 The significantly better PPCP adsorptive removal performance by the rGO-nZVI NH,

306 compared to parent nanomaterials, can be attributed to increased adsorption sites resulting from

307 nZVI acting as a spacer resisting restacking of rGO nanosheets, whereas rGO simultaneously

16
Environmental Science: Water Research & Technology Page 18 of 40
View Article Online
DOI: 10.1039/D0EW00140F

308 resiting nZVI aggregation by grafting nZVI on its surface.69-71 In addition, the heterojunction at

309 metal-graphene interface can act as specific adsorption sites and contribute to the overall increase

Environmental Science: Water Research & Technology Accepted Manuscript


310 of adsorption capactity in the NH.72, 73 The rGO-nZVI NH, along with the parent nanomaterials,

311 exhibited different removal performance for different PPCPs in the matrix. This can be attributed
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

312 to the physicochemical properties of both the adsorbent and also the PPCPs. Mostly, the

313 hydrophobicity of the studied PPCPs appears to govern the adsorption process. In the case of

314 PPCPs with high log Kow value (>3, SSRIs: citalopram, paroxetine, fluoxetine; NSAIDs:

315 diclofenac and naproxen; Table 1), except ibuprofen, the rGO-nZVI NH was able to remove more

316 than 98% of the initial concentration, suggesting hydrophobic interaction a dictating adsorption

317 mechanism. The relatively lower removal of ibuprofen (89%) can be attributed to the presence of

318 a single aromatic ring and a hydrophilic carboxylic group, which may cause reduced π-π

319 interaction (compared to others), ultimately resulting in less competitive adsorption within the

320 PPCP matrix.74 The two least removed PPCPs, sulfamethoxazole (antibiotic, 74%) and

321 acetaminophen (analgesic, 83%), as expected, have significantly low log Kow value (0.9 and 0.5

322 respectively). In addition to low hydrophobicity, both of these PPCPs also remains predominantly

323 as neutral species in the experimental pH range (~3), further negating any possibility for

324 electrostatic intercation with charged adsorent surface.75-77

325 However, despite having the lowest log Kow value, caffeine showed significantly high

326 removal percentage (93%) when adsorbed by the rGO-nZVI NH. Electronegative nitrogen and

327 oxygen atoms present in caffeine induce polarity even in its neutral form (pKa ~ 10.4) causing

328 electrosataic interaction with rGO-nZVI NH.78, 79 Additionally, the presence of nitrogen atom in

329 aromatic ring also facilitates π- π electron donor-acceptor interaction with NH resulting in

330 significantly high removal of caffeine.31, 80 This suggests that other adsorption mechanisms (e.g.,

17
Page 19 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

331 electrostatic and π-π electron donor acceptor interaction, hydrogen bond), in addition to the

332 hydrophobic interaction, can contribute to PPCP removal by the rGO-nZVI NH. In the case of

Environmental Science: Water Research & Technology Accepted Manuscript


333 SSRIs i.e., for citalopram, paroxetine, and fluoxetine, along with their relatively high logKow value,

334 the presence of 3, 4, and 5 hydrogen bond acceptors, respectively, (identified with black dots in
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

335 respective chemical structures in Figure S3) can contribute to adsorptive removal due to the

336 formation of hydrogen bond with functional groups of rGO-nZVI NH.81-83 Additionally, these

337 three SSRIs – citalopram, paroxetine, and fluoxetine, unlike other PPCPs in the matrix, have 1, 1,

338 and 3 highly electronegative fluorine atoms, respectively (as shown in Figure S3), which facilitate

339 higher electrostatic interaction with NH compared to other PPCPs not containing fluorine,

340 ultimately resulting in high removal efficiency ( > 98%).84

341 3.2.2. AOP augmented adsorptive removal of PPCPs

342 To evaluate the synergistic adsorptive and AOP based removal of the PPCPs by the rGO-nZVI

343 NH and parent nanomaterials, similar removal tests were carried out in the presence of H2O2 (15.6

344 µL, 28-30%). Figure 3b and Table S2 presents the results for PPCP removal using this AOP

345 augmented adsorption. nZVI (present in the NH or bare) works as a heterogeneous Fenton catalyst

346 for activating H2O2 to generate reactive oxygen species (e.g., hydroxyl radical (•OH)), which can

347 further augment PPCP removal by oxidative degradation in addition to adsorption.32, 51 We found

348 that rGO with H2O2 performed PPCP removal in the range of 3-73%, with similar removal for

349 individual PPCPs compared to only rGO, confirming the absence of any Fenton catalytic activity.

350 Bare nZVI, as a heterogeneous Fenton catalyst in the presence of H2O2, showed significant

351 increase ( ≥ 10% removal) in removal only for ibuprofen and acetaminophen compared to without

352 addition of H2O2. The highest removal of ~22% with bare nZVI with H2O2 for acetaminophen can

353 be accounted only for oxidative degradation, as bare nZVI did not show any adsorptive removal

18
Environmental Science: Water Research & Technology Page 20 of 40
View Article Online
DOI: 10.1039/D0EW00140F

354 for acetaminophen. The rGO-nZVI NH showed the best PPCP removal performance for AOP

355 augmented adsorption process, with more than 93% removal for all the PPCPs, which ranged upto

Environmental Science: Water Research & Technology Accepted Manuscript


356 20% higher than only adsorptive removal of PPCPs by the rGO-nZVI NH (Figure S5). In the case

357 of sulfamethoxazole, which was least removed in case of only adsorption by rGO-nZVI NH, the
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

358 removal percentage increased by 20% in AOP augmented condition. This can be attributed to the

359 isoxazole ring of sulfamethoxazole structure, which is susceptible to degradation under •OH

360 attack.85 However, bare nZVI, as Fenton catalyst in presence of H2O2, could not improve

361 sulfamethoxazole removal compared to only adsorptive removal performed in absence of H2O2.

362 This indicates better Fenton catalytic activity of nZVI when hybridized on graphene support. The

363 other PPCPs in the mixture are also susceptible to •OH, which initiates degradation through

364 subsequent hydroxylation and decarboxylation reaction pathways.86-89 This resulted in enhanced

365 removal of PPCPs in AOP augmented condition by the rGO-nZVI NH. However, for bare nZVI,

366 the overall removal efficiency across all PPCPs was still less than 25% in AOP augmented

367 condition. The limited catalytic performance of bare nZVI compared to rGO-nZVI NH can be

368 attributed to aggregation, passive oxide layer formation and, moreover, absence of synergistic

369 adsorption by a support medium.90, 91

19
Page 21 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

Environmental Science: Water Research & Technology Accepted Manuscript


Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

370

371 Figure 3. (a) Adsorptive and (b) AOP augmented adsorptive removal of PPCPs by rGO-nZVI NH

372 and parent nanomaterials rGO and nZVI after 30 minutes.

373 3.3 PPCP removal kinetics for rGO-nZVI NHs and their parent nanomaterials

374 To determine the PPCP removal kinetics for parent nanomaterials and rGO-nZVI NH, a second

375 set of PPCP removal tests were carried out using a mixture of six PPCPs which included 3 SSRIs:

376 citalopram, paroxetine, and fluoxetine; 1 NSAID: diclofenac; 1 anti-seizure: carbamazepine; 1

377 antibiotic: sulfamethoxazole. This subset of PPCPs were selected from the first set, keeping

20
Environmental Science: Water Research & Technology Page 22 of 40
View Article Online
DOI: 10.1039/D0EW00140F

378 PPCPs from different classes and including the three SSRIs. To evaluate and compare the kinetic

379 performance, the removal of PPCPs after first 10 minutes by the nanomaterials are reported in

Environmental Science: Water Research & Technology Accepted Manuscript


380 Figure 4 (also in Table S2). These confirm the faster removal kinetics for NH compared to the

381 parent nanomaterials for both adsorptive (Figure 4a) and AOP augmented adsorptive (Figure 4b)
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

382 removal. Furthermore, a significant increase in removal efficiency (27-58 %) across all the 6

383 PPCPs at 10 minutes can be observed for rGO- nZVI NH with H2O2 compared to only rGO-nZVI

384 NH. For the adsorptive removal of 6 PPCPs without H2O2, 52-77% of the total removal occurred

385 in the first 10 minutes, whereas the removal percentage was 94-100% in case of AOP augmented

386 adsorption condition after 10 minutes. This confirms, in AOP augmented condition, synergistic

387 adsorption and advanced oxidation is taking place simultaneously, which is enabling faster

388 removal kinetics for the rGO-nZVI NH compared to only adsorptive removal. As far as the

389 removal efficiency is concerned after 30 minutes, for both adsorptive and AOP augmented

390 adsorptive removal, rGO-nZVI NH performed significantly better than the parent nanomaterials

391 (Figure S6 (a) and (b)) which is similar to the results for the first set of PPCPs (12 PPCPs).

392 However, for this smaller set of 6 PPCPs, the parent nanomaterials seem to perform better

393 compared to their performance for the larger set of 12 PPCPs due to less competition for removal.

21
Page 23 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

Environmental Science: Water Research & Technology Accepted Manuscript


Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

394

395 Figure 4. (a) Adsorptive and (b) AOP augmented adsorptive removal of PPCPs by rGO-nZVI NH

396 and parent nanomaterials rGO and nZVI after 10 minutes.

397 To further evaluate the time dependent performance of the nanomaterials, the PPCP

398 removal was reported through kinetic plots at 5 minute time intervals over 30 minutes (Figures 5

399 and S7). The kinetic plots for two PPCPs, fluoxetine (a representative of SSRI group) and

400 sulfamethoxazole (antibiotic, which had the least removal in the first set of experiments) are

401 reported in Figure 5 and the rest are in the electronic supplementary information (Figure S7). Each

402 row in these figures presents removal kinetics results for one specific PPCP. In each row of Figures

403 5 and S7, the left and middle figures present the results of PPCP removal kinetics, respecitively,

404 by adsorptive (only) and AOP augmented adsorption for all three nanomaterials i.e., rGO, nZVI,

405 and rGO-nZVI NH. The right figure in each row presents the comparison between PPCP removal

406 kinetics by adsorption (only) and AOP augmented adsorption for the rGO-nZVI NH. For all the

407 6 PPCPs, parent nanomaterials showed slower kinetics compared to the rGO-nZVI NH for both,

408 adsorptive and AOP augmented adsorptive condition, over the entire period of 30 minutes. For

409 example, under adsorptive (only) condition, parent nanomaterial rGO could remove ~85% of

22
Environmental Science: Water Research & Technology Page 24 of 40
View Article Online
DOI: 10.1039/D0EW00140F

410 paroxetine at the end of 30 minutes which was relatively close to the removal performance by NH

411 (~100%). However, the kinetics was much faster for the rGO-nZVI NH, as the removal by NH

Environmental Science: Water Research & Technology Accepted Manuscript


412 was 75% in the first 5 minutes, whereas that by rGO was only 38%. In case of AOP augmented

413 condition, paroxetine and diclofenac seem to be more prone to oxidative degradation, as
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

414 significantly high removal of these two PPCPs was possible by bare nZVI (74% and 99%

415 respectively), although less compared to the rGO-nZVI NH (~100%). However, NH with H2O2

416 still exhibited faster removal kinetics for these two PPCPs compared to bare nZVI, with all the

417 100% removal in the first 5 minutes compared to 27% and 21% removal respectively by nZVI.

418

419 Figure 5. Kinetics of (a, d) adsorptive (only) and (b, e) AOP augmented adsorptive removal of

420 PPCPs by parent nanomaterials and rGO-nZVI NH; (c, f) Comparison of PPCP removal by rGO-

421 nZVI NH via adsorption (only) and AOP augmented adsorption for fluoxetine and

422 sulfamethoxazole.

423 To further quantify the removal kinetics, the rate constants for individual PPCP removal

424 by both the adsorption (k1) and AOP augmented adsorption (k2) by rGO-nZVI NH were estimated

23
Page 25 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

425 considering both processes as pseudo-first order reaction since the R2 values suggested better

426 conformity to pseudo-first order reaction than any other reaction order. However, deviation from

Environmental Science: Water Research & Technology Accepted Manuscript


427 exact conformity to pseudo-first order reaction rate constant (i.e., low R2 values) in case of few

428 PPCPs can be attributed to various reasons including the presence of multiple PPCPs in the
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

429 mixture.52, 92-94 Moreover, the adsorbate sample collection, processing, and analyses within an

430 interval shorter than 5 minutes was significantly challenging (for multiple NH-PPCP

431 combinations) which may have contributed to the low R2 values as well. The rate constants along

432 with the R2 values are shown in Table 2 and Figure S8. The reaction rate constants for AOP

433 augmented adsorption by rGO-nZVI NH was ~3-5.5 times higher than that for only adsorption for

434 individual PPCPs (Table 2). This can be attributed to the effect of heterogeneous Fenton catalysis,

435 which further illustrates the merit of synergistic AOP augmented adsorptive removal of PPCPs.

436

437 Table 2. Rate constants for individual PPCP removal by adsorption (red) and AOP augmented

438 adsorption for rGO-nZVI NH.

Citalopram Paroxetine Fluoxetine Diclofenac Carbamazepine Sulfamethoxazole


Pseudo First Order
Rate Constant, 0.16 0.21 0.16 0.17 0.09 0.06
kNH (min-1)
R2 0.93 0.96 0.98 0.98 0.97 0.99
Pseudo First Order
Rate Constant, 0.88 0.87 0.61 0.51 0.35 0.27
kNH + H2O2 (min-1)
R2 0.79 0.88 0.27 0.98 0.85 0.98
kNH + H2O2 / KNH 5.50 4.14 3.81 3.00 3.89 4.50
439

440

24
Environmental Science: Water Research & Technology Page 26 of 40
View Article Online
DOI: 10.1039/D0EW00140F

441 3.4 Mechanism of AOP augmented adsorption: H2O2 activation

442 To better comprehend the enhanced performance of rGO-nZVI NH in AOP augmented condition,

Environmental Science: Water Research & Technology Accepted Manuscript


443 we quantified residual H2O2 concentration with time in similar batch test in presence of all these

444 nanomaterials individually to analyze their catalytic property (Figure 6). AOP based contaminant
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

445 degradation (including PPCPs) are known to be caused by reactive oxygen species (ROS)

446 generated through activation of H2O2 in the presence of homo- or heterogeneous catalysts.32, 95, 96

447 These ROS may include super oxide radical anion (O2·-), hydroperoxyl radical (HO2·), and ·OH

448 radical.97, 98 We found that in the case of control (without any nanomaterial) and rGO, there was

449 no H2O2 decay, suggesting no catalytic activation of H2O2 to form various ROS. However, both

450 nZVI and rGO-nZVI NH showed H2O2 decay over time confirming the presence of catalytic

451 activation of H2O2. While the rGO-nZVI NH showed a high H2O2 decay (82%) over 30 minutes

452 with the fastest kinetics, nZVI showed a steady decay of H2O2 reaching 49% decay at the end of

453 30 minutes. This suggests that the rGO-nZVI NH could perform catalytic oxidation much faster

454 and better than the parent nanomaterials which may have resulted in better and faster PPCP

455 removal by the rGO-nZVI NH mediated AOP augmented adsorption, as shown above.

456

457 Figure 6. H2O2 decay over time by rGO-nZVI NH and parent nanomaterials rGO and nZVI.

25
Page 27 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

458 It is also important to note that both the adsorption and AOP processes in this study was

459 performed at pH 3. We did this because pH 3 has been described as the most favorable operating

Environmental Science: Water Research & Technology Accepted Manuscript


460 conditions for both homo- and heterogeneous Fenton reactions in the literature.32, 56, 99 Most

461 previous studies involving nZVI as heterogeneous Fenton catalyst have shown better catalytic
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

462 performance at pH 3 than at neutral and alkaline pH.32, 56, 99 There are several reasons for such

463 behavior of nZVI catalyst. Firstly, at low pH, Fe0 at the nZVI surface reacts with H2O2 or O2

464 present in water via a two-electron transfer process to continuously generate surface bound Fe2+

465 ions, which further reacts with H2O2 to generate ROS and Fe3+.99-102 Moreover, the nZVI core

466 (Fe0) can transfer electron to the oxidized Fe3+ ions at the nZVI surface to repeat the cycle of Fe2+

467 generation on the surface and can continuously generate ROS.51, 103 Overall, this leads to relatively

468 higher H2O2 activation and more ROS formation by nZVI based catalysts at acidic pH (as shown

469 in the following equations) than at alkaline pH.

470 Fe0 + H2O2 + 2H+ → Fe2+ + 2H2O

471 Fe0 + O2 + 2H+ → Fe2+ + 2H2O2

472 Fe2+ + H2O2 → Fe3+ + •OH + OH−

473 2 Fe3+ + Fe0 → 3Fe2+

474 Moreover, at alkaline pH, the Fe2+/Fe3+ ions generated from nZVI convert into passive iron

475 oxides (e.g., Fe(OH)3, FeOOH) either at nZVI surface or as precipitates.100, 104 Acidic pH inhibits

476 such nZVI passivation which in turn resulting in better performance of nZVI catalyst. This

477 phenomenon can also be observed in case of homogenous Fenton processes (that involves Fe2+

478 ions), making acidic pH (̴ 2.8-3.2) favorable operating condition.105-107

26
Environmental Science: Water Research & Technology Page 28 of 40
View Article Online
DOI: 10.1039/D0EW00140F

479 3.5 Hybridization induced enhancement in PPCP removal performance of rGO-nZVI

480 NH

Environmental Science: Water Research & Technology Accepted Manuscript


481 To further elucidate the synergistic effect of hybridization, similar removal tests with mixture of

482 parent nanomaterials were carried out. The mixture contained 1.9 mg of rGO and 6.1 mg of nZVI
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

483 to match the ratio (~1:3.2) of these nanomaterials in the synthesized rGO-nZVI NH. This mixture

484 of rGO and nZVI were also tested for removal of the same 6 PPCPs in the presence and absence

485 of H2O2 and compared with the performance of equivalent dose of rGO-nZVI NH (Figure 7).

486 Overall, the nanomaterial mixture showed limited PPCP removal when compared to the removal

487 performance by the rGO-nZVI NH. In case of solely adsorptive removal of these PPCPs, removal

488 percentage by rGO-nZVI NH was 66- 88% higher than the removal by the nanomaterial mixture.

489 Similarly, for AOP augmented adsorptive removal, the difference in removal performance was

490 56%, 57%, 60%, and 54% in favor of the rGO-nZVI NH for citalopram, fluoxetine,

491 carbamazepine, and sulfamethoxazole, respectively. However, for paroxetine and diclofenac, the

492 mixture with H2O2 itself can remove 87% and 98% respectively comparable to the ~100% removal

493 by the rGO-nZVI NH. These two PPCPs are highly susceptible to oxidative degradation as already

494 observed in the removal test with bare nZVI with H2O2 (Figure S7), removing 75% and 99% of

495 the initial concentration respectively. This indicates the nZVI alone can be attributed to the high

496 removal of paroxetine and diclofenac by the parent nanomaterial mixture.

27
Page 29 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

Environmental Science: Water Research & Technology Accepted Manuscript


Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

497

498 Figure 7. Comparative PPCP removal by rGO-nZVI NH and mixture of parent nanomaterials rGO

499 and nZVI.

500 Kinetic tests were also carried out with the mixture of parent nanomaterials. The kinetic

501 test further showed that the hybridization triggered faster PPCP removal kinetics than the mixture

502 of parent nanomaterials. The kinetic plots for fluoxetine and sulfamethoxazole are reported in

503 Figure 8, the rest are shown in Figure S9.

28
Environmental Science: Water Research & Technology Page 30 of 40
View Article Online
DOI: 10.1039/D0EW00140F

Environmental Science: Water Research & Technology Accepted Manuscript


Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

504

505 Figure 8. Comparative kinetics of (a, c) adsorptive, (b, d) AOP augmented adsorptive removal by

506 rGO-nZVI NH and mixture of rGO and nZVI for fluoxetine and sulfamethoxazole.

507 4. Conclusion

508 In this study, we used rGO-nZVI NH to remove a complex matrix of PPCPs including novel

509 antidepressants that are found in trace (ppb level) concentrations within wastewater treatment

510 plants. We compared the ability of rGO-nZVI NH to enhance PPCP removal efficiencies and

511 kinetics via two water treatment methods, i.e., adsorption and AOP. We found that rGO-nZVI NH

512 can significantly enhance the PPCP removal via these processes when compared to its parent

513 materials rGO and nZVI. Moreover, synergistic AOP augmented adsorption process showed much

514 better (more than 93% removal efficiency) and faster PPCP removal for almost all of the PPCPs

515 than only adsorption process when rGO-nZVI NH was used. We also found that the differences

29
Page 31 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

516 in removal efficiencies between individual PPCPs within the mixture was caused by their

517 competitive adsorption that depended upon their individual chemical nature and differences

Environmental Science: Water Research & Technology Accepted Manuscript


518 between hydrophobic and electrostatic interactions with the rGO-nZVI NHs. The results of this

519 study are very encouraging and have important implications for de facto potable water reuse,
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

520 majorly within an urban/industrial water infrastructure. The high efficiency and fast kinetics for

521 PPCP removal reveals the potential of utilizing rGO-nZVI NH at point of high concentration

522 emission (e.g., hospital, pharmaceutical industry effluent) and even at point of use/entry at the user

523 end to mitigate the huge risk associated with PPCP contaminated water. Moreover, rGO-nZVI

524 NH mediated AOP can be included within the already existing municipal wastewater treatment

525 train possibly right before the biological wastewater treatment processes, since rGO-nZVI NH has

526 the potential to breakdown the recalcitrant PPCPs, through heterogeneous Fenton catalysis, to non-

527 aromatic small-chain organic acids108 which can be further mineralized by microbes. However,

528 for achieving this, identifying the PPCP degradation products for rGO-nZVI NH based AOP and

529 understanding associated by-product toxicity to aquatic organisms and microbes will be important.

530 These product identification and toxicity studies can only be performed using individual PPCPs

531 instead of using PPCP mixture due to technical challenges associated with correlating and

532 connecting degradation products with their respective source PPCPs within a mixture.

533 Conflict of Interests

534 The authors declare no conflict of interest.

535 References

536 1. Gros, M.; Rodríguez-Mozaz, S.; Barceló, D., Fast and comprehensive multi-residue
537 analysis of a broad range of human and veterinary pharmaceuticals and some of their metabolites
538 in surface and treated waters by ultra-high-performance liquid chromatography coupled to

30
Environmental Science: Water Research & Technology Page 32 of 40
View Article Online
DOI: 10.1039/D0EW00140F

539 quadrupole-linear ion trap tandem mass spectrometry. Journal of Chromatography A 2012, 1248,
540 104-121.
541 2. Collado, N.; Rodriguez-Mozaz, S.; Gros, M.; Rubirola, A.; Barceló, D.; Comas, J.;

Environmental Science: Water Research & Technology Accepted Manuscript


542 Rodriguez-Roda, I.; Buttiglieri, G., Pharmaceuticals occurrence in a WWTP with significant
543 industrial contribution and its input into the river system. Environmental pollution 2014, 185,
544 202-212.
545 3. Schultz, M. M.; Furlong, E. T.; Kolpin, D. W.; Werner, S. L.; Schoenfuss, H. L.; Barber,
546 L. B.; Blazer, V. S.; Norris, D. O.; Vajda, A. M., Antidepressant pharmaceuticals in two US
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

547 effluent-impacted streams: occurrence and fate in water and sediment, and selective uptake in
548 fish neural tissue. Environmental science & technology 2010, 44, (6), 1918-1925.
549 4. Schultz, M. M.; Furlong, E. T., Trace analysis of antidepressant pharmaceuticals and
550 their select degradates in aquatic matrixes by LC/ESI/MS/MS. Analytical chemistry 2008, 80,
551 (5), 1756-1762.
552 5. Boxall, A. B.; Rudd, M. A.; Brooks, B. W.; Caldwell, D. J.; Choi, K.; Hickmann, S.;
553 Innes, E.; Ostapyk, K.; Staveley, J. P.; Verslycke, T.; Ankley, G. T.; Beazley, K. F.; Belanger, S.
554 E.; Berninger, J. P.; Carriquiriborde, P.; Coors, A.; Deleo, P. C.; Dyer, S. D.; Ericson, J. F.;
555 Gagne, F.; Giesy, J. P.; Gouin, T.; Hallstrom, L.; Karlsson, M. V.; Larsson, D. G.; Lazorchak, J.
556 M.; Mastrocco, F.; McLaughlin, A.; McMaster, M. E.; Meyerhoff, R. D.; Moore, R.; Parrott, J.
557 L.; Snape, J. R.; Murray-Smith, R.; Servos, M. R.; Sibley, P. K.; Straub, J. O.; Szabo, N. D.;
558 Topp, E.; Tetreault, G. R.; Trudeau, V. L.; Van Der Kraak, G., Pharmaceuticals and personal
559 care products in the environment: what are the big questions? Environ Health Perspect 2012,
560 120, (9), 1221-9.
561 6. Blair, B. D.; Crago, J. P.; Hedman, C. J.; Klaper, R. D., Pharmaceuticals and personal
562 care products found in the Great Lakes above concentrations of environmental concern.
563 Chemosphere 2013, 93, (9), 2116-2123.
564 7. Miller, K. J.; Meek, J., Helena Valley ground water: pharmaceuticals, personal care
565 products, endocrine disruptors (PPCPs), and microbial indicators of fecal contamination.
566 Citeseer: 2006.
567 8. Arpin-Pont, L.; Bueno, M. J. M.; Gomez, E.; Fenet, H., Occurrence of PPCPs in the
568 marine environment: a review. Environmental Science and Pollution Research 2016, 23, (6),
569 4978-4991.
570 9. Arnnok, P.; Singh, R. R.; Burakham, R.; Pérez-Fuentetaja, A.; Aga, D. S., Selective
571 uptake and bioaccumulation of antidepressants in fish from effluent-impacted Niagara River.
572 Environmental science & technology 2017, 51, (18), 10652-10662.
573 10. Daughton, C. G.; Ternes, T. A., Pharmaceuticals and personal care products in the
574 environment: Agents of subtle change? Environmental Health Perspectives 1999, 107, 907-938.
575 11. Cizmas, L.; Sharma, V. K.; Gray, C. M.; McDonald, T. J., Pharmaceuticals and personal
576 care products in waters: occurrence, toxicity, and risk. Environmental Chemistry Letters 2015,
577 13, (4), 381-394.
578 12. Hernando, M. D.; Mezcua, M.; Fernandez-Alba, A. R.; Barcelo, D., Environmental risk
579 assessment of pharmaceutical residues in wastewater effluents, surface waters and sediments.
580 Talanta 2006, 69, (2), 334-342.
581 13. Silva, L. J. G.; Lino, C. M.; Meisel, L. M.; Pena, A., Selective serotonin re-uptake
582 inhibitors (SSRIs) in the aquatic environment: An ecopharmacovigilance approach. Science of
583 The Total Environment 2012, 437, 185-195.

31
Page 33 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

584 14. Péry, A.; Gust, M.; Vollat, B.; Mons, R.; Ramil, M.; Fink, G.; Ternes, T.; Garric, J.,
585 Fluoxetine effects assessment on the life cycle of aquatic invertebrates. Chemosphere 2008, 73,
586 (3), 300-304.

Environmental Science: Water Research & Technology Accepted Manuscript


587 15. Lazzara, R.; Blázquez, M.; Porte, C.; Barata, C., Low environmental levels of fluoxetine
588 induce spawning and changes in endogenous estradiol levels in the zebra mussel Dreissena
589 polymorpha. Aquatic toxicology 2012, 106, 123-130.
590 16. Weinberger II, J.; Klaper, R., Environmental concentrations of the selective serotonin
591 reuptake inhibitor fluoxetine impact specific behaviors involved in reproduction, feeding and
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

592 predator avoidance in the fish Pimephales promelas (fathead minnow). Aquatic toxicology 2014,
593 151, 77-83.
594 17. Yu, Y.; Huang, Q.; Wang, Z.; Zhang, K.; Tang, C.; Cui, J.; Feng, J.; Peng, X.,
595 Occurrence and behavior of pharmaceuticals, steroid hormones, and endocrine-disrupting
596 personal care products in wastewater and the recipient river water of the Pearl River Delta, South
597 China. Journal of Environmental Monitoring 2011, 13, (4), 871-878.
598 18. Archer, E.; Wolfaardt, G. M.; van Wyk, J. H., Pharmaceutical and personal care products
599 (PPCPs) as endocrine disrupting contaminants (EDCs) in South African surface waters. Water Sa
600 2017, 43, (4), 684-706.
601 19. Stackelberg, P. E.; Furlong, E. T.; Meyer, M. T.; Zaugg, S. D.; Henderson, A. K.;
602 Reissman, D. B., Persistence of pharmaceutical compounds and other organic wastewater
603 contaminants in a conventional drinking-water-treatment plant. Science of The Total
604 Environment 2004, 329, (1), 99-113.
605 20. Vieno, N. M.; Härkki, H.; Tuhkanen, T.; Kronberg, L., Occurrence of pharmaceuticals in
606 river water and their elimination in a pilot-scale drinking water treatment plant. Environmental
607 science & technology 2007, 41, (14), 5077-5084.
608 21. Lin, T.; Yu, S.; Chen, W., Occurrence, removal and risk assessment of pharmaceutical
609 and personal care products (PPCPs) in an advanced drinking water treatment plant (ADWTP)
610 around Taihu Lake in China. Chemosphere 2016, 152, 1-9.
611 22. Xu, Y.; Liu, T.; Zhang, Y.; Ge, F.; Steel, R. M.; Sun, L., Advances in technologies for
612 pharmaceuticals and personal care products removal. Journal of Materials Chemistry A 2017, 5,
613 (24), 12001-12014.
614 23. Yang, Y.; Ok, Y. S.; Kim, K.-H.; Kwon, E. E.; Tsang, Y. F., Occurrences and removal of
615 pharmaceuticals and personal care products (PPCPs) in drinking water and water/sewage
616 treatment plants: A review. Science of the Total Environment 2017, 596, 303-320.
617 24. García-Galán, M. J.; Anfruns, A.; Gonzalez-Olmos, R.; Rodríguez-Mozaz, S.; Comas, J.,
618 UV/H2O2degradation of the antidepressants venlafaxine and O-desmethylvenlafaxine:
619 elucidation of their transformation pathway and environmental fate. Journal of hazardous
620 materials 2016, 311, 70-80.
621 25. Bhadra, B. N.; Seo, P. W.; Jhung, S. H., Adsorption of diclofenac sodium from water
622 using oxidized activated carbon. Chemical Engineering Journal 2016, 301, 27-34.
623 26. Freyria, F.; Geobaldo, F.; Bonelli, B., Nanomaterials for the abatement of
624 pharmaceuticals and personal care products from wastewater. Applied Sciences 2018, 8, (2), 170.
625 27. Cincinelli, A.; Martellini, T.; Coppini, E.; Fibbi, D.; Katsoyiannis, A., Nanotechnologies
626 for Removal of Pharmaceuticals and Personal Care Products from Water and Wastewater. A
627 Review. Journal of Nanoscience and Nanotechnology 2015, 15, (5), 3333-3347.

32
Environmental Science: Water Research & Technology Page 34 of 40
View Article Online
DOI: 10.1039/D0EW00140F

628 28. Cai, Z.; Dwivedi, A. D.; Lee, W.-N.; Zhao, X.; Liu, W.; Sillanpää, M.; Zhao, D.; Huang,
629 C.-H.; Fu, J., Application of nanotechnologies for removing pharmaceutically active compounds
630 from water: development and future trends. Environmental Science: Nano 2018, 5, (1), 27-47.

Environmental Science: Water Research & Technology Accepted Manuscript


631 29. Liu, F.-f.; Zhao, J.; Wang, S.; Du, P.; Xing, B., Effects of solution chemistry on
632 adsorption of selected pharmaceuticals and personal care products (PPCPs) by graphenes and
633 carbon nanotubes. Environmental science & technology 2014, 48, (22), 13197-13206.
634 30. Gao, Y.; Li, Y.; Zhang, L.; Huang, H.; Hu, J.; Shah, S. M.; Su, X., Adsorption and
635 removal of tetracycline antibiotics from aqueous solution by graphene oxide. Journal of colloid
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

636 and interface science 2012, 368, (1), 540-546.


637 31. Al-Khateeb, L. A.; Almotiry, S.; Salam, M. A., Adsorption of pharmaceutical pollutants
638 onto graphene nanoplatelets. Chemical Engineering Journal 2014, 248, 191-199.
639 32. Zha, S.; Cheng, Y.; Gao, Y.; Chen, Z.; Megharaj, M.; Naidu, R., Nanoscale zero-valent
640 iron as a catalyst for heterogeneous Fenton oxidation of amoxicillin. Chemical Engineering
641 Journal 2014, 255, 141-148.
642 33. Mondal, S. K.; Saha, A. K.; Sinha, A., Removal of ciprofloxacin using modified
643 advanced oxidation processes: kinetics, pathways and process optimization. Journal of Cleaner
644 Production 2018, 171, 1203-1214.
645 34. Lefevre, E.; Bossa, N.; Wiesner, M. R.; Gunsch, C. K., A review of the environmental
646 implications of in situ remediation by nanoscale zero valent iron (nZVI): behavior, transport and
647 impacts on microbial communities. Science of The Total Environment 2016, 565, 889-901.
648 35. Masud, A.; Cui, Y.; Atkinson, J. D.; Aich, N., Shape matters: Cr (VI) removal using iron
649 nanoparticle impregnated 1-D vs 2-D carbon nanohybrids prepared by ultrasonic spray pyrolysis.
650 Journal of Nanoparticle Research 2018, 20, (3), 64.
651 36. Mehrabi, N.; Masud, A.; Afolabi, M.; Hwang, J.; Ortiz, G. A. C.; Aich, N., Magnetic
652 graphene oxide-nano zero valent iron (GO–nZVI) nanohybrids synthesized using biocompatible
653 cross-linkers for methylene blue removal. RSC advances 2019, 9, (2), 963-973.
654 37. Rostamian, R.; Behnejad, H., A comparative adsorption study of sulfamethoxazole onto
655 graphene and graphene oxide nanosheets through equilibrium, kinetic and thermodynamic
656 modeling. Process Safety and Environmental Protection 2016, 102, 20-29.
657 38. Wang, P.; Zhou, X.; Zhang, Y.; Yang, L.; Zhi, K.; Wang, L.; Zhang, L.; Guo, X.,
658 Unveiling the mechanism of electron transfer facilitated regeneration of active Fe 2+ by nano-
659 dispersed iron/graphene catalyst for phenol removal. RSC Advances 2017, 7, (43), 26983-26991.
660 39. Yang, B.; Tian, Z.; Zhang, L.; Guo, Y.; Yan, S., Enhanced heterogeneous Fenton
661 degradation of Methylene Blue by nanoscale zero valent iron (nZVI) assembled on magnetic
662 Fe3O4/reduced graphene oxide. Journal of Water Process Engineering 2015, 5, 101-111.
663 40. Wan, Z.; Wang, J. L., Degradation of sulfamethazine using Fe3O4-Mn3O4/reduced
664 graphene oxide hybrid as Fenton-like catalyst. Journal of Hazardous Materials 2017, 324, 653-
665 664.
666 41. Wan, Z.; Hu, J.; Wang, J. L., Removal of sulfamethazine antibiotics using Ce-Fe-
667 graphene nanocomposite as catalyst by Fenton-like process. Journal of Environmental
668 Management 2016, 182, 284-291.
669 42. Shan, D.; Deng, S.; Jiang, C.; Chen, Y.; Wang, B.; Wang, Y.; Huang, J.; Yu, G.;
670 Wiesner, M. R., Hydrophilic and strengthened 3D reduced graphene oxide/nano-Fe 3 O 4 hybrid
671 hydrogel for enhanced adsorption and catalytic oxidation of typical pharmaceuticals.
672 Environmental Science: Nano 2018, 5, (7), 1650-1660.

33
Page 35 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

673 43. Zhuang, Y.; Liu, Q.; Kong, Y.; Shen, C.; Hao, H.; Dionysiou, D. D.; Shi, B., Enhanced
674 antibiotic removal through a dual-reaction-center Fenton-like process in 3D graphene based
675 hydrogels. Environmental Science: Nano 2019, 6, (2), 388-398.

Environmental Science: Water Research & Technology Accepted Manuscript


676 44. He, J.; Yang, X.; Men, B.; Wang, D., Interfacial mechanisms of heterogeneous Fenton
677 reactions catalyzed by iron-based materials: A review. Journal of environmental sciences 2016,
678 39, 97-109.
679 45. Ghadim, E. E.; Manouchehri, F.; Soleimani, G.; Hosseini, H.; Kimiagar, S.; Nafisi, S.,
680 Adsorption properties of tetracycline onto graphene oxide: equilibrium, kinetic and
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

681 thermodynamic studies. PLoS One 2013, 8, (11), e79254.


682 46. Chen, H.; Gao, B.; Li, H., Removal of sulfamethoxazole and ciprofloxacin from aqueous
683 solutions by graphene oxide. Journal of hazardous materials 2015, 282, 201-207.
684 47. Liu, F.-F.; Zhao, J.; Wang, S.; Xing, B., Adsorption of sulfonamides on reduced graphene
685 oxides as affected by pH and dissolved organic matter. Environmental pollution 2016, 210, 85-
686 93.
687 48. Khan, A.; Wang, J.; Li, J.; Wang, X.; Chen, Z.; Alsaedi, A.; Hayat, T.; Chen, Y.; Wang,
688 X., The role of graphene oxide and graphene oxide-based nanomaterials in the removal of
689 pharmaceuticals from aqueous media: a review. Environmental Science and Pollution Research
690 2017, 24, (9), 7938-7958.
691 49. Zhuang, S.; Zhu, X.; Wang, J., Kinetic, equilibrium, and thermodynamic performance of
692 sulfonamides adsorption onto graphene. Environmental Science and Pollution Research 2018,
693 25, (36), 36615-36623.
694 50. Xia, S.; Gu, Z.; Zhang, Z.; Zhang, J.; Hermanowicz, S. W., Removal of chloramphenicol
695 from aqueous solution by nanoscale zero-valent iron particles. Chemical Engineering Journal
696 2014, 257, 98-104.
697 51. Wang, L.; Yang, J.; Li, Y.; Lv, J.; Zou, J., Removal of chlorpheniramine in a nanoscale
698 zero-valent iron induced heterogeneous Fenton system: Influencing factors and degradation
699 intermediates. Chemical Engineering Journal 2016, 284, 1058-1067.
700 52. Peng, B.; Chen, L.; Que, C.; Yang, K.; Deng, F.; Deng, X.; Shi, G.; Xu, G.; Wu, M.,
701 Adsorption of antibiotics on graphene and biochar in aqueous solutions induced by π-π
702 interactions. Scientific reports 2016, 6, 31920.
703 53. Wang, C.; Li, H.; Liao, S.; Zheng, H.; Wang, Z.; Pan, B.; Xing, B., Coadsorption,
704 desorption hysteresis and sorption thermodynamics of sulfamethoxazole and carbamazepine on
705 graphene oxide and graphite. Carbon 2013, 65, 243-251.
706 54. Chang, E.-E.; Wan, J.-C.; Kim, H.; Liang, C.-H.; Dai, Y.-D.; Chiang, P.-C., Adsorption
707 of selected pharmaceutical compounds onto activated carbon in dilute aqueous solutions
708 exemplified by acetaminophen, diclofenac, and sulfamethoxazole. The Scientific World Journal
709 2015, 2015.
710 55. Jabeen, H.; Kemp, K. C.; Chandra, V., Synthesis of nano zerovalent iron nanoparticles–
711 graphene composite for the treatment of lead contaminated water. Journal of environmental
712 management 2013, 130, 429-435.
713 56. Lin, C.-C.; Hsu, S.-T., Performance of nZVI/H2O2 process in degrading polyvinyl
714 alcohol in aqueous solutions. Separation and Purification Technology 2018, 203, 111-116.
715 57. Xu, L.; Wang, J., Fenton-like degradation of 2, 4-dichlorophenol using Fe3O4 magnetic
716 nanoparticles. Applied Catalysis B: Environmental 2012, 123, 117-126.

34
Environmental Science: Water Research & Technology Page 36 of 40
View Article Online
DOI: 10.1039/D0EW00140F

717 58. Xu, L.; Wang, J., Degradation of 4-chloro-3, 5-dimethylphenol by a heterogeneous
718 fenton-like reaction using nanoscale zero-valent iron catalysts. Environmental engineering
719 science 2013, 30, (6), 294-301.

Environmental Science: Water Research & Technology Accepted Manuscript


720 59. Hollman, J.; Dominic, J. A.; Achari, G., Degradation of pharmaceutical mixtures in
721 aqueous solutions using UV/peracetic acid process: Kinetics, degradation pathways and
722 comparison with UV/H2O2. Chemosphere 2020, 125911.
723 60. Mullen, R. A.; Wigginton, K. R.; Noe-Hays, A.; Nace, K.; Love, N. G.; Bott, C. B.; Aga,
724 D. S., Optimizing extraction and analysis of pharmaceuticals in human urine, struvite, food
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

725 crops, soil, and lysimeter water by liquid chromatography-tandem mass spectrometry. Analytical
726 Methods 2017, 9, (41), 5952-5962.
727 61. Wang, Q.; Masud, A.; Aich, N.; Wu, Y., In vitro pulmonary toxicity of reduced graphene
728 oxide-nano zero valent iron nanohybrids and comparison with parent nanomaterial attributes.
729 ACS Sustainable Chemistry & Engineering 2018, 6, (10), 12797-12806.
730 62. Bagheri, M.; Jafari, S. M.; Eikani, M. H., Development of ternary nanoadsorbent
731 composites of graphene oxide, activated carbon, and zero‐valent iron nanoparticles for food
732 applications. Food Science & Nutrition 2019.
733 63. Park, O. K.; Choi, Y. M.; Hwang, J. Y.; Yang, C. M.; Kim, T. W.; You, N. H.; Koo, H.
734 Y.; Lee, J. H.; Ku, B. C.; Goh, M., Defect healing of reduced graphene oxide via intramolecular
735 cross-dehydrogenative coupling. Nanotechnology 2013, 24, (18).
736 64. Zhu, T.; Chen, J. S.; Lou, X. W., Glucose-Assisted One-Pot Synthesis of FeOOH
737 Nanorods and Their Transformation to Fe3O4@Carbon Nanorods for Application in Lithium Ion
738 Batteries. The Journal of Physical Chemistry C 2011, 115, (19), 9814-9820.
739 65. Lysenko, E. N.; Surzhikov, A. P.; Zhuravkov, S. P.; Vlasov, V. A.; Pustovalov, A. V.;
740 Yavorovsky, N. A., The oxidation kinetics study of ultrafine iron powders by thermogravimetric
741 analysis. Journal of Thermal Analysis and Calorimetry 2014, 115, (2), 1447-1452.
742 66. Fan, M.; Li, T.; Hu, J.; Cao, R.; Wei, X.; Shi, X.; Ruan, W., Artificial neural network
743 modeling and genetic algorithm optimization for cadmium removal from aqueous solutions by
744 reduced graphene oxide-supported nanoscale zero-valent iron (nZVI/rGO) composites. Materials
745 2017, 10, (5), 544.
746 67. Dong, H.; Zhao, F.; He, Q.; Xie, Y.; Zeng, Y.; Zhang, L.; Tang, L.; Zeng, G.,
747 Physicochemical transformation of carboxymethyl cellulose-coated zero-valent iron
748 nanoparticles (nZVI) in simulated groundwater under anaerobic conditions. Separation and
749 Purification Technology 2017, 175, 376-383.
750 68. Li, T.; Wang, T.; Qu, G.; Liang, D.; Hu, S., Synthesis and photocatalytic performance of
751 reduced graphene oxide–TiO 2 nanocomposites for orange II degradation under UV light
752 irradiation. Environmental Science and Pollution Research 2017, 24, (13), 12416-12425.
753 69. Jang, B.; Chae, O. B.; Park, S.-K.; Ha, J.; Oh, S. M.; Na, H. B.; Piao, Y., Solventless
754 synthesis of an iron-oxide/graphene nanocomposite and its application as an anode in high-rate
755 Li-ion batteries. Journal of Materials Chemistry A 2013, 1, (48), 15442-15446.
756 70. Peik-See, T.; Pandikumar, A.; Ngee, L. H.; Ming, H. N.; Hua, C. C., Magnetically
757 separable reduced graphene oxide/iron oxide nanocomposite materials for environmental
758 remediation. Catalysis Science & Technology 2014, 4, (12), 4396-4405.
759 71. Benjwal, P.; Kumar, M.; Chamoli, P.; Kar, K. K., Enhanced photocatalytic degradation of
760 methylene blue and adsorption of arsenic (iii) by reduced graphene oxide (rGO)–metal oxide
761 (TiO 2/Fe 3 O 4) based nanocomposites. Rsc Advances 2015, 5, (89), 73249-73260.

35
Page 37 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

762 72. Donarelli, M.; Ottaviano, L., 2D materials for gas sensing applications: A review on
763 graphene oxide, MoS2, WS2 and phosphorene. Sensors 2018, 18, (11), 3638.
764 73. Zou, C.; Hu, J.; Su, Y.; Shao, F.; Tao, Z.; Huo, T.; Zhou, Z.; Hu, N.; Yang, Z.; Kong, E.

Environmental Science: Water Research & Technology Accepted Manuscript


765 S.-W., Three-dimensional Fe3O4@ reduced graphene oxide heterojunctions for high-
766 performance room-temperature NO2 sensors. Frontiers in Materials 2019, 6, 195.
767 74. Jung, C.; Boateng, L. K.; Flora, J. R. V.; Oh, J.; Braswell, M. C.; Son, A.; Yoon, Y.,
768 Competitive adsorption of selected non-steroidal anti-inflammatory drugs on activated biochars:
769 Experimental and molecular modeling study. Chemical Engineering Journal 2015, 264, 1-9.
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

770 75. Chen, H.; Gao, B.; Li, H.; Ma, L. Q., Effects of pH and ionic strength on
771 sulfamethoxazole and ciprofloxacin transport in saturated porous media. Journal of Contaminant
772 Hydrology 2011, 126, (1), 29-36.
773 76. Chen, H.; Gao, B.; Li, H., Functionalization, pH, and ionic strength influenced sorption
774 of sulfamethoxazole on graphene. Journal of Environmental Chemical Engineering 2014, 2, (1),
775 310-315.
776 77. Prescott, L., Kinetics and metabolism of paracetamol and phenacetin. British journal of
777 clinical pharmacology 1980, 10, (S2), 291S-298S.
778 78. Sotelo, J. L.; Ovejero, G.; Rodríguez, A.; Álvarez, S.; Galán, J.; García, J., Competitive
779 adsorption studies of caffeine and diclofenac aqueous solutions by activated carbon. Chemical
780 Engineering Journal 2014, 240, 443-453.
781 79. Torrellas, S. Á.; García Lovera, R.; Escalona, N.; Sepúlveda, C.; Sotelo, J. L.; García, J.,
782 Chemical-activated carbons from peach stones for the adsorption of emerging contaminants in
783 aqueous solutions. Chemical Engineering Journal 2015, 279, 788-798.
784 80. Lee, H.-J.; Kim, G.; Kwon, Y.-K., Molecular adsorption study of nicotine and caffeine on
785 single-walled carbon nanotubes from first principles. Chemical Physics Letters 2013, 580, 57-61.
786 81. Calisto, V.; Ferreira, C. I. A.; Oliveira, J. A. B. P.; Otero, M.; Esteves, V. I., Adsorptive
787 removal of pharmaceuticals from water by commercial and waste-based carbons. Journal of
788 Environmental Management 2015, 152, 83-90.
789 82. Raza, A., Development and application of spectrophotometric methods for the
790 determination of citalopram hydrobromide in dosage forms. Chemical and pharmaceutical
791 bulletin 2006, 54, (4), 432-434.
792 83. Bebawy, L. I.; El-Kousy, N.; Suddik, J. K.; Shokry, M., Spectrophotometric
793 determination of fluoxetine and sertraline using chloranil, 2, 3 dichloro-5, 6 dicyano
794 benzoquinone and iodine. Journal of pharmaceutical and biomedical analysis 1999, 21, (1), 133-
795 142.
796 84. O'Hagan, D., Understanding organofluorine chemistry. An introduction to the C–F bond.
797 Chemical Society Reviews 2008, 37, (2), 308-319.
798 85. Trovó, A. G.; Nogueira, R. F. P.; Agüera, A.; Fernandez-Alba, A. R.; Sirtori, C.; Malato,
799 S., Degradation of sulfamethoxazole in water by solar photo-Fenton. Chemical and toxicological
800 evaluation. Water Research 2009, 43, (16), 3922-3931.
801 86. Li, W.; Nanaboina, V.; Zhou, Q.; Korshin, G. V., Effects of Fenton treatment on the
802 properties of effluent organic matter and their relationships with the degradation of
803 pharmaceuticals and personal care products. Water Research 2012, 46, (2), 403-412.
804 87. Pérez-Estrada, L. A.; Malato, S.; Gernjak, W.; Agüera, A.; Thurman, E. M.; Ferrer, I.;
805 Fernández-Alba, A. R., Photo-Fenton degradation of diclofenac: identification of main
806 intermediates and degradation pathway. Environmental science & technology 2005, 39, (21),
807 8300-8306.

36
Environmental Science: Water Research & Technology Page 38 of 40
View Article Online
DOI: 10.1039/D0EW00140F

808 88. Yu, H.; Nie, E.; Xu, J.; Yan, S.; Cooper, W. J.; Song, W., Degradation of diclofenac by
809 advanced oxidation and reduction processes: kinetic studies, degradation pathways and toxicity
810 assessments. Water research 2013, 47, (5), 1909-1918.

Environmental Science: Water Research & Technology Accepted Manuscript


811 89. Méndez-Arriaga, F.; Esplugas, S.; Giménez, J., Degradation of the emerging contaminant
812 ibuprofen in water by photo-Fenton. Water Research 2010, 44, (2), 589-595.
813 90. Gu, M.; Farooq, U.; Lu, S.; Zhang, X.; Qiu, Z.; Sui, Q., Degradation of trichloroethylene
814 in aqueous solution by rGO supported nZVI catalyst under several oxic environments. Journal of
815 Hazardous Materials 2018, 349, 35-44.
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

816 91. Rezaei, F.; Vione, D., Effect of pH on zero valent iron performance in heterogeneous
817 fenton and fenton-like processes: A review. Molecules 2018, 23, (12), 3127.
818 92. Tovar-Gómez, R.; Rivera-Ramírez, D. A.; Hernández-Montoya, V.; Bonilla-Petriciolet,
819 A.; Durán-Valle, C. J.; Montes-Morán, M. A., Synergic adsorption in the simultaneous removal
820 of acid blue 25 and heavy metals from water using a Ca(PO3)2-modified carbon. Journal of
821 Hazardous Materials 2012, 199-200, 290-300.
822 93. Leodopoulos, C.; Doulia, D.; Gimouhopoulos, K.; Triantis, T. M., Single and
823 simultaneous adsorption of methyl orange and humic acid onto bentonite. Applied Clay Science
824 2012, 70, 84-90.
825 94. Nguyen, K. M.; Nguyen, B. Q.; Nguyen, H. T.; Nguyen, H. T., Adsorption of arsenic and
826 heavy metals from solutions by unmodified iron-ore sludge. Applied Sciences 2019, 9, (4), 619.
827 95. Litter, M. I.; Slodowicz, M., An overview on heterogeneous Fenton and photoFenton
828 reactions using zerovalent iron materials. Journal of Advanced Oxidation Technologies 2017, 20,
829 (1).
830 96. Malik, P. K.; Saha, S. K., Oxidation of direct dyes with hydrogen peroxide using ferrous
831 ion as catalyst. Separation and Purification Technology 2003, 31, (3), 241-250.
832 97. Zhuang, Y.; Wang, X.; Zhang, L.; Dionysiou, D. D.; Kou, Z.; Shi, B., Double-network
833 hydrogel templated FeS/graphene with enhanced PMS activation performance: considering the
834 effect of the template and iron species. Environmental Science: Nano 2020, 7, (3), 817-828.
835 98. Zhuang, Y.; Wang, X.; Zhang, L.; Kou, Z.; Shi, B., Confinement Fenton-like degradation
836 of perfluorooctanoic acid by a three dimensional metal-free catalyst derived from waste. Applied
837 Catalysis B: Environmental 2020, 275, 119101.
838 99. Xu, L.; Wang, J., A heterogeneous Fenton-like system with nanoparticulate zero-valent
839 iron for removal of 4-chloro-3-methyl phenol. Journal of Hazardous Materials 2011, 186, (1),
840 256-264.
841 100. Mao, Q.; Zhou, Y.; Yang, Y.; Zhang, J.; Liang, L.; Wang, H.; Luo, S.; Luo, L.;
842 Jeyakumar, P.; Ok, Y. S.; Rizwan, M., Experimental and theoretical aspects of biochar-supported
843 nanoscale zero-valent iron activating H2O2 for ciprofloxacin removal from aqueous solution.
844 Journal of Hazardous Materials 2019, 380, 120848.
845 101. Keenan, C. R.; Sedlak, D. L., Factors Affecting the Yield of Oxidants from the Reaction
846 of Nanoparticulate Zero-Valent Iron and Oxygen. Environmental Science & Technology 2008,
847 42, (4), 1262-1267.
848 102. Lee, C.; Keenan, C. R.; Sedlak, D. L., Polyoxometalate-Enhanced Oxidation of Organic
849 Compounds by Nanoparticulate Zero-Valent Iron and Ferrous Ion in the Presence of Oxygen.
850 Environmental Science & Technology 2008, 42, (13), 4921-4926.
851 103. Hu, S.; Yao, H.; Wang, K.; Lu, C.; Wu, Y., Intensify removal of nitrobenzene from
852 aqueous solution using nano-zero valent iron/granular activated carbon composite as Fenton-like
853 catalyst. Water, Air, & Soil Pollution 2015, 226, (5), 155.

37
Page 39 of 40 Environmental Science: Water Research & Technology
View Article Online
DOI: 10.1039/D0EW00140F

854 104. Herney-Ramirez, J.; Vicente, M. A.; Madeira, L. M., Heterogeneous photo-Fenton
855 oxidation with pillared clay-based catalysts for wastewater treatment: A review. Applied
856 Catalysis B: Environmental 2010, 98, (1), 10-26.

Environmental Science: Water Research & Technology Accepted Manuscript


857 105. Jain, B.; Singh, A. K.; Kim, H.; Lichtfouse, E.; Sharma, V. K., Treatment of organic
858 pollutants by homogeneous and heterogeneous Fenton reaction processes. Environmental
859 Chemistry Letters 2018, 16, (3), 947-967.
860 106. Bello, M. M.; Raman, A. A. A.; Asghar, A., A review on approaches for addressing the
861 limitations of Fenton oxidation for recalcitrant wastewater treatment. Process Safety and
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

862 Environmental Protection 2019.


863 107. Pignatello, J. J.; Oliveros, E.; MacKay, A., Advanced oxidation processes for organic
864 contaminant destruction based on the Fenton reaction and related chemistry. Critical reviews in
865 environmental science and technology 2006, 36, (1), 1-84.
866 108. Munoz, M.; Conde, J.; de Pedro, Z. M.; Casas, J. A., Antibiotics abatement in synthetic
867 and real aqueous matrices by H2O2/natural magnetite. Catalysis Today 2018, 313, 142-147.
868

869

870

871

872

38
Published on 24 June 2020. Downloaded by University of Exeter on 6/25/2020 9:04:16 PM.

Table of Content
Environmental Science: Water Research & Technology
DOI: 10.1039/D0EW00140F
View Article Online
Page 40 of 40

Environmental Science: Water Research & Technology Accepted Manuscript

You might also like