Subscriber access provided by Iowa State University | Library
Sustainability Engineering and Green Chemistry
Degradation of ofloxacin by perylene diimide supramolecular nanofiber sunlight-driven photocatalysis Ping Chen, Lee Blaney, Giovanni Cagnetta, Jun Huang, Bin Wang, Yujue Wang, Shubo Deng, and Gang Yu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b05827 • Publication Date (Web): 03 Jan 2019 Downloaded from http://pubs.acs.org on January 5, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 28
Environmental Science & Technology
1
Degradation of ofloxacin by perylene diimide supramolecular nanofiber sunlight-driven
2
photocatalysis
3 4
Ping Chena#, Lee Blaneya,b#, Giovanni Cagnettaa, Jun Huanga, Bin Wanga,
5
Yujue Wanga, Shubo Denga, Gang Yua*
6 7
a School
8
Beijing Key Laboratory for Emerging Organic Contaminants Control, Tsinghua University, Beijing 100084, China
9
b University
of Environment, State Key Joint Laboratory of Environmental Simulation and Pollution Control (SKLESPC),
of Maryland Baltimore County, Department of Chemical, Biochemical, and Environmental Engineering, 1000
10
Hilltop Circle, Engineering 314, Baltimore, Maryland 21250, United States
11
# The first two authors contributed equally to this work.
12 13
* Corresponding Author:
14
Gang YU (Ph.D.)
15
School of Environment, Tsinghua University, Beijing, 100084, China
16
Tel: +86-10-62787137; Fax: +86-10-62785687; E-mail:
[email protected] 17
ACS Paragon Plus Environment
Environmental Science & Technology
18
Page 2 of 28
Abstract
19 20
This study describes a promising sunlight-driven photocatalyst for the treatment of ofloxacin and other fluoroquinolone
21
antibiotics in water and wastewater. Perylene diimide (PDI) supramolecular nanofibers, which absorb a broad spectrum of
22
sunlight, were prepared via a facile acidification polymerization protocol. Under natural sunlight, the PDI photocatalysts
23
achieved rapid treatment of fluoroquinolone antibiotics, including ciprofloxacin, enrofloxacin, norfloxacin, and ofloxacin.
24
The fastest degradation was observed for ofloxacin, which had a half-life of 2.08 min for the investigated conditions.
25
Various light sources emitting in the UV-visible spectrum were tested, and blue light was found to exhibit the fastest
26
ofloxacin transformation kinetics due to the strong absorption by the PDI catalyst. Reactive species, namely h+, 1O2, and
27
O2•-, comprised the primary photocatalytic mechanisms for ofloxacin degradation. Frontier electron density calculations
28
and mass spectrometry were used to verify the major degradation pathways of ofloxacin by the PDI-sunlight
29
photocatalytic system and identify the transformation products of ofloxacin, respectively. Degradation mainly occurred
30
through demethylation at the piperazine ring, ketone formation at the morpholine moiety, and aldehyde reaction at the
31
piperazinyl group. An overall mechanism was proposed for ofloxacin degradation in the PDI-sunlight photocatalytic
32
system, and the effects of water quality constituents were examined to determine performance in real water/wastewater
33
systems. Ultimately, the aggregate results from this study highlight the suitability of the PDI-sunlight photocatalytic
34
system to treat antibiotics in real water and wastewater systems.
35 36
TOC GRAPHIC
37
ACS Paragon Plus Environment
Page 3 of 28
38
Environmental Science & Technology
INTRODUCTION
39 40
Antibiotics are common organic micropollutants that are routinely detected in surface water, wastewater, and drinking
41
water.1,2 Over-prescription and -consumption of antibiotics, along with their incomplete removal in wastewater treatment
42
plants, have increased concerns associated with antibiotic presence in water resources. Ofloxacin is a fluoroquinolone
43
antibiotic that has been frequently detected in surface water at concentrations as high as 31.7 μg L-1. 3,4 Despite the
44
relatively low concentration, ofloxacin occurrence can adversely affect ecological systems by inhibiting the growth of
45
microorganisms, inducing the selection of antibiotic resistant bacteria, and leading to other ecotoxicological effects.5,6 For
46
these reasons, novel treatment processes are required for the selective and efficient removal of ofloxacin and other
47
fluoroquinolone antibiotics from diverse water and wastewater matrices.
48 49
Conventional activated sludge systems and advanced oxidation process are commonly used for wastewater treatment.7
50
However, conventional activated sludge generally demonstrates low removal for organic micropollutants, such as
51
ofloxacin.8 The treatment of ofloxacin and/or other fluoroquinolone antibiotics has also been studied in various systems,
52
such as electrochemical oxidation, ultraviolet (UV)-coupled systems (e.g., UV-H2O2), acoustic chemical oxidation, ozone-
53
coupled systems (e.g., O3-H2O2), Fenton oxidation, persulfate oxidation, laccase-mediated oxidation, and photocatalytic
54
oxidation.9-13 Of these, environmentally friendly photocatalytic materials can utilize solar energy without the need for
55
exogenous chemical reagents or electrical energy.14-18 These advantages make novel sunlight-photocatalysis systems a
56
desirable technology for treatment of micropollutants.
57 58
Traditional semiconductor photocatalysts are characterized by high cost, poor structural adjustability, low activity under
59
solar irradiation, and toxicity concerns from leachable metal content. For these reasons, new photocatalytic materials need
60
to be explored. Perylene diimide (PDI) and its derivatives are considered to be the best n-type organic semiconductors.19
61
Recently, PDI has been used in solar cells,20 fluorescent materials,21,22 single-molecule sensors,23 and organic transistors24
62
due to its unique photoelectrical properties. Importantly, PDI can be formed into supramolecular nanofibers through
63
acidification polymerization,25 and the resultant nanofibers have strong photothermal stability and notable photo-oxidation
64
and -reduction characteristics. Ongoing research has focused on the synthesis, characterization, and application of PDI-
65
based materials.26 Since 2014, traditional semiconductor photocatalysts have been modified with PDI to split water via ACS Paragon Plus Environment
Environmental Science & Technology
Page 4 of 28
66
composite catalysts, such as PDI/Ni,27 PDI/Ru,28 PDI/TiO2,29 and PDI/C3N4.30 The results of these studies have shown that
67
PDI incorporation improved the activity of traditional semiconductor materials under visible light. However, these
68
composite catalysts were still primarily comprised of traditional semiconductor materials but with a small amount of PDI
69
nanofibers to facilitate the migration of charge carriers.
70 71
In recent years, Zhu and coworkers improved the preparation and synthesis protocols for novel PDI supramolecular
72
nanofibers that catalyze phenol degradation with visible light excitation.31-33 Under identical experimental conditions, the
73
phenol degradation rate with PDI supramolecular nanofibers was 20 and 13 times that of non-supramolecular PDI and g-
74
C3N4, respectively. The PDI photocatalyst possessed unique one-dimensional π conjugation and internal electric field
75
effects, which enabled electrons to migrate to the nanofiber surface. At the surface, electrons reacted with dissolved
76
oxygen to form reactive oxygen species that resulted in strong oxidation activity.34 Because many organic micropollutants
77
contain phenol moieties or other functional groups that interact with reactive oxygen species, the application of PDI
78
photocatalysis to these emerging contaminants merits further study. Given the high energetic costs of UV light in UV-
79
based advanced oxidation processes,35 sunlight-activated photocatalysts are needed to ensure sustainable treatment. No
80
previous efforts have explored this facet of PDI-based catalysts.
81 82
To date, the synthesis, application, and understanding of PDI supramolecular nanomaterials involve the following
83
shortcomings: (i) the metal modification poses a potential threat to aquatic ecosystems; (ii) applications have been limited
84
to production of hydrogen and degradation of simple chemicals (e.g., phenol); (iii) natural sunlight has not been employed
85
for photocatalytic treatment of contaminants; and, (iv) mechanistic insight into the photocatalytic reactions has not been
86
studied. These shortcomings represent knowledge gaps in (1) production of organic PDI nanofibers without transition
87
metals, (2) application of PDI supramolecular nanomaterials to treat organic micropollutants in water and wastewater, (3)
88
utilization of solar irradiation, and (4) identification of reaction mechanisms, pathways, and transformation products. For
89
structurally-complex environmental contaminants, such as ofloxacin, the transformation products generated by PDI
90
photocatalytic reactions may retain toxicological concerns and, therefore, need to be identified. We note that no previous
91
reports have explored the potential for PDI-based materials to transform micropollutants.
92 93
In this study, PDI supramolecular, sunlight-activated photocatalysts were prepared by acidification polymerization, and ACS Paragon Plus Environment
Page 5 of 28
Environmental Science & Technology
94
the catalytic mechanism of these novel materials was investigated for degradation of the ofloxacin micropollutant. The
95
research content includes the preparation and characterization of PDI, the catalyst activity for different light sources, the
96
adsorption and degradation of ofloxacin on the surface of the catalyst, the generation of reactive species, and the impacts
97
of water quality parameters (e.g., pH, cations, anions, and dissolved organic matter (DOM)) in the PDI-sunlight
98
photocatalytic system. The transformation products of ofloxacin were determined and used to inform the photocatalytic
99
mechanism and ofloxacin degradation pathway. Ultimately, these results highlight the potential application of novel,
100
sustainable, sunlight-driven photocatalysts to address growing concerns with micropollutants in water and wastewater.
101 102
MATERIALS AND METHODS
103 104
Preparation and analysis of PDI supramolecular photocatalyst nanofibers. A brief overview of the reagents and ultrapure
105
water is provided in Text S1 of the Supporting Information (SI). Preparation of the PDI supramolecular photocatalyst
106
generally followed the protocol of Zhu et al.,31 although key improvements and modifications were implemented here. Briefly,
107
6 mM 3,4,9,10-perylenetetracarboxylic diimide, 40 mM β-alanine, and 25 g imidazole were accurately weighed and heated at
108
100 °C for 4 h under argon. After cooling to room temperature, the mixture was dispersed in 180 mL of ethanol and 400 mL of
109
2 M hydrochloric acid (HCl) and stirred for 12 h. The resulting alanine-modified material was filtered through a 0.45-μm
110
membrane (Beijing Ruifeng Tongchuang Analysis Instrument Co. Ltd.; Beijing, China). After washing with ultrapure water and
111
drying at 60 °C, the solid substance was dissolved in 250 mL solution containing 0.6% (v/v) triethylamine. Then, 35 mL of 4 M
112
HCl was added to spontaneously form nanofibers. After two repeated cycles of washing with ultrapure water and drying at
113
60 °C, the PDI supramolecular nanofibers were produced.
114 115
The morphology of PDI supramolecular nanofibers was investigated by transmission electron microscopy (TEM) on a
116
JEM-2100F instrument at an acceleration voltage of 200 kV (Japan Electron Optics Laboratory Company; Japan). X-ray
117
diffraction (XRD) patterns were obtained with an Ultima III diffractometer (Rigaku, Japan). The d-spacing from XRD
118
patterns was calculated using Eq. 1.
119 120
2𝑑sin 𝜃 = 𝑛𝜆
121
(Eq. 1)
122 ACS Paragon Plus Environment
Environmental Science & Technology
Page 6 of 28
123
In Eq. 1, 𝜃 is the diffraction angle (recorded from the XRD pattern), 𝜆 is the x-ray wavelength, and 𝑛 was the order of
124
diffraction (first). Other structural, morphological, and absorbance characterization methods are described in Text S2 of
125
the SI.
126 127
Photocatalysis experiments. Experimental solutions with 200 mg L-1 of PDI and 8 mg L-1 of ofloxacin were generated in
128
200 mL beakers and ultrasonically dispersed for 30 min to ensure that the PDI catalyst and ofloxacin contaminant
129
achieved adsorption equilibrium.18 The temperature was 28 ± 3 °C, and a magnetic stirrer was used to ensure well-mixed
130
conditions. Photocatalysis experiments were carried out under actual sunlight. Tests were conducted at 10:00 a.m. to
131
14:00 p.m. on sunny days in May through September in 2017 and 2018 in Beijing, China (N 40.00°, E 116.33°). Light
132
intensity was measured using the CEL-NF2000 light power meter from Beijing Zhongjiao Jinyuan Technology Co.
133
(China).
134 135
To examine the photocatalyst activity under different light sources, six 9 W light-emitting diodes (LEDs; Epileds
136
Technologies, Taiwan) were used to generate UV light (390-400 nm; 3.4 mW cm-2), blue light (455-460 nm; 5.8 mW
137
cm-2), green light (515-530 nm; 5.8 mW cm-2), yellow light (590-595 nm; 3.7 mW cm-2), red light (655-660 nm; 13.0 mW
138
cm-2), and white light (400-840 nm; 9.2 mW cm-2). The LEDs were placed 10 cm away from the experimental solutions.
139 140
Reactive oxygen species were investigated by electron spin resonance (ESR) spectroscopy (JES-FA200, Bruker;
141
Germany). Apparent, time-based ofloxacin rate constants were measured in the presence and absence of quenching agents
142
to inform reaction mechanisms. Detailed information about reactive species determination is provided in Text S3 of the
143
SI. These data were used to calculate the contribution of reactive species (e.g., •OH, 1O2, O2•-, h+) to the photocatalytic
144
degradation of ofloxacin using Eqs. 2-5.
145 146
R•OH =
k•OH 𝑘app
≈
(𝑘app ― 𝑘IPA) 𝑘app
147
(Eq. 2)
148 149
R O2 = 1
k1O2 𝑘app
≈
(𝑘IPA ― 𝑘NaN3) 𝑘app
ACS Paragon Plus Environment
Page 7 of 28
Environmental Science & Technology
150
(Eq. 3)
151 152
kO•2 -
RO•2 - =
𝑘app
≈
(𝑘app ― 𝑘TEMPL) 𝑘app
153
(Eq. 4)
154 155
Rh + =
kh + 𝑘app
≈
(𝑘app ― 𝑘Na2C2O2) 𝑘app
156
(Eq. 5)
157 158
In Eqs. 2-5, R is the fractional contribution of a particular reactive species (e.g., •OH, 1O2, O2-•, h+) to the apparent rate
159
constant (kapp) and ki is the rate constant for the degradation of ofloxacin by a certain reactive species (indicated by the
160
•OH, 1O
161
sodium azide (NaN3), 4-hydroxy-2,2,6,6-tetramethylpiperidinyloxy (TEMPL), and sodium oxalate (Na2C2O4) subscripts).
2,
O2-•, and h+ subscripts) or in the presence of a quenching agent (indicated by the isopropyl alcohol (IPA),
162 163
To investigate pH effects on ofloxacin degradation, the fraction of each protonated/deprotonated species in solution (i.e.,
164
α0, α1, and α2) was calculated using Eqs. 6-9.36
165 166
∑ α [OFL]
[OFL]total = [OFL + ] + [OFL +/ - ] + [OFL - ] =
i
total
i = 0,1,2
167
(Eq. 6)
168 169
(
α0 = 1 +
𝐾a1 [H + ]
+
)
𝐾a1𝐾a2 [H + ]
2
―1
170
(Eq. 7)
171
172
(
𝐾a2 [H + ] +1+ α1 = 𝐾a1 [H + ]
)
―1
173
(Eq. 8) ACS Paragon Plus Environment
Environmental Science & Technology
Page 8 of 28
174
(
2
)
[H + ] [H + ] + +1 α2 = 𝐾a1𝐾a2 𝐾a2
175
―1
176
(Eq. 9)
177 178
In Eq. 6, [OFL] is the molar ofloxacin concentration, α is the ionization factor, the “+”, “+/-”, and “-” superscripts indicate
179
the ofloxacin cation, zwitterion, and anion, respectively, the i subscript represents the ofloxacin species from most (0) to
180
least (2) protonated, and the “total” subscript indicates the total concentration. In Eqs. 7-9, 𝐾a1 and 𝐾a2 are the acid
181
dissociation constants for ofloxacin and [H + ] is the molar hydronium concentration.
182 183
When studying the effects of environmental factors, SO42- (Na2SO4, 10 mg L-1), NO3- (NaNO3, 10 mg L-1), Cu2+ (CuSO4, 5
184
mg L-1), DOM (fulvic acid, 100 mg C L-1), Cl- (NaCl, 10 mg L-1), or HCO3- (NaHCO3, 10 mg L-1) were added to the
185
experimental solution containing 200 mg L-1 of PDI and 8 mg L-1 of ofloxacin. To identify the effects of real water
186
matrices on ofloxacin degradation by the PDI-sunlight photocatalytic system, experimental solutions were generated by
187
adding ofloxacin powder to 0.45-μm membrane-filtered tap water (Beijing, China), river water (Pearl River water;
188
Guangzhou, China), and wastewater effluent (Beijing, China). For all conditions, the solution pH (3.0-8.0) and the optical
189
power density of sunlight (45-55 mW cm-2) were measured during experimentation. Samples were collected at pre-
190
determined times. After the catalyst was filtered out, the ofloxacin concentration was determined by high performance
191
liquid chromatography (HPLC; LC-20AT, Shimadzu, Japan).
192 193
Ofloxacin analysis and identification of transformation products. Ofloxacin concentrations were measured by HPLC
194
with photodiode array (PDA) detection (Shimadzu UV model with SPD-M20A) and the transformation products were
195
analyzed by ultraperformance liquid chromatography with quadrupole time-of-flight mass spectrometry (UPLC-Q-TOF-
196
MS; Agilent Technologies 1290UHPLC/6540QTOF-MS; United States). Detailed information on these techniques is
197
provided in Text S4 and Table S1 of the SI.
198 199
Computational methodology to predict attack sites. The conduction band (CB) value was calculated using Eq. 10.
200 ACS Paragon Plus Environment
Page 9 of 28
201
Environmental Science & Technology
CB = VB ― [GAP]
202
(Eq. 10)
203 204
In Eq. 10, VB is the valence band value and [GAP] is the band gap width of PDI, which was calculated from the UV-
205
visible/diffuse reflectance spectroscopy (UV-vis/DRS) spectrum.
206 207
The B3LYP density functional theory model and the 6-311+g (d, p) pople basis set were used to process the ofloxacin
208
molecular orbital calculations in Gaussian 09. In particular, these tools were used to calculate the frontier electron
209
densities (FEDs) of the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital
210
(LUMO), along with the point charges. These data were subsequently used to predict the site of reactive species attack on
211
ofloxacin.
212 213 214
RESULTS AND DISCUSSION
215 216
Structure, morphology, and absorbance spectrum of the PDI supramolecular nanofibers.
217
Following the protocols described above, the PDI supramolecular nanofibers were self-assembled with carboxy-
218
substituted PDI molecules through H-type π-π stacking and hydrogen bonding; additional information is available in
219
Wang et al.31 The incorporation of β-alanine improved the growth of the supramolecular system by providing short
220
carboxylic acid side chains that facilitated columnar stacking of perylene groups and increased solubility.31 The phase
221
structures of the synthesized PDI supramolecular nanofibers were elucidated by XRD analysis (Fig. 1a). The PDI material
222
had a high degree of crystallinity and, according to Eq. 1, the typical d-spacing of π-π stacking was 3.44 Å for the labeled,
223
Pa peak. The small d-spacing of π-π stacking in the PDI nanofibers is expected to benefit the migration of charge
224
carriers.29 Fig. 1b–d depict the morphologies of the prepared PDI materials by TEM and atomic force microscopy (AFM).
225
The TEM images revealed fairly uniform nanofibers with ~20-50 nm diameters and ~100-400 nm lengths. The height
226
profile was determined by AFM and confirmed the ~20 nm diameter of individual nanofibers. Following acidification
227
polymerization, the PDI materials exhibited a small surface area (Fig. S1 in the SI), suggesting a low adsorption capacity
228
that may negatively affect photocatalytic mechanisms. ACS Paragon Plus Environment
Environmental Science & Technology
(a)
(b)
Intensity (a.u.)
PDI 6.2
14 9.9
5
10
229
15
Page 10 of 28
Pa, 25.9 20 22.9
20
2 Theta (degree)
25
30
(d)
(c)
~40 ~40 nm nm ~20 nm
~20 nm
230 231
Figure 1. (a) XRD pattern of the prepared PDI photocatalyst; (b) and (c) TEM images of PDI supramolecular nanofibers; and,
232
(d) AFM analysis of an individual PDI nanofiber.
233 234
The functional groups of the prepared PDI supramolecular nanofibers were identified through Fourier-transform infrared
235
(FTIR) spectroscopy (Fig. S2a in the SI). The stretching vibration band at ~3450 cm-1 was assigned to C-OH in carboxylic
236
acid moieties; accordingly, the absorption band at 1700 cm-1 was attributed to the C=O of the same group. Another C=O
237
group at 1650 cm-1 was assigned to keto functionalities in the PDI molecular structure.37 The stretching vibration band at
238
1440-1600 cm-1 was attributed to the C=C groups from the aromatic carbon skeleton. The stretching vibration peaks of the
239
C-H (1350 cm-1), CH2 (2850-3090 cm-1), and σC-H (615-1015 cm-1) correspond to in-plane bending vibrations from the
240
aromatic rings, the aliphatic group, and the out-of-plane bending vibrations in the aromatic rings, respectively. The
241
absorption peaks of -N-C=O at ~1400 cm-1 and -N-CH2 at ~1030-1240 cm-1 indicated the presence of amide groups in the ACS Paragon Plus Environment
Page 11 of 28
Environmental Science & Technology
242
PDI nanofibers and chemical reactions between β-alanine and PDI, respectively.38 Overall, the stretching vibration peaks
243
demonstrated the existence of carboxyl, -N-CH2, and -CH2 groups in the PDI supramolecular nanofibers. Among these
244
groups, the carboxyl and -N-CH2 functionalities may aid in sorption of ofloxacin cations/zwitterions and
245
zwitterions/anions, respectively. The innovative acidification polymerization protocol contributed to these newly
246
discovered groups. X-ray photoelectron spectroscopy (XPS) analysis further confirmed that multiple functional groups
247
were present on the surface of the PDI supramolecular nanofibers (Fig. S2b-e in the SI). These results, along with the
248
FTIR spectroscopy analysis, confirmed that β-alanine was successfully introduced into the PDI supramolecular
249
nanofibers.
250 251
Fig. S2f in the SI displays the UV-vis/DRS spectrum of the PDI nanofibers. The PDI photocatalyst presented a wide
252
absorption edge at ~730 nm, suggesting that the supramolecular nanofibers are capable of being activated by visible light
253
(i.e., 390-760 nm). Compared to traditional photocatalysts, such as TiO2 (~320 nm), BiPO4 (~300 nm), and g-C3N4 (~450
254
nm), the absorption spectrum of the PDI photocatalyst demonstrated an obvious red shift and, therefore, enhancement of
255
visible-light absorption. The band gap energy of the PDI material was calculated to be 1.73 eV (see inset of Fig. S2f in the
256
SI). Overall, the structure, morphology, and absorption spectrum of the PDI supramolecular nanofibers demonstrate
257
successful preparation via the modified acidification polymerization protocol and highlight the potential application of
258
these novel materials under sunlight irradiation.
259 260
Photocatalytic degradation kinetics and mechanisms.
261 262
Degradation of fluoroquinolones by the PDI-sunlight photocatalytic system. The activity of the sunlight-activated
263
PDI nanofibers was evaluated for the degradation of four fluoroquinolone antibiotics. Fig. 2a shows the observed time-
264
based rate constant for ciprofloxacin, enrofloxacin, norfloxacin, and ofloxacin degradation by the PDI-sunlight
265
photocatalysis system. The rate constants for ciprofloxacin, enrofloxacin, norfloxacin, and ofloxacin were (9.4 ± 0.8)×10-
266
2,
267
fluoroquinolones were rapidly transformed by the PDI nanofiber photocatalyst. The fastest degradation was obtained for
268
ofloxacin, and this finding was attributed to the chemical structure as discussed below in the degradation pathways
269
section. As ofloxacin showed the fastest degradation, further testing was completed with this fluoroquinolone antibiotic to
(12.6 ± 0.6)×10-2, (7.0 ± 0.7)×10-2, and (33.4 ± 2.0)×10-2 min-1, respectively. These results indicated that all four
ACS Paragon Plus Environment
Environmental Science & Technology
Page 12 of 28
270
understand the reaction kinetics, mechanisms, and products; however, we note that similar mechanisms are expected for
271
the other compounds in this structurally-similar class of antibiotics.
272
(a)
H N
OFL
(b)
O N
N
F O
O
NOR
C/C0 for OFL
H2N N
N
OH
F O
O
H N
ENR
N
N OH
F O
O
H2N N
CIP
Dark
O
0.2
0.3 -1
kobs (min )
0.4
0.5
Sunlight
0.4
PDI-adsorption Photolysis PDI-photocatalysis
0.0
-30 -1
0
1
2
Absorbance (a.u.)
0.25
4
Time (min)
5
6
7
0.00
(c) 0.35 0.30
3
0.05 Blue light
UV light
Green light
Yellow light
Red light
0.10 0.15
kobs (min-1)
273
0.6
OH
F
0.1
0.8
0.2
N
O
0.0
1.0
OH
0.20
0.20
0.25
0.15
0.30 0.10
0.35
0.05
0.40
0.00
0.45
350
400
450
500
550
Wavelength (nm)
600
650
274 275
Figure 2. (a) Rate constants for photocatalytic degradation of four fluoroquinolones (e.g., OFL, ofloxacin; NOR, norfloxacin;
276
ENR, enrofloxacin; CIP, ciprofloxacin) by sunlight-activated PDI; (b) Degradation of ofloxacin by photolysis, PDI adsorption,
277
and PDI photocatalysis under sunlight irradiation; and, (c) The UV-vis absorption spectra of PDI supramolecular nanofibers
278
overlaid with the rate constants for ofloxacin degradation by the PDI photocatalyst. In all cases, the initial concentration of
279
fluoroquinolone antibiotic was 8 mg L-1, the concentration of PDI nanofibers was 200 mg L-1, the solution pH was 5.6, and error
280
bars indicate one standard deviation.
281 282
The contributions of hydrolysis, direct photolysis, adsorption, and photocatalysis to the apparent degradation of ofloxacin ACS Paragon Plus Environment
Page 13 of 28
Environmental Science & Technology
283
are presented in Fig. 2b. No change in ofloxacin concentration was recorded in the absence of PDI photocatalysts and
284
light. A 6% change in the ofloxacin concentration was observed over the 7-min irradiation period in the absence of PDI
285
nanofibers, suggesting that photolysis was not a major degradation mechanism for the experimental conditions. The
286
adsorption and photocatalytic processes were carried out following the methods described in the photocatalysis
287
experiments section. After 30 min adsorption under dark conditions, the 200 mg L-1 of PDI nanofibers adsorbed
288
approximately 15% of the ofloxacin (C0 = 8 mg L-1), indicating an adsorption capacity of 40 mg g-1 for the experimental
289
conditions. With sunlight irradiation, the PDI supramolecular nanofibers enhanced the ofloxacin degradation to ~91%,
290
verifying the ability of the PDI material to rapidly catalyze ofloxacin degradation using solar energy. Importantly, the
291
observed rate constants for ofloxacin degradation by PDI are one to three orders of magnitude higher than those observed
292
in other photocatalytic systems,39-44 suggesting that this material may have unique advantages over previously reported
293
catalysts for micropollutant treatment.
294 295
To further examine the photocatalytic activity of the PDI supramolecular nanofibers, different light sources were
296
employed. The experimental details were included in the photocatalysis experiments section, and the results are displayed
297
in Fig. 2c and Table S2 of the SI. The UV-visible absorption spectra for the PDI materials demonstrated broad absorbance
298
of light at 350-650 nm, with the highest absorption at 470 nm. These results indicate that the PDI catalyst absorbs not only
299
UV-A light, but also visible light. Given the absorption maximum at ~415-550 nm, the PDI catalyst was expected to
300
demonstrate optimal activity under blue light. This hypothesis was confirmed. In fact, all light sources showed effective
301
ofloxacin degradation. The time-based rate constants for ofloxacin degradation in the PDI-UV light, -blue light, -green
302
light, -yellow light, -red light, -white light, and -sunlight photocatalytic systems were 0.223±0.014, 0.403±0.029,
303
0.271±0.027, 0.174±0.007, 0.150±0.012, 0.295±0.009, and 0.334±0.020 min-1, respectively. From these results, the PDI
304
nanofibers exhibited the highest photocatalytic activity for blue light, followed by sunlight and, then, white light. The
305
photocatalysis results were in general agreement with the UV-visible absorption spectra, highlighting the benefits of the
306
broad absorbance spectrum of the PDI nanofibers prepared through the acidification polymerization protocol. Importantly,
307
the PDI-sunlight photocatalytic system yielded a degradation half-life of 2.08 min for ofloxacin. The stability of the PDI
308
supramolecular nanofibers was tested to ensure that this performance was reproducible over sequential treatment cycles.
309
As shown in Fig. S3 of the SI, four sequential cycles showed similar ofloxacin degradation without any apparent change
310
in performance. In combination, these results highlight the potential advantages for implementation of the PDI-sunlight ACS Paragon Plus Environment
Environmental Science & Technology
311
Page 14 of 28
photocatalytic system in real systems.
312 313
Role of reactive species in ofloxacin degradation. Reactive species generated from the PDI-sunlight photocatalytic system
314
were probed using the ESR spin-trapping technique. As shown in Fig. S4a-c of the SI, no ESR signals were recorded for
315
the dark condition, while three ESR signals were observed under sunlight irradiation. The intensity ratios for two of these
316
signals were 1:2:2:1 (Fig. S4a in the SI) and 1:1:1:1 (Fig. S4b in the SI), indicating the presence of the dimethyl pyridine
317
N-oxide (DMPO)-•OH and DMPO-O2•- adducts,45-46 respectively. These findings confirm the role of •OH and O2•- in the
318
photocatalytic system. In addition to the DMPO-•OH and DMPO-O2•- adducts, another signal (1:1:1 intensity ratio) was
319
detected with 4-oxo-2,2,6,6-tetramethylpiperidine (TMPO) spin-trapping ESR and attributed to the TMPO-1O2 adduct (Fig.
320
S4c in the SI).47,48 The signal intensity of the three adducts increased with irradiation time, suggesting their continuous
321
production by the PDI supramolecular nanofiber photocatalyst during irradiation.
322 323
Chemical quenchers were applied to the experimental solutions to further elucidate the contribution of specific reactive
324
species to ofloxacin degradation. The quenching agents for h+, •OH, •OH and 1O2, and O2•- were Na2C2O4, IPA, NaN3, and
325
TEMPL, respectively. Fig. S5a of the SI shows that the various quenching agents had different effects on ofloxacin
326
degradation by the sunlight-activated PDI nanofibers. In particular, the extents of ofloxacin degradation after 7 min of
327
irradiation for the blank control and solutions containing IPA (100 mM), NaN3 (10 mM), TEMPL (5 mM), and Na2C2O4
328
(100 mM) were approximately 91%, 88%, 78%, 69%, and 53%, respectively; furthermore, the observed rate constants for
329
ofloxacin degradation under the same conditions were 0.334±0.195, 0.302±0.013, 0.216±0.013, 0.171±0.010, and
330
0.120±0.011 min-1, respectively. These results indicate that h+, O2•-, and 1O2 played important roles in ofloxacin
331
degradation by the PDI-sunlight photocatalytic system. Using Eqs. 2-5, the relative contributions of •OH, 1O2, O2•-, and h+
332
to the overall ofloxacin degradation were calculated to be 9.5 ± 4.0 %, 16.4 ± 3.5 %, 48.8 ± 2.9 %, and 64.1 ± 3.4 %,
333
respectively (Fig. S5b in the SI). Note that the contributions sum to more than 100% due to the convoluted radical
334
chemistry involved in the photocatalytic mechanism.39,49,50 Nevertheless, h+ production by the PDI-sunlight photocatalytic
335
system was found to be the major mechanism of ofloxacin degradation. The detailed mechanism for ofloxacin degradation
336
is discussed further in the photocatalytic mechanism section. Table S3 of the SI reports the contributions of •OH, 1O2, O2•-,
337
and h+ to the degradation of ciprofloxacin, norfloxacin, and enrofloxacin. The results followed the same general trend as
338
observed for ofloxacin, although the contribution of O2•- was slightly lower than that of 1O2 for enrofloxacin. For all ACS Paragon Plus Environment
Page 15 of 28
339
Environmental Science & Technology
fluoroquinolones, h+ exerted the largest impact on the overall degradation kinetics.
340 341
Degradation pathways of ofloxacin. For the 7-min irradiation experiments, the overall ofloxacin degradation was ~91%
342
but the change in total organic carbon (TOC) was just 20%, indicating a relatively low degree of mineralization for this
343
level of treatment (Fig. S6). Instead, ofloxacin degradation in the PDI-sunlight photocatalytic system produced a variety of
344
transformation products that retained structural characteristics of the ofloxacin molecule. The analytical response of these
345
products tended to, first, increase with irradiation time and, then, decrease upon further irradiation. To elucidate the
346
photocatalytic mechanisms associated with ofloxacin degradation, transformation products were identified by mixing
347
samples collected at 3, 7, and 10 min of irradiation for UPLC-Q-TOF MS analysis, which confirmed 13 major
348
transformation products. The total ion current diagram and MS/MS results are shown in Fig. S7, Table S4, and Fig. S8 of
349
the SI. Importantly, the majority of these products (i.e., P2, P4, P6, P7, P8, P9, P10, P11, P12, and P13) retain the
350
fluoroquinolone pharmacophore (i.e., the basic chemical structure needed to convey antibiotic properties) and, therefore,
351
the ability to inhibit bacterial growth when present at high enough concentrations.9 Further efforts to identify the relative
352
potency of individual transformation products will help to elucidate the extent of treatment needed to optimize operating
353
conditions and ensure removal of antimicrobial activity.
354 355
To examine the mechanisms that produced the 13 identified transformation products during photocatalysis of ofloxacin,
356
the FED method was used to predict the sites of reactive species attack. The results of this analysis are summarized in
357
Table S5 of the SI. According to frontline orbital theory: a high FED2HOMO+FED2LUMO value indicates that an atom is more
358
likely to be attacked by •OH (electrophilic reaction); a high, positive point charge indicates the likelihood of O2•- attack
359
(nucleophilic reaction) and H-abstraction; and, a high 2FED2HOMO value suggests that an atom is likely to be attacked by
360
h+.50,51 The C10 and N12 positions of the ofloxacin molecule displayed high FED2HOMO+FED2LUMO values, demonstrating
361
the potential for •OH addition in the quinolone and N-methylpiperazine rings. The C3, C7, C10, C14, C23, and C29
362
positions exhibited high, positive point charges, and so these positions are expected to undergo preferential attack by O2•-;
363
furthermore, the C10, C14, and C29 atoms may undergo H-abstraction reactions. High 2FED2HOMO values were calculated
364
for the C2, N12, and N19 positions, suggesting that these sites may be directly oxidized by h+. These analyses were used
365
with the identified transformation products to propose a tentative ofloxacin photocatalytic pathway for the PDI-sunlight
366
system (see Fig. 3). ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 28
367 HN
O N
Attack N19 hVB+
Pathway I
-CH2
Attack N12
OFL O
O
O
O
-H2
O N
N
+
O
OH
F
O
P10 O
N O
OH
F
O
P2
O
NH
O
P1
O
PDIs
N
O
H
N
Attack C29
F
N N
O
H
1
O2
Attack C14 (-2H)
O N
O O
P8
O
O
-CO
H2O
OH P11 O
OH
O
+ CO2
N
F
N
F
O P3
O HN
O
O OH
P13 O N
N
N
F
O
Pathway III
OH
O N
+O OH
O
O N
Sunlight
Pathway II
N H
N H
O
O N O
368 369
OH O
P4 HN
F
H2N
N
F
OH P9
OH
O
N
H2N
N
F
O
O HN
-CH2O
O
O N
OH P7
HN
O2·-
F
-C5H9N N
N
-CO
N
O N
HN
OH
F P6
O
O
OH
F P12 O
O
Figure 3. Proposed degradation pathway for ofloxacin transformation in the PDI-sunlight photocatalytic system.
370 371
The MS analysis demonstrated that ofloxacin and a majority of the transformation products (e.g., P2, P4, P6, P7 P8, P9,
372
P10, P11, and P12) contained an MS/MS fragment with m/z 261, suggesting that these compounds retained the quinolone
373
moiety. Further analysis of the degradation products indicated that the N-methylpiperazinyl functionality was the main
374
site of attack for reactive species. In particular, three main pathways were determined (see Fig. 3). Pathway I was the
375
demethylation reaction, which has been widely reported for photolysis of fluoroquinolones.9,52,53 The photogenerated h+
376
attacked N19, releasing the methyl group and producing P7. Then, P7 underwent dehydrogenation to form P2.54 The
377
generation of P4 and P10 was proposed to begin with the direct oxidation of P7, as an electron donor, by h+, a reaction
378
that produces the demethylated cation radical [OFL-CH3]•+.55,56 For heterogeneous photocatalysts, the presence of
379
dissolved oxygen influences the degradation mechanisms and products.57 For instance, dissolved oxygen can react with
380
conduction band e- to produce nucleophilic O2•-. From the FED analysis in Table S5 of the SI, O2•- was predicted to attack
381
the C3, C7, C10, C14, C23, and C29 positions. However, the retained m/z 261 fragment suggests that O2•- mainly attacked
382
the C14 position on the N-methylpiperazinyl group. In addition, O2•- may preferentially interact with [OFL-CH3]•+,
383
forming a zwitterion or a diradical. The detailed mechanism for O2•- reaction with the piperazine ring is shown in Eq. 11.51
384
However, the species produced in the third reaction step was unstable in water and decomposed into acetaldehyde. Similar ACS Paragon Plus Environment
Page 17 of 28
Environmental Science & Technology
385
ring cleavage products were observed in previous reports.58,59 Under acidic conditions, the acetaldehyde products undergo
386
H+ addition and hydrolysis to generate the P4 and P10 isomers. In a separate reaction, h+ attacked the N12 position on
387
ofloxacin to form P9 by effectively releasing the N-methylpiperazine.
388 O2·-
HN
HN
HN
Open ring
N
N
N
O
389
O
O
H+
HN N
O
O
H2N
Hydrolysis
N
O
HN N
H2O
390
HN
O
P10 O
O
O
H2N
P4
(Eq. 11)
391 392
Pathway II was the ketone formation reaction. Since electron-withdrawing groups can weaken the stability of the
393
transition state, the H-abstraction reaction greatly affects the formation of degradation products.60 According to the FED
394
calculations, the H-abstraction reaction was most likely to occur at C10, C14, and C29. These reactions form unstable
395
molecules that are likely to quickly react with •OH to produce hydroxylated ofloxacin products (Eq. 12). Depending on
396
the atoms adjacent to the hydroxylated site, subsequent dehydrogenation generates a keto-ofloxacin product. These
397
reactions could possibly generate m/z 378 (C10), m/z 376 (C14), or m/z 376 (C29). An m/z 378 product was not identified
398
in the UPLC-Q-TOF MS analysis; therefore, C10 was ruled out as an attack site. Only P13 (m/z 376) exhibited the
399
characteristics of the ketone formation reaction, namely the H-abstraction (-1 H), hydroxyl addition (+1 OH), and
400
dehydrogenation (-2 H) to yield a +14 change in m/z relative to ofloxacin. These results indicate that H-abstraction
401
occurred on C14 or C29. The fragmentation pattern from UPLC-Q-TOF MS analysis did not include the quinolone
402
structure (m/z 261); therefore, the H-abstraction reaction likely occurred at the C29 position according to the mechanism
403
shown in Eq. 12.
404 N
O N
N N
H OH Attack C29
F
405
OFL O
O
N
O
O N
OH
F O
O
N
O N
2H
N
O
O N
OH
F
406
N
N OH
F P13 O
O
O
(Eq. 12)
407 408
Pathway III was the aldehyde reaction. In this case, 1O2 attacked the C14 position in ofloxacin to form endoperoxide (see
409
Eq. 13), which decomposed to form acetaldehyde groups in a similar manner as that reported by Wahab et al. for pACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 28
410
chlorophenol.51 The resulting product is P8 in Fig. 6. As indicated in Eq. 13, the acetaldehyde groups in P8 were prone to
411
H+ attack and subsequent hydrolysis, resulting in formation of the P11 and P12 isomers. Further decomposition of the
412
aldehyde groups generated P6, which exhibited a strong response in the total ion current chromatogram (see Fig. S7 of the
413
SI). These mechanisms align with previous reports of photocatalysis of fluoroquinolone antibiotics by TiO2.54 Ultimately,
414
the identified transformation products from all three pathways were degraded into smaller, oxygen-containing compounds
415
(e.g., P1 and P3) before being mineralized into water and carbon dioxide.
416 N
1
O2
N N
N
417
OFL
O
O
O P8 O
H N
H+
N
Open ring
N
Hydrdysis
N
O
H2O
N H
O
N N
O
O P12
418
HN P11
N H
HN P6
(Eq. 13)
419 420
Photocatalytic mechanism governing the PDI-sunlight photocatalytic system. Based on the above results and discussion,
421
the mechanism shown in Fig. S10 of the SI (also shown in the graphical abstract) was proposed for ofloxacin degradation
422
in the PDI-sunlight photocatalytic system. An organic semiconductor supramolecular structure was created through self-
423
assembly of PDI molecules via acidification polymerization. The terminal carboxyl group aided in the formation of H-type
424
π-π stacking of supramolecular nanofibers. Under sunlight irradiation, the PDI was excited through absorption of UV-A and
425
visible light, causing the formation of photogenerated electrons and holes (see Eq. 14).
426 427
PDI + hv
PDI + h + + e -
428
(Eq. 14)
429 430
The PDI supramolecular structures demonstrate nanocrystalline, high-order H-type π-π stacking effects, and the electric
431
field caused by the carboxyl-rich side facilitated the transport of electrons to the surface, leading to the high separation of
432
the conduction band e- and valence band h+. Since the conduction band potential of PDI (-0.34 eV, calculated by Eq. 10
433
and Fig. S11 in the SI) was lower than the standard reduction potential of O2/O2•- (-0.33 eV), the conduction band e- can
434
be captured by dissolved oxygen to form O2•- (see Eq. 15).
435 ACS Paragon Plus Environment
Page 19 of 28
Environmental Science & Technology
e - + O2
436
O•2 -
437
(Eq. 15)
438 439
Importantly, the long-range electron delocalization in PDI π-π stacking greatly improved the migration efficiency of
440
photogenerated e-, resulting in a high yield of O2•- at the catalyst-solution interface.31 The O2•- reacts with H+ (Eq. 16) to
441
produce 1O2 (Eq. 17) and •OH (Eq. 18).
442 443
O•2 - + H +
HO•2
444
(Eq. 16)
445 446
2HO•2
H 2O 2 + 1O 2
447
(Eq. 17)
448 449
H 2O 2 + e -
•
OH + OH -
450
(Eq. 18)
451 452
The h+ had a valence band potential of 1.39 eV, suggesting minimal reaction with water or HO- to form •OH. In this case,
453
the •OH measured by ESR likely derived from O2•--initiated reactions. Nevertheless, the reactive species quenching
454
experiments demonstrated the negligible role of •OH in the photocatalysis of ofloxacin. The main contribution to
455
ofloxacin transformation was attributed to direct reaction with h+.
456 457
Effects of water quality and interfering substances on ofloxacin photocatalysis. As shown in Fig. 4a, the observed rate
458
constant for ofloxacin degradation decreased from 0.632 ± 0.033 min-1 at pH 3.0 to 0.095 ± 0.008 min-1 at pH 8.0,
459
indicating that ofloxacin degradation was inhibited at higher pH in the PDI-sunlight photocatalytic system. The acid-base
460
speciation reactions for ofloxacin reveal that the cation, zwitterion, and anion dominate at pH less than 5.6, pH between
461
5.6 and 7.9, and pH above 7.9, respectively (Eqs. 6-9). Ofloxacin deprotonation from the cation to the zwitterion increases
462
the density of the electron cloud around the carboxylic acid group, making it more vulnerable to attack by reactive
463
species. This mechanism was confirmed with the g-C3N4-visible light system (see Fig. 4a); note that the preparation of ACS Paragon Plus Environment
Environmental Science & Technology
Page 20 of 28
464
this material is reported in Text S5 of the SI. However, the PDI-sunlight photocatalytic system showed the opposite trend,
465
suggesting that the stimulatory effects of ofloxacin deprotonation were suppressed by another phenomenon. One possible
466
explanation may involve concurrent deprotonation of PDI nanofibers, which could cause depolymerization of the
467
supramolecular nanofibers with increasing pH and, thus, impede photocatalytic activity. However, the strong hydrogen
468
bond interactions between carboxyl groups of the supramolecular fibers were stable for the tested experimental
469
conditions, as indicated in a previous report.31 For this reason, the mechanism involved with pH effects requires further
470
investigation.
471
(a) 1.2
0.07 0.7 OFL+
OFL-
OFL+/-
0.06 0.6 0.05 0.5
PDI g-C3N4
0.6
0.04 0.4 0.03 0.3
0.4
0.02 0.2
0.2
0.3
kobs (min-1)
0.8
kobs (min-1)
Ionization fraction
1.0
(b) 0.4
0.2
0.1
0.01 0.1
0.0
(c)
Absorbance (a.u.)
472
2
4
6
8
pH
3.5
10
12
0.00 0.0
2.0
DOM (100 mg C L-1) Cl- (10 mg L-1) HCO3- (10 mg L-1)
Visible light absorption
1.5
Milli-Q water Tap water Pearl River water Wastewater effluent
1.0
Cl-
HCO3-
Milli-Q water Tap water Pearl River water Wastewater effluent
0.8 0.6 0.4 0.2
0.5 0.0 200
SO42- NO3 DOM
1.0
NO3- (10 mg L-1)
2.5
Control Cu2+
(d) 1.2
Cu2+ (5 mg L-1) SO42- (10 mg L-1)
3.0
0.0
C/C0 for OFL
0
300
400
500
600
700
0.0
0
1
2
3
4
5
6
7
8
473 474
Wavelength (nm) Time (min) Figure 4. (a) Ionization fractions (primary y-axis) and rate constants for ofloxacin (8 mg L-1) photocatalysis by the g-C3N4-
475
white light system (400-800 nm, 200 mg L-1, secondary y-axis, orange bar) and the PDI-sunlight photocatalytic system (200 mg
476
L-1, secondary y-axis, blue bar) as a function of solution pH; the curves through the rate constant data stem from a species-
477
specific kinetics model similar to that reported by Snowberger et al. (2016)9; (b) rate constants for ofloxacin degradation by
478
sunlight-activated PDI nanofibers in the presence of water quality constituents (e.g., [Cu2+] = 5 mg L-1, [DOM] = 100 mg C L-1,
479
concentrations of all other water quality constituents were 10 mg L-1); (c) UV-vis absorbance spectra of solutions containing the ACS Paragon Plus Environment
Page 21 of 28
Environmental Science & Technology
480
water quality constituents included in (b); and, (d) ofloxacin (8 mg L-1) degradation in different water and wastewater matrices
481
by 200 mg L-1 of PDI nanofibers for sunlight irradiation at pH 5.6. All error bars indicate one standard deviation.
482 483
The effects of water quality constituents on the rate constant for ofloxacin photocatalysis are shown in Fig. 4b. In the
484
presence of SO42- and NO3-, the rate constant for ofloxacin degradation was not significantly affected; furthermore, only
485
minor differences were observed for Cl- and HCO3-. However, the Cu2+ cation significantly reduced the transformation
486
kinetics of ofloxacin, a finding that was attributed to Cu2+ scavenging of O2•- (𝑘"Cu2 + ,O•2 ― = 8.1×109 M−1s−1) and 1O2
487
(𝑘"Cu2 + ,1O2 = 1.2×109 M-1s-1).61,62 Martínez et al. reported the second-order rate constant for ofloxacin reaction with 1O2 to
488
be 5.6×106 M-1s-1.63 Then, for the experimental conditions used here, Cu2+ reacted with approximately 1.5× more 1O2 than
489
ofloxacin, resulting in slower ofloxacin degradation (kobs = 0.334 ± 0.195 min-1) compared to the control solution (kobs =
490
0.126 ± 0.007 min-1). The relative impacts of other metals on ofloxacin degradation can be investigated by comparing the
491
rate constants for metal ion reaction with 1O2 and O2•-. For example, the second-order rate constant for Fe3+ with O2•- is
492
2.4×106 M−1s−1, and rate constants for reaction of Fe3+-complexes with 1O2 are on the order of 109 M−1s−1.64,65 In this case,
493
competition effects from Fe3+ will be similar to Cu2+ for 1O2-based reactions and less than Cu2+ for O2•-. So, Fe3+ will exert
494
a lower impact than Cu2+ (at the same molar concentration). However, we note that iron can also (i) precipitate on the PDI
495
supramolecular photocatalyst to reduce activity or (ii) enable Fenton reactions in the presence of hydrogen peroxide to
496
increase activity. The impacts of environmentally-relevant metal concentrations are also inherent in the results described
497
below for ofloxacin degradation in real surface water and wastewater effluent. The presence of DOM inhibited the
498
photocatalytic degradation of ofloxacin, presumably due to competition for reaction with the reactive species and light
499
screening effects in the visible region (see Fig. 4c).
500 501
To evaluate the suitability of the PDI-sunlight photocatalytic system to degrade antibiotics under environmental
502
conditions, ofloxacin transformation was studied in different water matrices. The primary chemical properties associated
503
with known impacts on ofloxacin degradation kinetics (from Fig. 4b) are shown in Table S6 of the SI for the four water
504
sources. As indicated in Fig. 4d and Table S7 in the SI, the observed rate constants for ofloxacin degradation in ultrapure
505
water, tap water, Pearl River water, and wastewater effluent were 0.334 ± 0.195, 0.250 ± 0.018, 0.208 ± 0.014, and 0.224
506
± 0.009 min-1, respectively. Unsurprisingly, ultrapure water, which contains negligible background inorganic or organic
507
species, demonstrated the fastest ofloxacin transformation due to the absence of reactive species scavenging or light ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 28
508
screening. Ofloxacin degradation in tap water was slightly inhibited compared to ultrapure water, presumably due to the
509
low TOC, Cl-, and NO3- content that caused some scavenging effects. A greater extent of interference was observed in the
510
Pearl River water and wastewater effluent, for which the ofloxacin transformation decreased by 17.8% and 15.9%,
511
respectively, over the 7-min irradiation period. The TOC concentrations in the Pearl River water and wastewater effluent
512
were 1.75× and 1.09× higher than that of the tap water, increasing the scavenging of reactive species by DOM and
513
inhibiting ofloxacin degradation. The higher DOM content of the Pearl River water and wastewater effluent likely exerted
514
light-screening effects that further inhibited ofloxacin transformation. Nevertheless, the PDI-sunlight photocatalytic
515
system was successfully applied to ofloxacin treatment in tap water, surface water, and wastewater effluent with half-lives
516
of 2.77, 3.33, and 3.09 min, respectively. These results confirm the suitability of this novel system to ameliorate concerns
517
with antibiotics in real drinking water and wastewater matrices.
518 519
ACKNOWLEDGEMENTS.
520 521
This work was supported by the Major Science and Technology Program for Water Pollution Control and Treatment in
522
China (No. 2017ZX07202001), Program for Changjiang Scholars and Innovative Research Team in University, and China
523
Postdoctoral Science Foundation (Grant No. 2017M620796).
524 525
SUPPORTING INFORMATION.
526 527
Five sections of text describe chemical reagents, characterization methods, reactive species determination, ofloxacin analysis
528
and identification of transformation products, and preparation of g-C3N4. Seven tables provide fluoroquinolone analysis
529
parameters, effects of different light sources, contributions of reactive species, transformation product information, FED
530
calculations, water quality, and rate constants for ofloxacin degradation in different wastewater matrices. Eleven figures
531
show the characterization results from BET, FTIR, XPS, and DRS analyses, sequential photocatalysis experiments, ESR
532
spectra, quenching experiments, TOC, total ion current chromatogram, MS/MS fragmentation, atomic numbering,
533
proposed photocatalytic mechanism, and valence band XPS spectra.
534 535
REFERENCES ACS Paragon Plus Environment
Page 23 of 28
Environmental Science & Technology
536 537 538 539 540 541
(1) Calamari, D.; Zuccato, E.; Castiglioni, S.; Bagnati, R.; Fanelli, R., Strategic survey of therapeutic drugs in the rivers Po and Lambro in northern Italy. Environ. Sci. Technol. 2003, 37, (7), 1241-1248. (2) He, K.; Blaney, L., Systematic optimization of an SPE with HPLC-FLD method for fluoroquinolone detection in wastewater. J. Hazard. Mater. 2015, 282, 96-105. (3) Fatta-Kassinos, D.; Vasquez, M. I.; Kummerer, K., Transformation products of pharmaceuticals in surface waters and
542
wastewater formed during photolysis and advanced oxidation processes - Degradation, elucidation of byproducts and
543
assessment of their biological potency. Chemosphere 2011, 85, (5), 693-709.
544
(4) Radjenovic, J.; Petrovic, M.; Barcelo, D., Fate and distribution of pharmaceuticals in wastewater and sewage sludge of the
545
conventional activated sludge (CAS) and advanced membrane bioreactor (MBR) treatment. Water Res. 2009, 43, (3), 831-
546
841.
547
(5) He, K.; Soares, A. D.; Adejumo, H.; McDiarmid, M.; Squibb, K.; Blaney, L., Detection of a wide variety of human and
548
veterinary fluoroquinolone antibiotics in municipal wastewater and wastewater-impacted surface water. J. Pharmaceut.
549
Biomed. 2015, 106, 136-143.
550 551 552 553 554 555 556
(6) Volmer, D. A.; Mansoori, B.; Locke, S. J., Study of 4-quinolone antibiotics in biological samples by short-column liquid chromatography coupled with electrospray ionization tandem mass spectrometry. Anal. Chem. 1997, 69, (20), 4143-4155. (7) Salgado, P.; Melin, V.; Duran, Y.; Mansilla, H.; Contreras, D., The Reactivity and Reaction Pathway of Fenton Reactions Driven by Substituted 1,2-Dihydroxybenzenes. Environ. Sci. Technol. 2017, 51, (7), 3687-3693. (8) Li, B.; Zhang, T., Biodegradation and Adsorption of Antibiotics in the Activated Sludge Process. Environ. Sci. Technol. 2010, 44, (9), 3468-3473. (9) Snowberger, S.; Adejumo, H.; He, K.; Mangalgiri, K. P.; Hopanna, M.; Soares, A. D.; Blaney, L., Direct Photolysis of
557
Fluoroquinolone Antibiotics at 253.7 nm: Specific Reaction Kinetics and Formation of Equally Potent Fluoroquinolone
558
Antibiotics. Environ. Sci. Technol. 2016, 50, (17), 9533-9542.
559
(10) Feng, M. B.; Wang, X. H.; Chen, J.; Qu, R. J.; Sui, Y. X.; Cizmas, L.; Wang, Z. Y.; Sharma, V. K., Degradation of
560
fluoroquinolone antibiotics by ferrate(VI): Effects of water constituents and oxidized products. Water Res. 2016, 103, 48-
561
57.
562
(11) Gao, Y. X.; Yu, G.; Liu, K.; Deng, S. B.; Wang, B.; Huang, J.; Wang, Y. J., Integrated adsorption and visible-light
563
photodegradation of aqueous clofibric acid and carbamazepine by a Fe-based metal-organic framework. Chem. Eng. J.
564
2017, 330, 157-165.
565
(12) Ding, H. Y.; Wu, Y. X.; Zou, B. C.; Lou, Q.; Zhang, W. H.; Zhong, J. Y.; Lu, L.; Dai, G. F., Simultaneous removal and ACS Paragon Plus Environment
Environmental Science & Technology
Page 24 of 28
566
degradation characteristics of sulfonamide, tetracycline, and quinolone antibiotics by laccase-mediated oxidation coupled
567
with soil adsorption. J. Hazard. Mater. 2016, 307, 350-358.
568
(13) Peng, G. L.; Li, T.; Ai, B. X.; Yang, S. G.; Fu, J. W.; He, Q.; Yu, G.; Deng, S. B., Highly efficient removal of enrofloxacin
569
by magnetic montmorillonite via adsorption and persulfate oxidation. Chem. Eng. J. 2018,
570
https://doi.org/10.1016/j.cej.2018.10.190.
571 572 573 574 575 576 577
(14) Kim, H. I.; Choi, Y.; Hu, S.; Choi, W.; Kim, J. H., Photocatalytic hydrogen peroxide production by anthraquinoneaugmented polymeric carbon nitride. Appl. Catal. B-Environ. 2018, 229(5), 121-129. (15) Miao, H.; Yang, J.; Wei, Y.; Li, W.; Zhu, Y., Visible-light photocatalysis of PDI nanowires enhanced by plasmonic effect of the gold nanoparticles. Appl. Catal. B-Environ. 2018, 239(30), 61-67. (16) Zhao, G.; Cheng, Y.; Wu, Y.; Xu, X.; Hao, X. New 2D carbon nitride organic materials synthesis with huge‐Application prospects in CN photocatalyst. Small, 2018, 14(15), e1704138. (17) Zhang, Q. X.; Chen, P.; Zhuo, M. H.; Wang, F. L.; Su, Y. H.; Chen, T. S.; Yao, K.; Cai, Z. W.; Lv, W. Y.; Liu, G. G.,
578
Degradation of indometacin by simulated sunlight activated CDs-loaded BiPO4 photocatalyst: Roles of oxidative species.
579
Appl. Catal. B-Environ. 2018, 221, 129-139.
580
(18) Chen, P.; Wang, F. L.; Chen, Z. F.; Zhang, Q. X.; Su, Y. H.; Shen, L. Z.; Yao, K.; Liu, Y.; Cai, Z. W.; Lv, W. Y.; Liu, G.
581
G., Study on the photocatalytic mechanism and detoxicity of gemfibrozil by a sunlight-driven TiO2/carbon dots
582
photocatalyst: The significant roles of reactive oxygen species. Appl. Catal. B-Environ. 2017, 204, 250-259.
583
(19) Hedstrom, S.; Chaudhuri, S.; La Porte, N. T.; Rudshteyn, B.; Martinez, J. F.; Wasielewski, M. R.; Batista, V. S.,
584
Thousandfold Enhancement of Photoreduction Lifetime in Re(bpy)(CO)3 via Spin-Dependent Electron Transfer from a
585
Perylenediimide Radical Anion Donor. J. Am. Chem. Soc. 2017, 139, (46), 16466-16469.
586
(20) Guo, Y. K.; Li, Y. K.; Awartani, O.; Han, H.; Zhao, J. B.; Ade, H.; Yan, H.; Zhao, D. H., Improved Performance of All-
587
Polymer Solar Cells Enabled by Naphthodiperylenetetraimide-Based Polymer Acceptor. Adv. Mater. 2017, 29, (1700309),
588
1-6.
589 590 591
(21) Zang, L.; Liu, R. C.; Holman, M. W.; Nguyen, K. T.; Adams, D. M., A single-molecule probe based on intramolecular electron transfer. J. Am. Chem. Soc. 2002, 124, (36), 10640-10641. (22) Ego, C.; Marsitzky, D.; Becker, S.; Zhang, J. Y.; Grimsdale, A. C.; Mullen, K.; MacKenzie, J. D.; Silva, C.; Friend, R. H.,
592
Attaching perylene dyes to polyfluorene: Three simple, efficient methods for facile color tuning of light-emitting
593
polymers. J. Am. Chem. Soc. 2003, 125, (2), 437-443.
594
(23) Sauer, M., Single-molecule-sensitive fluorescent sensors based on photoinduced intramolecular charge transfer. Angew. ACS Paragon Plus Environment
Page 25 of 28
595
Environmental Science & Technology
Chem. Int. Edit. 2003, 42, (16), 1790-1793.
596
(24) Zhao, Z.; Gao, S. M.; Zheng, X. Y.; Zhang, P. F.; Wu, W. T.; Kwok, R. T. K.; Xiong, Y.; Leung, N. L. C.; Chen, Y. C.;
597
Gao, X. K.; Lam, J. W. Y.; Tang, B. Z., Rational Design of Perylenediimide-Substituted Triphenylethylene to Electron
598
Transporting Aggregation-Induced Emission Luminogens (AIEgens) with High Mobility and Near-Infrared Emission.
599
Adv. Funct. Mater. 2018, 28, (11), 9.
600
(25) Kolhe, N. B.; Devi, R. N.; Senanayak, S. P.; Jancy, B.; Narayan, K. S.; Asha, S. K., Structure engineering of naphthalene
601
diimides for improved charge carrier mobility: self-assembly by hydrogen bonding, good or bad? J. Mater. Chem. 2012,
602
22, (30), 15235-15246.
603
(26) Zeng, L.; Liu, T.; He, C.; Shi, D. Y.; Zhang, F. L.; Duan, C. Y., Organized Aggregation Makes Insoluble Perylene Diimide
604
Efficient for the Reduction of Aryl Halides via Consecutive Visible Light-Induced Electron-Transfer Processes. J. Am.
605
Chem. Soc. 2016, 138, (12), 3958-3961.
606
(27) Weingarten, A. S.; Kazantsev, R. V.; Palmer, L. C.; McClendon, M.; Koltonow, A. R.; Samuel, A. P. S.; Kiebala, D. J.;
607
Wasielewski, M. R.; Stupp, S. I., Self-assembling hydrogel scaffolds for photocatalytic hydrogen production. Nat. Chem.
608
2014, 6, (11), 964-970.
609 610 611 612
(28) Kunz, V.; Stepanenko, V.; Wurthner, F., Embedding of a ruthenium(II) water oxidation catalyst into nanofibers via selfassembly. Chem. Commun. 2015, 51, (2), 290-293. (29) Chen, S.; Li, Y. X.; Wang, C. Y., Visible-light-driven photocatalytic H2 evolution from aqueous suspensions of perylene diimide dye-sensitized Pt/TiO2 catalysts. RSC Adv. 2015, 5, (21), 15880-15885.
613
(30) Chen, S.; Wang, C.; Bunes, B. R.; Li, Y. X.; Wang, C. Y.; Zang, L., Enhancement of visible-light-driven photocatalytic H2
614
evolution from water over g-C3N4 through combination with perylene diimide aggregates. Appl. Catal. A-Gen. 2015, 498,
615
63-68.
616 617 618 619 620 621 622 623 624
(31) Wang, J.; Shi, W.; Liu, D.; Zhang, Z. J.; Zhu, Y. F.; Wang, D., Supramolecular organic nanofibers with highly efficient and stable visible light photooxidation performance. Appl. Catal. B-Environ. 2017, 202, 289-297. (32) Wei, W. Q.; Liu, D.; Wei, Z.; Zhu, Y. F., Short-Range pi-pi Stacking Assembly on P25 TiO2 Nanoparticles for Enhanced Visible-Light Photocatalysis. ACS Catal. 2017, 7, (1), 652-663. (33) Wei, W.; Wei, Z.; Liu, D.; Zhu, Y., Enhanced visible-light photocatalysis via back-electron transfer from palladium quantum dots to perylene diimide. Appl. Catal. B-Environ. 2018, 230, 49-57.. (34) Liu, D.; Wang, J.; Bai, X.; Zong, R.; Zhu, Y., Self-Assembled PDINH Supramolecular System for Photocatalysis under Visible Light. Adv. Mater. 2016, 28, (33), 7284-7290. (35) Duran, A.; Monteagudo, J. M.; San Martin, I., Photocatalytic treatment of an industrial effluent using artificial and solar ACS Paragon Plus Environment
Environmental Science & Technology
625 626 627 628
Page 26 of 28
UV radiation: an operational cost study on a pilot plant scale. J. Environ. Manage. 2012, 98, 1-4. (36) Stumm W, Morgan J J. Aquatic chemistry: chemical equilibria and rates in natural waters. Cram101 Textbook Outlines to Accompany, 1995, 179(11):A277. (37) Chen, P.; Zhang, Q.; Su, Y.; Shen, L.; Wang, F.; Liu, H.; Liu, Y.; Cai, Z.; Lv, W.; Liu, G., Accelerated photocatalytic
629
degradation of diclofenac by a novel CQDs/BiOCOOH hybrid material under visible-light irradiation: Dechloridation,
630
detoxicity, and a new superoxide radical model study. Chem. Eng. J. 2018, 332, 737-748.
631
(38) Wang, Y.; Bayazit, M. K.; Moniz, S. J. A.; Ruan, Q.; Lau, C. C.; Martsinovich, N.; Tang, J., Linker-controlled polymeric
632
photocatalyst for highly efficient hydrogen evolution from water. Energ. & Environ. Sci. 2017, 10, (7), 1643-1651.
633
(39) Wang, F. L.; Feng, Y. P.; Chen, P.; Wang, Y. F.; Su, Y. H.; Zhang, Q. X.; Zeng, Y. Q.; Xie, Z. J.; Liu, H. J.; Liu, Y.; Lv,
634
W. Y.; Liu, G. G., Photocatalytic degradation of fluoroquinolone antibiotics using ordered mesoporous g-C3N4 under
635
simulated sunlight irradiation: Kinetics, mechanism, and antibacterial activity elimination. Appl. Catal. B-Environ. 2018,
636
227, 114-122.
637
(40) Hapeshi, E.; Achilleos, A.; Vasquez, M. I.; Michael, C.; Xekoukoulotakis, N. P.; Mantzavinos, D.; Kassinos, D., Drugs
638
degrading photocatalytically: Kinetics and mechanisms of ofloxacin and atenolol removal on titania suspensions. Water
639
Res. 2010, 44, (6), 1737-1746.
640 641 642 643 644 645 646 647 648 649 650 651 652
(41) Bhatia, V.; Ray, A. K.; Dhir, A., Enhanced photocatalytic degradation of ofloxacin by co-doped titanium dioxide under solar irradiation. Sep. Purif. Technol. 2016, 161, 1-7. (42) Zhang, W.; Liu, Y.; Li, C., Photocatalytic degradation of ofloxacin on Gd2Ti2O7, supported on quartz spheres[J]. J. Phys. Chem. Solids. 2018, 118, 144-149. (43) Kaur, M.; Mehta, S. K.; Kansal, S. K., Visible light driven photocatalytic degradation of ofloxacin and malachite green dye using cadmium sulphide nanoparticles. J. Environ. Chem. Eng. 2018, 6, (3), 3631-3639. (44) Kaur, A.; Umar, A.; Anderson, W. A.; Kansal, S. K., Facile synthesis of CdS/TiO2 nanocomposite and their catalytic activity for ofloxacin degradation under visible illumination. J. Photoch. Photobio. A 2018, 360, 34-43. (45) Di, J.; Xia, J.; Ji, M.; Li, H.; Xu, H.; Li, H.; Chen, R., The synergistic role of carbon quantum dots for the improved photocatalytic performance of Bi2MoO6. Nanoscale 2015, 7, (26), 11433-11443. (46) Bing, J. S.; Hu, C.; Zhang, L. L., Enhanced mineralization of pharmaceuticals by surface oxidation over mesoporous gamma-Ti-Al2O3 suspension with ozone. Appl. Catal. B-Environ. 2017, 202, 118-126. (47) Chen, P.; Wang, F.; Zhang, Q.; Su, Y.; Shen, L.; Yao, K.; Chen, Z.-F.; Liu, Y.; Cai, Z.; Lv, W.; Liu, G., Photocatalytic
653
degradation of clofibric acid by g-C3N4/P-25 composites under simulated sunlight irradiation: The significant effects of
654
reactive species. Chemosphere 2017, 172, 193-200. ACS Paragon Plus Environment
Page 27 of 28
655
Environmental Science & Technology
(48) Fang, H.; Gao, Y.; Li, G.; An, J.; Wong, P.-K.; Fu, H.; Yao, S.; Nie, X.; An, T., Advanced Oxidation Kinetics and
656
Mechanism of Preservative Propylparaben Degradation in Aqueous Suspension of TiO2 and Risk Assessment of Its
657
Degradation Products. Environ. Sci. Technol. 2013, 47, (6), 2704-2712.
658
(49) Zhang, Z.; Wang, J.; Liu, D.; Luo, W.; Zhang, M.; Jiang, W.; Zhu, Y., Highly Efficient Organic Photocatalyst with Full
659
Visible Light Spectrum through π−π Stacking of TCNQ-PTCDI. Acs Appl. Mater. Inter. 2016, 8, (44), 30225-30231.
660
(50) Ji, Y.; Zhou, L.; Ferronato, C.; Yang, X.; Salvador, A.; Zeng, C.; Chovelon, J.-M., Photocatalytic degradation of atenolol in
661
aqueous titanium dioxide suspensions: Kinetics, intermediates and degradation pathways. J. Photoch. Photobio. A. 2013,
662
254, 35-44.
663 664 665 666 667 668 669
(51) Wahab, H. S.; Bredow, T.; Aliwi, S. M., A computational study on the adsorption and ring cleavage of para-chlorophenol on anatase TiO2 surface. Surf. Sci. 2009, 603, (4), 664-669. (52) Kusari, S.; Prabhakaran, D.; Lamshoeft, M.; Spiteller, M., In vitro residual anti-bacterial activity of difloxacin, sarafloxacin and their photoproducts after photolysis in water. Environ. Pollut. 2009, 157, (10), 2722-2730. (53) Wammer, K. H.; Korte, A. R.; Lundeen, R. A.; Sundberg, J. E.; McNeill, K.; Arnold, W. A., Direct photochemistry of three fluoroquinolone antibacterials: Norfloxacin, ofloxacin, and enrofloxacin. Water Res. 2013, 47, (1), 439-448. (54) Sturini, M.; Speltini, A.; Maraschi, F.; Profumo, A.; Pretali, L.; Irastorza, E. A.; Fasani, E.; Albini, A., Photolytic and
670
photocatalytic degradation of fluoroquinolones in untreated river water under natural sunlight. Appl. Catal. B-Environ.
671
2012, 119, 32-39.
672 673 674 675 676
(55) Moeller, M. N.; Hatch, D. M.; Kim, H.-Y. H.; Porter, N. A., Superoxide Reaction with Tyrosyl Radicals Generates paraHydroperoxy and para-Hydroxy Derivatives of Tyrosine. J. Am. Chem. Soc. 2012, 134, (40), 16773-16780. (56) Yin, L.; Shen, Z.; Niu, J.; Chen, J.; Duan, Y., Degradation of Pentachlorophenol and 2,4-Dichlorophenol by Sequential Visible-light Driven Photocatalysis and Laccase Catalysis. Environ. Sci. Technol. 2010, 44, (23), 9117-9122. (57) Zhang, Q.; Chen, P.; Tan, C.; Chen, T.; Zhuo, M.; Xie, Z.; Wang, F.; Liu, H.; Cai, Z.; Liu, G.; Lv, W., A photocatalytic
677
degradation strategy of PPCPs by a heptazine-based CN organic polymer (OCN) under visible light. Environ. Sci. Nano
678
2018, 5, 2325-2336.
679 680 681 682
(58) Wang, P.; He, Y.-L.; Huang, C.-H., Oxidation of fluoroquinolone antibiotics and structurally related amines by chlorine dioxide: Reaction kinetics, product and pathway evaluation. Water Res. 2010, 44, (20), 5989-5998. (59) Zhang, H. C.; Huang, C. H., Oxidative transformation of fluoroquinolone antibacterial agents and structurally related amines by manganese oxide. Environ. Sci. Technol. 2005, 39, (12), 4474-4483.
683
(60) An, T.; An, J.; Gao, Y.; Li, G.; Fang, H.; Song, W., Photocatalytic degradation and mineralization mechanism and toxicity
684
assessment of antivirus drug acyclovir: Experimental and theoretical studies. Appl. Catal. B-Environ. 2015, 164, 279-287. ACS Paragon Plus Environment
Environmental Science & Technology
685
(61) Munoz, V. A.; Ferrari, G. V.; Paulina Montana, M.; Miskoski, S.; Garcia, N. A., Effect of Cu2+-complexation on the
686
scavenging ability of chrysin towards photogenerated singlet molecular oxygen (O2((1△g)). Possible biological
687
implications. J. Photoch. Photobio. B 2016, 162, 597-603.
688 689 690 691 692
Page 28 of 28
(62) Bielski, B.H.; Cabelli, D.E.; Arudi, R.L.; Ross, A.B. Reactivity of HO2/O2•- radicals in aqueous solution. Journal of Physical and Chemical Reference Data, 1985, 14(4), 1041-1100. (63) Martinez, L. J.; Sik, R. H.; Chignell, C. F., Fluoroquinolone antimicrobials: Singlet oxygen, superoxide and phototoxicity. Photochem. Photobiol. 1998, 67, (4), 399-403. (64) Rush, J.D.; Bielski, B.H.J. Pulse radiolytic studies of the reaction of perhydroxyl/superoxide O2- with iron(II)/iron(III)
693
ions. The reactivity of HO2/O2- with ferric ions and its implication on the occurrence of the Haber-Weiss reaction. J.
694
Phys. Chem., 1985, 89(23), 5062-5066.
695 696
(65) Venediktov, E.A.; Krasnovskiĭ, A.A. Jr. Quenching of the luminescence of singlet molecular oxygen by complexes of porphyrins and highly-charged metal ions. Biofizika, 1980, 25(2), 336-337.
697
ACS Paragon Plus Environment