Subscriber access provided by TULANE UNIVERSITY
Remediation and Control Technologies
Efficient Destruction of Pollutants in Water by a Dual-Reaction-Center FentonLike Process over Carbon Nitride Compounds (CN)-Complexed Cu(II)-CuAlO2 Lai Lyu, Dengbiao Yan, Guangfei Yu, Wenrui Cao, and Chun Hu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b06545 • Publication Date (Web): 15 Mar 2018 Downloaded from http://pubs.acs.org on March 15, 2018
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 33
Environmental Science & Technology
1
Efficient Destruction of Pollutants in Water by a Dual-Reaction-Center
2
Fenton-Like Process over Carbon Nitride Compounds (CN)-Complexed
3
Cu(II)-CuAlO2
4
Lai Lyuab, Dengbiao Yanc, Guangfei Yub, Wenrui Caoa and Chun Hu*,ab
5 6
a
7
Water Quality and Conservation of the Pearl River Delta, Ministry of Education,
8
School of Environmental Science and Engineering, Guangzhou University,
9
Guangzhou 510006, China b
10 11 12 13
Research Institute of Environmental Studies at Greater Bay, Key Laboratory for
Key Laboratory of Drinking Water Science and Technology, Research Center for
Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China c
School of Environmental and Chemical Engineering, Tianjin Polytechnic University, Tianjin 300387, China
14 15
*Corresponding author Tel: +86-10-62849628; fax: +86-10-62923541;
16
e-mail:
[email protected] /
[email protected] 17
1 ACS Paragon Plus Environment
Environmental Science & Technology
18
ABSTRACT
19
Carbon nitride compounds (CN) complexed with the in-situ produced Cu(II) on
20
the surface of CuAlO2 substrate (CN-Cu(II)-CuAlO2) is prepared via a surface growth
21
process for the first time, which exhibits exceptionally high activity and efficiency for
22
the degradation of the refractory pollutants in water through a Fenton-like process in a
23
wide pH range. The reaction rate for BPA removal is ~25 times higher than that of the
24
CuAlO2. According to the characterization, Cu(II) generation on the surface of
25
CuAlO2 during the surface growth process result in the marked decrease of the
26
surface oxygen vacancies and the formation of the C-O-Cu bridges between CN and
27
Cu(II)-CuAlO2 in the catalyst. The electron paramagnetic resonance (EPR) analysis
28
and density functional theory (DFT) calculations demonstrate that the dual reaction
29
centers are produced around the Cu and C sites due to the cation–π interactions
30
through the C-O-Cu bridges in CN-Cu(II)-CuAlO2. During the Fenton-like reactions,
31
the electron-rich center around Cu is responsible for the efficient reduction of H2O2
32
to •OH, and the electron-poor center around C captures electrons from H2O2 or
33
pollutants and diverts them to the electron-rich area via the C-O-Cu bridge. Thus, the
34
catalyst exhibits excellent catalytic performance for the refractory pollutant
35
degradation. This study can deepen our understanding on the enhanced Fenton
36
reactivity for water purification through functionalizing with organic solid-phase
37
ligands on the catalyst surface.
38
2 ACS Paragon Plus Environment
Page 2 of 33
Page 3 of 33
Environmental Science & Technology
39
INTRODUCTION
40
Growing discharges of nonbiodegradable and persistent organic pollutants
41
adversely affect the environment and potentially threaten human health.1, 2 Advanced
42
oxidation processes (AOPs) have been applied for eliminating the recalcitrant organic
43
pollutants in water due to the generation of the highly reactive free radicals.3, 4 Among
44
various AOPs, Fenton reaction (Fe2+/H2O2) is an especially powerful and
45
environmentally friendly method,5, 6 since it does not require any special energy input
46
and can rapidly destruct the pollutant structure with the generated highly aggressive
47
free hydroxyl radical (•OH).7-9 Unfortunately, the classical Fenton reaction suffers
48
from its drawbacks such as the poor recyclability,10 the narrow working pH range
49
and the accumulation of Fe-containing sludge,12, 13 which restrict its wide applications.
50
To avoid or minimize these drawbacks, many researchers focus on looking for
51
efficient heterogeneous Fenton-like catalysts as alternatives to the homogeneous
52
process.14-18
11
53
However, few of the developed heterogeneous Fenton catalysts exhibit good
54
activity and high catalytic efficiency under neutral conditions, which is due to the
55
rate-limiting step upon the reduction of the stationary M(n+m)+ to Mn+ (M represent
56
metal species) by oxidizing H2O2 on the solid-liquid interface.9, 19-21 In addition, in
57
this step, H2O2 was finally decomposed into O2•− or O2, leading to invalid
58
consumption of H2O2.8, 12 Moreover, surface oxygen vacancies (Vo) were easy to be
59
produced in the synthesis process of the metal-containing catalyst, which often
60
promote the rapid decomposition of H2O2 to H2O and O2,19, 22, 23 leading low H2O2
61
utilization. The problems above are contributed to the dependency for the redox of the
62
metal ions which are difficult to solve by conventional means.24
63
Recently, our research20 has revealed that constructing dual reaction centers with 3 ACS Paragon Plus Environment
Environmental Science & Technology
Page 4 of 33
64
the non-uniform distribution of the electrons in a catalyst by lattice-doping method is
65
essential for overcoming the limitations of the classical Fenton reaction, which gives
66
us inspirations for in-depth exploration of more efficient dual-center Fenton catalysts
67
using other means for tuning the electron distribution. Cation–π interactions are
68
important intermolecular binding forces, which are relevant for the aromatic
69
recognition in the chemical and biological processes.25,
70
particularly crucial process in the cation–π interactions for transition-metal ions.26 It is
71
reasonable to believe that creating a special bonding bridge enhancing the cation–π
72
interaction through the charge transfer is the key for inducing the non-uniform
73
distribution of the electrons on the catalyst surface.
26
Charge transfer is a
74
Herein, a novel Fenton catalyst consisting of aromatic carbon nitride compounds
75
(CN) complexed with in-situ produced Cu(II) on the surface of CuAlO2 substrate
76
(CN-Cu(II)-CuAlO2)
77
CN-Cu(II)-CuAlO2 exhibits excellent Fenton-like activity and efficiency in a wide pH
78
range for the degradation of the refractory pollutants, as demonstrated with phenol,
79
2-chlorphenol, ibuprofen, phenytoin and bisphenol A (BPA). The generation of Cu(II)
80
on the surface of CuAlO2 and the formation of the key C-O-Cu bridges between CN
81
and Cu(II)-CuAlO2 in the catalyst during the surface complexation process were
82
verified and characterized by several characterization techniques. The electron-rich
83
Cu and electron-poor C centers (i.e., dual reaction centers) are produced due to the
84
cation–π interactions through the C-O-Cu bridges in CN-Cu(II)-CuAlO2, as
85
confirmed by the electron paramagnetic resonance (EPR) analysis and density
86
functional theory (DFT) calculations. A preliminary effort to identify a correlation
87
between the surface electron properties of CN-Cu(II)-CuAlO2 and catalytic
88
performance has been undertaken, and a dual-reaction-center mechanism for the
is
prepared
via
a
surface
4 ACS Paragon Plus Environment
complexation
process.
Page 5 of 33
Environmental Science & Technology
89
Fenton-like reaction has been proposed.
90 91
EXPERIMENTAL SECTION
92
Preparation of catalysts. CN-Cu(II)-CuAlO2 is prepared by two steps. The first one
93
is to synthesize CuAlO2. In a typical procedure, 4.0 g of Cu(CH3COO)2•H2O was
94
dissolved in 30 mL of ethylene glycol to form solution A. Then, 7.5 g of
95
Al(NO3)3•9H2O was dissolved in 15 mL of ethylene glycol to form solution B. Next,
96
solution B was added to solution A and stirred at room temperature for 2 h. After the
97
mixture is kept at 150 ºC, vaporizing the ethylene glycol and volatile substances and
98
getting the dry gel precursor. Then, the dry gel precursor was calcined in a muffle
99
furnace at 1200 ºC in air for 2 h. The resulting solid was washed with deionized water
100
several times. After drying at 60 ºC, CuAlO2 was obtained. Next, the prepared
101
CuAlO2 and urea were mixed in deionized water with the mass ratio of 1:5. After that,
102
the mixture was calcined for 4 h at 550 ºC in a muffle furnace. After natural cooling,
103
the composite was obtained, which was designated as CN-Cu(II)-CuAlO2. The
104
pristine g-C3N4 was prepared using the same procedure without the addition of
105
CuAlO2.
106
Procedures and analysis. The Fenton-like catalytic experiments were carried out
107
under ambient conditions by using BPA, phenol, 2-chlorphenol, ibuprofen and
108
phenytoin as model contaminants. Their structures are shown in Figure S1
109
(Supporting Information, SI). In a typical experiment, 1.0 g L-1 catalysts were
110
dispersed in 25 mg L-1 pollutant solutions and the whole process is keep 35 ºC
111
(summer room temperature). The mixture was magnetic stirred for 30 min to establish
112
the adsorption/desorption equilibrium. Then, H2O2 was added to the pollutant
113
suspensions under stirring throughout the experiment. At given time intervals, 4 mL 5 ACS Paragon Plus Environment
Environmental Science & Technology
Page 6 of 33
114
aliquots were collected and filtered using a Millipore filter (pore size 0.22 µm) for
115
analysis. Then, the enzyme catalase was added to destroy the residual H2O2.
116
The pollutant concentrations were measured by a high performance liquid
117
chromatography (HPLC, 1200 series; Agilent) with an auto-sampler, a Zorbax SB-Aq
118
column (4.6 × 250 mm, 5 µm; Agilent) and an UV detector at the wavelength of 225
119
nm. The mobile phase was a mixture of methanol/water and was operated at a
120
flow-rate of 1.0 mL min-1. The total organic carbon (TOC) was determined by a
121
TOC-VCPH analyzer (Shimadzu) using high-temperature combustion. The H2O2
122
concentration was determined using the reported DPD method.27 The amount of
123
metallic ions releasing from the catalysts during the reaction were measured using the
124
inductively coupled plasma optical emission spectrometry (ICP-OES) on an Optima
125
2000 (Perkin Elmer, U.S.A.). To test the catalyst stability and recyclability, the used
126
CN-Cu(II)-CuAlO2 sample was filtered, washed with water, and dried at 60 ◦C. The
127
catalyst was continued to be used in the second cycle.
128
BMPO-trapped
EPR
signals
were
detected
in
different
air-saturated
129
methanol/aqueous dispersions of the corresponding samples. Typically, in the absence
130
of H2O2, 0.05 g of the prepared powder sample was added to 1 mL of water (for
131
detecting •OH) or water/methanol (10%/90%, V/V, for detecting O2•−). 20 µL of
132
BMPO (250 mM) was added and held for 5 min. Then, the solution was sucked into
133
the capillary. The EPR spectra were recorded on a Bruker A300-10/12 EPR
134
spectrometer at room temperature. In the presence of H2O2, 0.01 g of the prepared
135
powder sample was added to 1 mL of water (for detecting •OH) or methanol (for
136
detecting O2•−). Then, 100 µL of the above suspension, 10 µL of BMPO (250 mM)
137
and 10 µL of H2O2 (30%, w/w) were mixed and held for 5 min. The solution was
138
sucked into the capillary to carry out EPR detection. 6 ACS Paragon Plus Environment
Page 7 of 33
Environmental Science & Technology
139
The attenuated total reflection Fourier-transform infrared spectroscopy
140
(ATR-FTIR) was recorded using a Nicolet 8700 FTIR spectrophotometer (Thermo
141
Fisher Scientific Inc., USA) equipped with a Universal ATR accessory. To prepare an
142
ATR sample, 0.1 g catalyst and 2 mL D2O were added to a 5 mL centrifuge tube and
143
sonicated for 5 min. Subsequently, the solid in the stock was separated by
144
centrifugation and treated again with another aliquot of D2O in order to eliminate
145
residual H2O. The sample was sealed and centrifuged immediately. Half of the
146
supernatant was used as the reference, and the solid resuspended in the other half was
147
used as the sample. The BPA solution (100 ppm) was also prepared by D2O and the
148
prepared process of the sample with BPA was the same as that of the pure D2O. The
149
chemicals and reagents, characterization and density functional theory (DFT)
150
calculations are presented in the SI.
151 152
RESULTS AND DISCUSSION
153
Characterization of catalysts. Figure 1a shows the powder X-ray diffraction (XRD)
154
patterns of the as-synthesized samples. The XRD peaks in the samples of CuAlO2 and
155
CuAlO2 (550 ◦C, 4 h) were indexed to the standard CuAlO2 (e.g., 006, 101, 012, 009
156
and 018 planes at 2θ = 31.6°, 36.5°, 37.8°, 48.2° and 56.9°, respectively; JCPDS No.
157
73-9485),28 indicating that the second calcination procedure did not affect the
158
structural properties of the catalyst without the addition of urea. However, in addition
159
to the typical CuAlO2 diffraction pattern, two evident peaks indexed to CuO (200) at
160
2θ = 35.8°29 and CuO (111) at 2θ = 39.1°30 were observed for CN-Cu(II)-CuAlO2,
161
indicating that the second calcination procedure significantly influenced the crystal
162
structure of catalyst in the presence of urea, which led to the formation of Cu(II).
163
Differently, although the surface growth process of CN-Cu(II)-CuAlO2 was prepared 7 ACS Paragon Plus Environment
Environmental Science & Technology
164
using the g-C3N4 synthesis process, there was no significant XRD peaks were indexed
165
to g-C3N4, suggesting that the produced surface species was not the standard g-C3N4
166
structure with the characteristic (100) and (002) planes due to the influence of the
167
surface Cu species of CuAlO2 during this surface growth process.
168
The field emission scanning electron microscopy (FESEM) images of Figure
169
S2a,b (SI) show that CuAlO2 consists of loosely packed grains with irregular shapes.
170
The size of the fine particles is less than 100 nm and that of coarse particles is about 2
171
µm. CN-Cu(II)-CuAlO2 (Figure S2c,d, SI) shows a larger particle size (0.4−3 µm)
172
and a smoother surface generally. This is due to the in-situ growth of urea and
173
covering the carbon nitride compounds on the CuAlO2 surface. Figure S3a (SI)
174
shows the transmission electron microscopy (TEM) of CN-Cu(II)-CuAlO2. As seen
175
from the micrographs, the CN-Cu(II)-CuAlO2 grain exhibits a plate like elongated
176
morphology, which may form due to anisotropic crystal growth observed in CuAlO2.
177
The HR-TEM image of CN-Cu(II)-CuAlO2 (Figure S3b, SI) exhibits clear lattice
178
fringes. The measured distances of 0.21 and 0.24 nm can be attributed to the dhkl of
179
(104) and (012) planes of the CuAlO2 substrate structure. The EDS linescan (Figure
180
1b) and TEM elemental mappings (Figure 1c) of CN-Cu(II)-CuAlO2 show that Cu,
181
Al and O were the main elements and uniformly distributed in the particle bulk
182
horizontally and vertically. C was mainly distributed on the surface and edge of the
183
particles. The top and bottom parts of C mapping were influenced by the C-containing
184
support grid for TEM measurements. Differently, the N signals (may affected by the
185
O element) were relatively weak on the whole scale. These results revealed the in-situ
186
growth and formation of CN on the catalyst surface during the synthesis process.
187
The X-ray photoelectron spectroscopy (XPS) spectra on Cu 2p3/2 of CuAlO2 and
188
CN-Cu(II)-CuAlO2 samples were exhibited in Figure 2a,b. For CuAlO2 (Figure 2a), 8 ACS Paragon Plus Environment
Page 8 of 33
Page 9 of 33
Environmental Science & Technology
189
the obvious binding energy (BE) peak at 932.1 eV was attributed by the reduction
190
state copper species.31 The auger parameters obtained at 1848.8 eV confirmed the
191
reduction state copper species belonging to Cu(I). This result suggested that Cu
192
species mainly exist in the form of cuprous state in CuAlO2. The weak peak located at
193
934.6 eV accompanied by the appearance of shake-up satellite lines were attributed by
194
the small amount of Cu(II)
195
revealed that both Cu(I) and Cu(II) existed on the surface of CuAlO2 with the Cu(I) to
196
Cu(II) atomic ratio of 1:0.61. After the surface growth process, the surface atomic
197
ratio of Cu(I) to Cu(II) was significantly changed (Figure 2b), which obtained with
198
1:1.79 for CN-Cu(II)-CuAlO2. The results indicated that the surface growth of the
199
carbon nitride compounds (CN) on the CuAlO2 substrate greatly increased the
200
proportion of Cu(II) on the surface of CN-Cu(II)-CuAlO2, suggesting the surface CN
201
compounds had a strong interaction with the Cu species during the synthesis process.
20
on the surface of CuAlO2. The XPS curve fittings
202
CuAlO2 showed mixed bands in the BE range of 72 to 80 eV (Figure S4a, SI),
203
the deconvolution of which indicated the co-presence of Al 2p (73.3 eV), Cu 3p3/2
204
(74.4) and Cu 3p1/2 (76.8 eV).32 However, Al 2p did not appear, and only Cu 3p3/2
205
(74.2) and Cu 3p1/2 (77.0 eV) were observed in the same BE range of
206
CN-Cu(II)-CuAlO2 (Figure S4b, SI). This result revealed that the surface growth of
207
CN obscured the Al species in the substrate.
208
In the C 1s XPS spectra of CN-Cu(II)-CuAlO2 (Figure 2c), the distinct bands at
209
284.7 and 288.7 eV corresponded to the sp2 C-C bonds and the sp2-hybridized C
210
(N-C=N) in the aromatic tri-s-triazine rings,33 respectively, which were very close to
211
the characteristic peaks of the pristine g-C3N4 (Figure S5a, SI). This result suggested
212
that the produced surface species on the CN-Cu(II)-CuAlO2 sample formed the
213
g-C3N4-like configuration although they were not the standard g-C3N4 due to the 9 ACS Paragon Plus Environment
Environmental Science & Technology
214
influence of the surface Cu species of CuAlO2 during this surface growth process.
215
Differently, the intensity proportion of the peak corresponding to C-C at 284.7 eV
216
markedly increased, and the intensity proportion corresponding to N-C=N at 288.7 eV
217
significantly decreased in CN-Cu(II)-CuAlO2, which revealed that the excess C atoms
218
were incorporated into the tri-s-triazine of g-C3N4, and then occupied and replaced the
219
N atoms, resulting in the great increase of C-C and decrease of N-C=N. This result
220
could also be confirmed by the atomic ratio of C to N. The value of
221
CN-Cu(II)-CuAlO2 was ~63.9, which was much higher than that of the pristine
222
g-C3N4 (~0.79).
223
The above results suggested that the CN compounds on the surface of
224
CN-Cu(II)-CuAlO2 was actually the g-C3N4 containing the excess C (i. e. C-g-C3N4).
225
In addition, a new peak at 286.5 eV for the C 1s spectra emerged in
226
CN-Cu(II)-CuAlO2 (Figure 2c), which is ascribed to the C atoms of the aromatic ring
227
bonding to the deprotonated hydroxyl groups (C-O-M).34, 35 In combination with the
228
Cu XPS analysis of CuAlO2 and CN-Cu(II)-CuAlO2, the binding energy in Cu 2p3/2,
229
Cu 3p3/2 and Cu 3p1/2 were all shifted obviously after the surface growth of CN
230
compounds, indicating that the CuAlO2 substrate had a strong interaction with
231
C-g-C3N4, and the C-O-M bond occurs at the Cu sites. Thus, the evidence above
232
revealed that the connection between the surface CN compounds and the CuAlO2
233
substrate was achieved by the C-O-Cu bridge.
234
The N 1s XPS of CN-Cu(II)-CuAlO2 was fitted into three main peaks at 398.8,
235
400.1 and 402.3 eV (Figure S5b, SI), corresponding to the C=N-C in triazine rings,
236
the tertiary N in the form of N-(C)3 and the amino functional groups (C-N-H),33
237
respectively. The relative intensity proportion of C=N-C in triazine rings was greatly
238
decreased in CN-Cu(II)-CuAlO2 compared with the pristine g-C3N4 (Figure S5c, SI), 10 ACS Paragon Plus Environment
Page 10 of 33
Page 11 of 33
Environmental Science & Technology
239
which further confirmed that the N atoms were substantially replaced by excess C
240
atoms during the surface growth process.
241
The O1s XPS of CuAlO2 (Figure S5d, SI) shows strong signals in the range of
242
528−534 eV, which can be fitted into 3 deconvolutions around 530.2 eV (lattice
243
O-Cu(I) of CuAlO2),36, 37 530.9 eV (lattice O-Al(III) of CuAlO2)36 and 532.2 eV
244
(surface –OH on CuAlO2).38-40 For the O1s XPS of CN-Cu(II)-CuAlO2 (Figure 2d),
245
the three peaks were still retained. The peak at 530.2 eV obviously weakened
246
compared with that of CuAlO2, which was due to the conversion of Cu (I) to Cu (II)
247
after the surface growth process according to the foregoing analysis. The emergence
248
of a new peak at 532.5 eV futher confirmed the produced Cu(II) connecting with the
249
deprotonated hydroxyl group through C-O-Cu bonding bridge,40,
250
consistent with the result of C 1s XPS. In addition, the oxygen vacancy concentration
251
could be reflected from the variation of area ratio R of the lattice O 1s peaks to the
252
surface O 1s peaks (Plattice/Psurface).41 The value of R(CN-Cu(II)-CuAlO2)=2.68 was
253
much higher than that of R(CuAlO2)=1.89, indicating that the surface oxygen vacancy
254
(Vo) concentration of CN-Cu(II)-CuAlO2 was significantly lower than that of CuAlO2.
255
Formation of active dual reaction centers. The EPR technique was used to monitor
256
the electronic band structure information and investigate the presence of the single
257
electrons in the prepared samples. As shown in Figure 3a, g-C3N4 showed a sharp
258
EPR signal at g=1.996, which originated from the uncoupled electron of the skeleton
259
CN aromatic rings.42, 43 CuAlO2 showed no detectable EPR signals, indicating no
260
single electrons existing in CuAlO2. However, the as-prepared CN-Cu(II)-CuAlO2
261
displayed a very strong EPR signal at g=2.084, which was significantly enhanced
262
compared with the pure g-C3N4, revealing the conspicuous increase of the electron
263
density around Cu due to the Cu(II)-π interactions of the formed C-O-Cu bonding 11 ACS Paragon Plus Environment
41
which was
Environmental Science & Technology
264
bridge on CN-Cu(II)-CuAlO2.44 These results indicated that the C-O-Cu produced by
265
the Cu(II) connecting with the deprotonated hydroxyl group on C-g-C3N4 gathered a
266
large amount of single electrons around the Cu, forming the electron-rich reaction
267
centers.
268
The distributions of the valence-electron density can provide useful information
269
for the reaction center analysis. The negative and positive areas are expected to be the
270
promising reactive sites for the reduction and oxidation reaction,45 respectively.
271
Therefore, we investigated the distribution of the valence-electron density on
272
CN-Cu(II)-CuAlO2 through DFT calculations. Figure 3b,c shows the optimized
273
geometry structure (left) and the corresponding two-dimensional valence-electron
274
density color-filled maps (right) of the CN-Cu(II)-CuAlO2 model fragments. The
275
construction and optimization of the geometry structure model were achieved by
276
using the B3LYP functional and Gaussian 09 package. The dangling bonds were
277
terminated with H atoms to obtain a neutral cluster. The valence-electron density was
278
analyzed using the Multiwfn package. As shown in Figure 3b, the largest electron
279
distribution area appears around the Cu atom, and its maximum valence-electron
280
density is as high as 4.0 e/Å3 compared with the O, C and N atoms in the
281
CN-Cu(II)-CuAlO2 model fragment, which theoretically confirms that the
282
electron-rich center is formed around the Cu atom, being consistent with the
283
experimental results. The valence-electron densities around C and N on the aromatic
284
rings of CN are remarkably lower than those around Cu and O atoms, revealing that
285
the π electrons on the surface of CN-Cu(II)-CuAlO2 shift to the Cu center (π→Cu, σ
286
donation) through the special Cu-O-C bonding bridge. Thus, the electron-poor centers
287
must be formed on the aromatic rings of CN. To accurately determine the electron
288
poor sites, the valence-electron density map of the CN fragments on the 12 ACS Paragon Plus Environment
Page 12 of 33
Page 13 of 33
Environmental Science & Technology
289
CN-Cu(II)-CuAlO2 model was magnified and presented on Figure 3c. Obviously, the
290
narrowest electron distribution area appears around the C atoms, and its maximum
291
valence-electron density is only ~0.2 e/Å3, which is much smaller than that of the N
292
atoms, indicating that the electron-poor sites is around the C atoms. In the aromatic
293
rings of CN, the π electrons tends to be concentrated around the N atoms, which
294
suggests that in addition to the π→Cu (σ donation) electron transfer process, C
295
replacing N in the CN-rings can also promote the formation of the electron-poor C
296
centers.
297
The highest occupied molecular orbital (HOMO) and lowest unoccupied
298
molecular orbital (LUMO) distribution information of the dual reaction center
299
fragment of CN-Cu(II)-CuAlO2 was obtained via DFT calculations (Figure S6, SI).
300
The energy separation of the HOMO and LUMO is often used as an indicator of
301
chemical reactivity.46 The HOMO–LUMO gap for pure g-C3N4 was calculated as 5.17
302
(α electron gap = β electron gap), representing a high kinetic stability and a low
303
chemical reactivity. However, the gaps of the α and β electrons for the dual reaction
304
center fragment of CN-Cu(II)-CuAlO2 evidently reduce to 3.62 eV and 2.24 eV,
305
respectively, which is energetically favorable to extract the electrons from a low-lying
306
HOMO and to add the electrons to a high-lying LUMO, forming an activated complex
307
of the potential reaction.47 This result theoretically demonstrates that the chemical
308
reactivity of CN-Cu(II)-CuAlO2 is significantly enhanced by the formation of the dual
309
reaction centers (electron-rich and electron-poor centers).
310
BMPO-trapped EPR spectra were used to detect •OH and HO2•/O2•− radicals in
311
the CN-Cu(II)-CuAlO2 air-saturated aqueous and methanol-aqueous dispersions
312
without adding H2O2, respectively. As shown in Figure S7a (SI), six distinct and
313
sharp
characteristic
peaks
of
BMPO-HO2•/O2•− 13
ACS Paragon Plus Environment
were
observed
in
the
Environmental Science & Technology
methanol-aqueous
dispersion,
which
Page 14 of 33
314
CN-Cu(II)-CuAlO2
315
electron-rich reaction centers on CN-Cu(II)-CuAlO2 with copious single electrons
316
could efficiently reduce the dissolved O2 to HO2•/O2•−. By using the method of
317
detecting •OH radicals in aqueous dispersion, a series of distinct EPR signals were
318
also detected in the CN-Cu(II)-CuAlO2 suspension without adding H2O2 (Figure S7b,
319
SI). Although the EPR signals are complex due to the effect of the unpaired electrons
320
on CN-Cu(II)-CuAlO2, they also can split into four single lines with a spacing of ~14
321
G in the magnetic field due to the hyperfine interaction between the electron spin
322
of •OH and the orbital spin of N atom in BMPO. This result suggested that
323
CN-Cu(II)-CuAlO2 donated the electrons to O2 in the electron-rich reaction centers
324
and simultaneously accepted the electrons from H2O to produce •OH in the
325
electron-poor reaction centers, forming an electron transfer cycle. This effect was
326
predominantly due to the orbital interactions involving the electron transfer of π→Cu
327
(σ donation) and Cu→π* (π back-donation). All of the above evidence shows that the
328
activated dual reaction centers have been successfully built on CN-Cu(II)-CuAlO2.
329
Efficient degradation of pollutants over CN-Cu(II)-CuAlO2/H2O2 suspension.
330
The catalytic activities of the prepared samples were initially evaluated through the
331
degradation of BPA in the presence of H2O2 under the mild conditions. As shown in
332
Figure 4a, no significant BPA degradation was observed in the g-C3N4/H2O2
333
suspension. The degradation of BPA in the CuAlO2/H2O2 suspension was only 15.8%
334
within 120 min. Astonishingly, in the CN-Cu(II)-CuAlO2/H2O2 suspension, the
335
degradation degree of BPA could reach 95.5% within 120 min under the neutral and
336
mild conditions, which was 25 and 45 times higher than that in the suspensions of
337
CuAlO2/H2O2 and g-C3N4/H2O2, respectively. In addition, the TOC removal rate
338
reached 41.5% in CN-Cu(II)-CuAlO2/H2O2 suspension within 180 min (Figure 4b), 14 ACS Paragon Plus Environment
indicated
that
the
Page 15 of 33
Environmental Science & Technology
339
which was greatly higher than that in the suspensions of g-C3N4 or CuAlO2 (almost no
340
any removal of TOC).
341
To further investigate the activity and adaptability of CN-Cu(II)-CuAlO2 for
342
different organic pollutants, four other refractory organic pollutants including phenol,
343
2-chlorphenol, ibuprofen and phenytoin were used to be degraded in the
344
CN-Cu(II)-CuAlO2 suspensions under natural conditions. As shown in Figure 4c, all
345
of the refractory compounds were substantially degraded and the degradation rate
346
could reached 88.3%-97.5% within 120 min. In addition, the degradation of higher
347
initial concentrations for BPA and 2-chlorphenol by CN-Cu(II)-CuAlO2 were also
348
carried out as shown in Figure S8 (SI). The degradation rates of both BPA (100 mg
349
L-1) and 2-CP (100 mg L-1) could reach ~90% within 120 min.
350
During the Fenton-like reactions, the maximum concentration of dissolved Cu
351
was ~0.8 mg L-1 (including the surface free Cu ions), which was much lower than the
352
limitations of the EU directives (< 2.0 mg L-1) and USA regulations (< 1.3 mg L-1).
353
No Al ions was detected in the CN-Cu(II)-CuAlO2 suspension. In addition, only 3.3%
354
TOC was removed for BPA degradation by the homogeneous Fenton reaction (Cu2+
355
concentration 0.8 mg L−1) within 120 min, which was much lower than that in the
356
CN-Cu(II)-CuAlO2 suspension, suggesting that the contribution for the pullutant
357
degradation of the released ions was negligible. The durability of CN-Cu(II)-CuAlO2
358
was tested by recovering the solid catalyst through filtration, washing and drying. As
359
shown in Figure S9 (SI), the degradation of BPA in the CN-Cu(II)-CuAlO2
360
suspension was not significantly decreased within the margin of tolerance even after 6
361
successive cycles of degradation testing. Cu2+ leaching from the catalyst significantly
362
decreased over the reuse cycles. These results suggested that CN-Cu(II)-CuAlO2 was
363
an efficient Fenton catalyst with a good repeatability and stability. 15 ACS Paragon Plus Environment
Environmental Science & Technology
364
The pH is often regarded as a sensitive factor for Fenton reaction, so the effect of
365
the initial pH value on BPA degradation in CN-Cu(II)-CuAlO2 suspension was
366
determined as presented in Figure 4d. Under the neutral conditions, the catalyst
367
showed the highest activity. In the range of pH 7-9, the degradation of BPA slightly
368
decreased with the increase of the initial pH, but it still could reach 70% within 120
369
min at pH 9. In the range of pH 5-7, the degradation rate of BPA was almost
370
unchanged with the change of the initial solution pH, which indicated that the activity
371
of the catalyst was not obviously affected by the pH values. In aqueous solution, the
372
metal-containing catalyst often forms surface hydroxyl groups by dissociative
373
chemisorptions of water molecules,48 leading to the activity of the metal-containing
374
catalyst strongly dependent on the solution pH, while CN-Cu(II)-CuAlO2 avoid this
375
effect due to the formation of the dual reaction centers. After the reaction, the final pH
376
values of the solutions (initial pH 6, 7, 8 and 9) generally declined (Figure S10, SI)
377
due to the production of the acidic intermediates, such as some low-molecular weight
378
organic acids, from the BPA degradation. For the solution of initial pH 5, the final pH
379
value did not change, indicating that the acidic intermediates had been degraded under
380
this condition. We also carried out the activity evaluation of CN-Cu(II)-CuAlO2 at
381
room temperature (Figure S11, SI). Within 120 min, the BPA degradation rates in the
382
CN-Cu(II)-CuAlO2/H2O2 suspensions at room temperature (~25°C), 30°C and 35°C
383
were ~90%, ~93% and ~96%, respectively. The results suggested that the activity of
384
the catalyst for BPA degradation did not change obviously at the three different
385
temperatures and the reactions were not highly temperature-dependent.
386
Fenton-like reaction mechanism on active dual reaction centers. The H2O2
387
decomposition experiment was performed in the presence of the catalysts and BPA
388
(Figure 5a). The decomposition rate of H2O2 in the CN-Cu(II)-CuAlO2 system was 16 ACS Paragon Plus Environment
Page 16 of 33
Page 17 of 33
Environmental Science & Technology
389
significantly slower than that in the CuAlO2 system, suggesting that CuAlO2 and
390
CN-Cu(II)-CuAlO2 exhibited the opposite reactivity for the decomposition of H2O2
391
compared to the degradation of BPA. CN-Cu(II)-CuAlO2 exhibited the highest
392
catalytic activity toward to the BPA degradation but consumed the least amount of
393
H2O2, revealing a high utilization efficiency of H2O2 in CN-Cu(II)-CuAlO2 system,
394
which indicated that the formation of the dual reaction centers were extremely
395
selective for promoting the pollutant degradation with minor spurious decomposition
396
of H2O2 to O2. Generally, oxygen vacancies often take part in the fast decomposition
397
of H2O2 into O2.19, 22 According to the previous XPS results, the oxygen vacancy
398
concentration of CuAlO2 was significantly more than that of CN-Cu(II)-CuAlO2.
399
Thus H2O2 was rapidly and invalidly decomposed in the CuAlO2 system with very
400
little pollutant degradation.
401
Figure 5b shows the in-situ ATR-FTIR spectra for the CuAlO2 and
402
CN-Cu(II)-CuAlO2 suspensions. In order to clearly distinguish the hydroxyl groups
403
from different sources, we carried out the experiments with D2O instead of H2O. The
404
bands in the range of 2200-2700 cm−1 are assigned to the -OD stretching of D2O ν(-OD)
405
on the catalyst surface, and the band at ~1190 cm−1 is assigned to the molecular
406
bending modes of D2O δ(D-O-D).49 For CuAlO2/D2O suspension, the ν(-OD) bands
407
centered at 2302 and 2560 cm-1. However, the bands toward adsorbing BPA on
408
CuAlO2 surface were shifted to low frequency (2550 and 2290 cm-1), which was due
409
to the deprotonation of the phenolic OH group of BPA complexing with the surface
410
Cu via σ bonding to the lone pairs of the oxygen atom.12 For CN-Cu(II)-CuAlO2/D2O,
411
the intensities of all the peaks corresponding to D2O greatly weakened, and the ν(-OD)
412
shifted to lower frequency (2285 cm-1) compared to the CuAlO2 suspension (2302
413
cm-1). The results indicated that the CN compounds had complexed with the surface 17 ACS Paragon Plus Environment
Environmental Science & Technology
414
Cu via σ bonding on CN-Cu(II)-CuAlO2, forming the dual reaction centers, which
415
greatly impedes the direct adsorption of D2O on the Cu sites. In the presence of BPA,
416
the ν(-OD) band did not continuously shift to low frequency but to a higher frequency
417
(2304 cm-1), which is the evidence that the pollutants do not directly adsorb to the
418
electron-rich Cu centers, but are mainly degraded in the electron-poor centers during
419
the Fenton-like reaction.
420
The conversion paths of H2O2 on different surfaces of the catalysts during the
421
Fenton-like reaction process were studied by the EPR spin-trap technique with BMPO.
422
As shown in Figure 5c, after adding H2O2, very weak BMPO-•OH signals were
423
detected in the CuAlO2/H2O2 system, suggesting the low conversion rate of H2O2 to
424
•
425
very strong BMPO-•OH signals, which was ~4 times higher than that in the
426
CuAlO2/H2O2 system. This result indicated that H2O2 were efficiently reduced to •OH
427
by the free electrons on the electron-rich centers of CN-Cu(II)-CuAlO2. Similarly, no
428
significant BMPO-HO2•/O2•− signals were observed in the CuAlO2/H2O2 system
429
(Figure 5d) due to the direct oxidation of H2O2 by the surface oxygen vacancies on
430
CuAlO2, whereas four distinct characteristic signals of BMPO-HO2•/O2•− were
431
observed in the CN-Cu(II)-CuAlO2/H2O2 system. The strong BMPO-HO2•/O2•−
432
signals might originate from three reactions: the reduction of the dissolved O2 on the
433
electron-rich reaction centers, the oxidation of H2O2 on the electron-poor reaction
434
centers and the activation of H2O2 in the small amount of surface copper ions. These
435
phenomena suggested that the creation of the dual reaction centers could efficiently
436
and rapidly activate H2O2 and produce the reactive free radicals, which was extremely
437
selective for promoting the pollutant degradation with minor spurious decomposition
438
of H2O2 to O2. However, CuAlO2 with a great amount of oxygen vacancies only
OH on the surface of CuAlO2. In contrast, CN-Cu(II)-CuAlO2/H2O2 system exhibited
18 ACS Paragon Plus Environment
Page 18 of 33
Page 19 of 33
Environmental Science & Technology
439
promoted the rapid and inefficient decomposition of H2O2 and did not result in the
440
generation of the reactive species in the solutions. This is the reason that
441
CN-Cu(II)-CuAlO2 exhibited the highest catalytic activity toward the pollutant
442
degradation but consumed the least amount of H2O2, whereas CuAlO2 exhibited low
443
activity for pollutant degradation but invalidly consumed a large amount of H2O2
444
during the Fenton-like reactions. These results confirmed that the formation of the
445
dual reaction centers on CN-Cu(II)-CuAlO2 was responsible for the high utilization
446
efficiency of H2O2 and the high Fenton-like reactivity. Moreover, the •OH signal
447
intensities for the CN-Cu(II)-CuAlO2/H2O2 system are not changed after adding BPA
448
(Figure 5c), indicating that the presence of the organic pollutants does not affect the
449
reduction of H2O2. However, the HO2•/O2•− signals are evidently weakened in the
450
CN-Cu(II)-CuAlO2/H2O2 system after adding BPA (Figure 5d), which is because the
451
electron-poor C sites tend to be complexed with the electron-rich organic compounds.
452
Thus, the organic pollutants can also act as electron donors at the electron-poor sites,
453
avoiding the oxidation of partial H2O2 into HO2•/O2•−.
454
The remained single electrons of CN-Cu(II)-CuAlO2 after reacting with the
455
dissolved O2/H2O and H2O2 were measured through EPR. As shown in Figure S12
456
(SI), the EPR signal intensity almost did not change after reaction with the dissolved
457
O2/H2O and H2O2. This result further confirmed the dual center reaction process and
458
revealed the electron compensation effect on the dual reaction centers that O2 or H2O2
459
captured electrons from the electron-rich centers, and simultaneously H2O or H2O2
460
donated electrons to the electron-poor centers. These donated electrons of H2O or
461
H2O2 were quickly diverted to the electron-rich centers from the electron-poor centers
462
due to the special connecting mode of Cu-O-C by the produced Cu(II) connecting
463
with the deprotonated hydroxyl group on C-g-C3N4. The EPR signal that did not 19 ACS Paragon Plus Environment
Environmental Science & Technology
464
weaken after reaction revealed the excellent electron cycle capability of the dual
465
reaction centers on CN-Cu(II)-CuAlO2.
466 467
ASSOCIATED CONTENT
468
Supporting Information
469
Supplementary Methods and Figures S1-S12. This material is available free of charge
470
via the Internet at http://pubs.acs.org.
471
AUTHOR INFORMATION
472
Corresponding Author
473
*
[email protected] /
[email protected] (Hu C.)
474
Notes
475
The authors declare no competing financial interest.
476
ACKNOWLEDGMENTS
477
This work was supported by the National Key Research and Development Plan
478
(2016YFA0203200), the National Natural Science Foundation of China (51538013)
479
and the Science Starting Foundation of Guangzhou University (27000503151 and
480
2700050302).
481 482
Reference
483
(1) Liu, Y. Y.; Liu, X. M.; Zhao, Y. P.; Dionysiou, D. D., Aligned α-FeOOH nanorods
484
anchored on a graphene oxide-carbon nanotubes aerogel can serve as an effective
485
Fenton-like oxidation catalyst. Appl. Catal., B 2017, 213, 74-86.
486
(2) Zhang, L. L.; Tu, J. J.; Lyu, L.; Hu, C., Enhanced catalytic degradation of
487
ciprofloxacin over Ce-doped OMS-2 microspheres. Appl. Catal., B 2016, 181,
488
561-569. 20 ACS Paragon Plus Environment
Page 20 of 33
Page 21 of 33
Environmental Science & Technology
489
(3) Shi, J. G.; Ai, Z. H.; Zhang, L. Z., Fe@Fe2O3 core-shell nanowires enhanced
490
Fenton oxidation by accelerating the Fe(III)/Fe(II) cycles. Water Res. 2014, 59,
491
145-153.
492
(4) Cheng, M.; Zeng, G. M.; Huang, D. L.; Lai, C.; Liu, Y.; Xu, P.; Zhang, C.; Wan, J.;
493
Hu, L.; Xiong, W. P.; Zhou, C. Y., Salicylic acid-methanol modified steel converter
494
slag as heterogeneous Fenton-like catalyst for enhanced degradation of alachlor.
495
Chem. Eng. J. 2017, 327, 686-693.
496
(5) Ma, J. Q.; Yang, Q. F.; Wen, Y. Z.; Liu, W. P., Fe-g-C3N4/graphitized mesoporous
497
carbon composite as an effective Fenton-like catalyst in a wide pH range. Appl. Catal.,
498
B 2017, 201, 232-240.
499
(6) Huang, X. P.; Hou, X. J.; Jia, F. L.; Song, F. H.; Zhao, J. C.; Zhang, L. Z.,
500
Ascorbate-promoted surface iron cycle for efficient heterogeneous Fenton alachlor
501
degradation with hematite nanocrystals. ACS Appl. Mater. Inter. 2017, 9, (10),
502
8751-8758.
503
(7) Qin, Y. X.; Zhang, L. Z.; An, T. C., Hydrothermal carbon-mediated Fenton-like
504
reaction mechanism in the degradation of alachlor: Direct electron transfer from
505
hydrothermal carbon to Fe(III). ACS Appl. Mater. Inter. 2017, 9, (20), 17116-17125.
506
(8) Navalon, S.; Martin, R.; Alvaro, M.; Garcia, H., Gold on diamond nanoparticles as
507
a highly efficient Fenton catalyst. Angew. Chem. Int. Edit. 2010, 49, (45), 8403-8407.
508
(9) Lyu, L.; Yu, G. F.; Zhang, L. L.; Hu, C.; Sun, Y., 4-Phenoxyphenol-functionalized
509
reduced graphene oxide nanosheets: A metal-free Fenton-like catalyst for pollutant
510
destruction. Environ. Sci. Technol. 2018, 52, (2), 747-756.
511
(10) Li, H.; Shang, J.; Yang, Z. P.; Shen, W. J.; Ai, Z. H.; Zhang, L. Z., Oxygen
512
vacancy associated surface Fenton chemistry: Surface structure dependent hydroxyl
513
radicals generation and substrate dependent reactivity. Environ. Sci. Technol. 2017, 51, 21 ACS Paragon Plus Environment
Environmental Science & Technology
514
(10), 5685-5694.
515
(11) Huang, X. P.; Hou, X. J.; Zhao, J. C.; Zhang, L. Z., Hematite facet confined
516
ferrous ions as high efficient Fenton catalysts to degrade organic contaminants by
517
lowering H2O2 decomposition energetic span. Appl. Catal., B 2016, 181, 127-137.
518
(12) Lyu, L.; Zhang, L. L.; Wang, Q. Y.; Nie, Y. L.; Hu, C., Enhanced Fenton catalytic
519
efficiency of γ-Cu-Al2O3 by σ-Cu2+-ligand complexes from aromatic pollutant
520
degradation. Environ. Sci. Technol. 2015, 49, (14), 8639-8647.
521
(13) Xiong, X. M.; Sun, Y. K.; Sun, B.; Song, W. H.; Sun, J. Y.; Gao, N. Y.; Qiao, J. L.;
522
Guan, X. H., Enhancement of the advanced Fenton process by weak magnetic field
523
for the degradation of 4-nitrophenol. RSC Adv. 2015, 5, (18), 13357-13365.
524
(14) Lyu, L.; Zhang, L. L.; Hu, C.; Yang, M., Enhanced Fenton-catalytic efficiency by
525
highly accessible active sites on dandelion-like copper-aluminum-silica nanospheres
526
for water purification. J. Mater. Chem. A 2016, 4, (22), 8610-8619.
527
(15) Li, X. F.; Liu, X.; Xu, L. L.; Wen, Y. Z.; Ma, J. Q.; Wu, Z. C., Highly dispersed
528
Pd/PdO/Fe2O3 nanoparticles in SBA-15 for Fenton-like processes: confinement and
529
synergistic effects. Appl. Catal., B 2015, 165, 79-86.
530
(16) Lin, K. Y. A.; Lin, J. T., Ferrocene-functionalized graphitic carbon nitride as an
531
enhanced heterogeneous catalyst of Fenton reaction for degradation of Rhodamine B
532
under visible light irradiation. Chemosphere 2017, 182, 54-64.
533
(17) Zhang, L. L.; Lyu, L.; Nie, Y. L.; Hu, C., Cu-doped Bi2O3/Bi0 composite as an
534
efficient Fenton-like catalyst for degradation of 2-chlorophenol. Sep. Purif. Technol.
535
2016, 157, 203-208.
536
(18) Wang, H. H.; Zhang, L. L.; Hu, C.; Wang, X. K.; Lyu, L.; Sheng, G. D.,
537
Enhanced degradation of organic pollutants over Cu-doped LaAlO3 perovskite
538
through heterogeneous Fenton-like reactions. Chem. Eng. J. 2018, 332, 572-581. 22 ACS Paragon Plus Environment
Page 22 of 33
Page 23 of 33
Environmental Science & Technology
539
(19) Lyu, L.; Zhang, L. L.; Hu, C., Enhanced Fenton-like degradation of
540
pharmaceuticals over framework copper species in copper-doped mesoporous silica
541
microspheres. Chem. Eng. J. 2015, 274, 298-306.
542
(20) Lyu, L.; Zhang, L. L.; Hu, C., Galvanic-like cells produced by negative charge
543
nonuniformity of lattice oxygen on d-TiCuAl-SiO2 nanospheres for enhancement of
544
Fenton-catalytic efficiency. Environ. Sci.: Nano 2016, 3, (6), 1483-1492.
545
(21) Yu, G. F.; Lyu, L.; Zhang, F. G.; Yan, D. B.; Cao, W. R.; Hu, C., Theoretical and
546
experimental evidence for rGO-4-PP Nc as a metal-free Fenton-like catalyst by tuning
547
the electron distribution. RSC Adv. 2018, 8, (6), 3312-3320.
548
(22) Lee, Y. N.; Lago, R. M.; Fierro, J. L. G.; Gonzalez, J., Hydrogen peroxide
549
decomposition over Ln(1-x)A(x)MnO(3) (Ln = La or Nd and A = K or Sr) perovskites.
550
Appl. Catal., A 2001, 215, (1-2), 245-256.
551
(23) Nie, Y. L.; Hu, C.; Qu, J. H.; Zhao, X., Photoassisted degradation of endocrine
552
disruptors over CuOx-FeOOH with H2O2 at neutral pH. Appl. Catal., B 2009, 87, (1-2),
553
30-36.
554
(24) Lyu, L.; Hu, C., Heterogeneous Fenton catalytic water treatment technology and
555
mechanism. Prog. Chem. 2017, 29, (9), 981-999.
556
(25) Lagutschenkov, A.; Sinha, R. K.; Maitre, P.; Dopfer, O., Structure and infrared
557
spectrum of the Ag+-phenol ionic complex. J. Phys. Chem. A 2010, 114, (42),
558
11053-11059.
559
(26) Lyu, L.; Zhang, L. L.; He, G. Z.; He, H.; Hu, C., Selective H2O2 conversion to
560
hydroxyl radicals in the electron-rich area of hydroxylated C-g-C3N4/ CuCo-Al2O3. J.
561
Mater. Chem. A 2017, 5, (15), 7153-7164.
562
(27) Bader, H.; Sturzenegger, V.; Hoigne, J., Photometric-method for the
563
determination of low concentrations of hydrogen-peroxide by the peroxidase 23 ACS Paragon Plus Environment
Environmental Science & Technology
564
catalyzed oxidation of N,N-Diethyl-P-Phenylenediamine (DPD). Water Res. 1988, 22,
565
(9), 1109-1115.
566
(28) Choi, S. Y.; Kim, C. D.; Han, D. S.; Park, H., Facilitating hole transfer on
567
electrochemically synthesized p-type CuAlO2 films for efficient solar hydrogen
568
production from water. J. Mater. Chem. A 2017, 5, (21), 10165-10172.
569
(29) Ehara, T.; Abe, H.; Iizaka, R.; Abe, K.; Sato, T., Crystalline orientation control in
570
sol-gel preparation of CuAlO2 thin films. J. Sol-Gel Sci. Techn. 2017, 82, (2),
571
363-369.
572
(30) Mosleh, S.; Rahimi, M. R.; Ghaedi, M.; Dashtian, K.; Hajati, S.,
573
Sonochemical-assisted synthesis of CuO/Cu2O/Cu nanoparticles as efficient
574
photocatalyst for simultaneous degradation of pollutant dyes in rotating packed bed
575
reactor: LED illumination and central composite design optimization. Ultrason.
576
Sonochem. 2018, 40, 601-610.
577
(31) Lopez-Suarez, F. E.; Parres-Esclapez, S.; Bueno-Lopez, A.; Illan-Gomez, M. J.;
578
Ura, B.; Trawczynski, J., Role of surface and lattice copper species in
579
copper-containing (Mg/Sr)TiO3 perovskite catalysts for soot combustion. Appl. Catal.,
580
B 2009, 93, (1-2), 82-89.
581
(32) Zhang, Y. J.; Liu, Z. T.; Zang, D. Y.; Feng, L. P., Structural and opto-electrical
582
properties of Cu-Al-O thin films prepared by magnetron sputtering method. Vacuum
583
2014, 99, 160-165.
584
(33) Zhang, J. S.; Zhang, M. W.; Yang, C.; Wang, X. C., Nanospherical carbon nitride
585
frameworks with sharp edges accelerating charge collection and separation at a soft
586
photocatalytic interface. Adv. Mater. 2014, 26, (24), 4121-4126.
587
(34) Tong, Z. W.; Yang, D.; Shi, J. F.; Nan, Y. H.; Sun, Y. Y.; Jiang, Z. Y.,
588
Three-dimensional porous aerogel constructed by g-C3N4 and graphene oxide 24 ACS Paragon Plus Environment
Page 24 of 33
Page 25 of 33
Environmental Science & Technology
589
nanosheets with excellent visible-light photocatalytic performance. ACS Appl. Mater.
590
Inter. 2015, 7, (46), 25693-25701.
591
(35) Li, H. Y.; Gan, S. Y.; Wang, H. Y.; Han, D. X.; Niu, L., Intercorrelated
592
superhybrid of AgBr supported on graphitic-C3N4-decorated nitrogen-doped graphene:
593
High engineering photocatalytic activities for water purification and CO2 reduction.
594
Adv. Mater. 2015, 27, (43), 6906-6907.
595
(36) Igbari, O.; Xie, Y. M.; Jin, Z. M.; Liao, L. S., Microstructural and electrical
596
properties of CuAlO2 ceramic prepared by a novel solvent-free ester elimination
597
process. J. Alloy Compd. 2015, 653, 219-227.
598
(37) Ji, H. H.; Lyu, L.; Zhang, L. L.; An, X. Q.; Hu, C., Oxygen vacancy enhanced
599
photostability and activity of plasmon-Ag composites in the visible to near-infrared
600
region for water purification. Appl. Catal., B 2016, 199, 230-240.
601
(38) Liu, Z. Y.; Duchon, T.; Wang, H. R.; Grinter, D. C.; Waluyo, I.; Zhou, J.; Liu, Q.;
602
Jeong, B.; Crumlin, E. J.; Matolin, V.; Stacchiola, D. J.; Rodriguez, J. A.; Senanayake,
603
S. D., Ambient pressure XPS and IRRAS investigation of ethanol steam reforming on
604
Ni-CeO2(111) catalysts: an in situ study of C-C and O-H bond scission. Phys. Chem.
605
Chem. Phys. 2016, 18, (25), 16621-16628.
606
(39) Maticiuc, N.; Katerski, A.; Danilson, M.; Krunks, M.; Hiie, J., XPS study of OH
607
impurity in solution processed CdS thin films. Sol. Energ. Mat. Sol. C 2017, 160,
608
211-216.
609
(40) Chen, L. J.; Chen, F.; Shi, Y. F.; Zhang, J. L., Preparation and visible light
610
photocatalytic activity of a graphite-like carbonaceous surface modified TiO2
611
photocatalyst. J Phys Chem C 2012, 116, (15), 8579-8586.
612
(41) Cao, W.; Tan, O. K.; Pan, J. S.; Zhu, W.; Reddy, C. V. G., XPS characterization of
613
x α-Fe2O3-(1-x)ZrO2 for oxygen gas sensing application. Mater. Chem. Phys. 2002, 75, 25 ACS Paragon Plus Environment
Environmental Science & Technology
614
(1-3), 67-70.
615
(42) Yu, X.; Li, Z. H.; Liu, J. W.; Hu, P. G., Ta-O-C chemical bond enhancing charge
616
separation between Ta4+ doped Ta2O5 quantum dots and cotton-like g-C3N4. Appl.
617
Catal., B 2017, 205, 271-280.
618
(43) Zhang, J. S.; Zhang, M. W.; Sun, R. Q.; Wang, X. C., A facile band alignment of
619
polymeric carbon nitride semiconductors to construct isotype heterojunctions. Angew.
620
Chem. Int. Edit. 2012, 51, (40), 10145-10149.
621
(44) Boukhvalov, D. W.; Osipov, V. Y.; Shames, A. I.; Takai, K.; Hayashi, T.; Enoki,
622
T., Charge transfer and weak bonding between molecular oxygen and graphene zigzag
623
edges at low temperatures. Carbon 2016, 107, 800-810.
624
(45) Chu, S.; Wang, Y.; Guo, Y.; Feng, J. Y.; Wang, C. C.; Luo, W. J.; Fan, X. X.; Zou,
625
Z. G., Band structure engineering of carbon nitride: In search of a polymer
626
photocatalyst with high photooxidation property. ACS Catal. 2013, 3, (5), 912-919..
627
(46) Zhang, L. P.; Xia, Z. H., Mechanisms of oxygen reduction reaction on
628
nitrogen-doped graphene for fuel cells. J. Phys. Chem. C 2011, 115, (22),
629
11170-11176.
630
(47) Aihara, J., Reduced HOMO-LUMO gap as an index of kinetic stability for
631
polycyclic aromatic hydrocarbons. J. Phys. Chem. A 1999, 103, (37), 7487-7495.
632
(48) Tamura, H.; Mita, K.; Tanaka, A.; Ito, M., Mechanism of hydroxylation of metal
633
oxide surfaces. J. Colloid. Interf. Sci. 2001, 243, (1), 202-207.
634
(49) Belhadj, H.; Melchers, S.; Robertson, P. K. J.; Bahnemann, D. W., Pathways of
635
the photocatalytic reaction of acetate in H2O and D2O: A combined EPR and
636
ATR-FTIR study. J. Catal. 2016, 344, 831-840.
637 638
26 ACS Paragon Plus Environment
Page 26 of 33
Page 27 of 33
Environmental Science & Technology
639
Figure Captions
640
Figure 1. (a) XRD patterns of the prepared samples. (b) EDS linescan (from TEM)
641
and (c) elemental mappings (from TEM) of CN-Cu(II)-CuAlO2.
642
Figure 2. XPS spectra in Cu 2p for (a) CuAlO2 and (b) CN-Cu(II)-CuAlO2, (c) C 1s
643
and (d) O 1s for CN-Cu(II)-CuAlO2.
644
Figure 3. (a) EPR spectra of the fresh g-C3N4, CuAlO2 and CN-Cu(II)-CuAlO2
645
samples. (b) and (c) DFT calculations for the optimized structure (left) and the
646
corresponding two-dimensional valence-electron density color-filled maps (right) of
647
the CN-Cu(II)-CuAlO2 model in -Cu(II)-CN vision fragment (b) and -CN vision
648
fragment (c). Orange, red, gray, blue and white circles denote Cu, O, C, N, and H
649
atoms, respectively. The valence-electron density is given in e/Å3.
650
Figure 4. (a) BPA degradation curves in various suspensions with H2O2 (Insert shows
651
the corresponding kinetic curves). (b) TOC removal curves during BPA degradation.
652
(c) Decomposition curves of different pollutants in CN-Cu(II)-CuAlO2 suspensions
653
with H2O2. (d) Effect of initial pH values for BPA degradation. Reaction conditions:
654
Natural initial pH (pH 6-7, except d), initial pollutants 25 mg/L, initial H2O2 10 mM,
655
catalyst 1.0 g L−1.
656
Figure 5. (a) H2O2 decomposition curves during BPA degradation in CuAlO2 and
657
CN-Cu(II)-CuAlO2 suspensions. Reaction conditions: Natural initial pH, initial
658
pollutants 25 mg/L, initial H2O2 10 mM, catalyst 1.0 g L−1. (b) ATR-FTIR spectra of
659
CuAlO2 and CN-Cu(II)-CuAlO2 in D2O environment with/without BPA. BMPO
660
spin-trapping EPR spectra for (c) •OH and (d) HO2•/O2•− in various suspensions in the
661
presence of H2O2 with/without pollutants.
662
27 ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 33
663 ∗ g-C3N4
018
♥
♥ ♥ ♥
10
♥
♥
1200
♥
♥ ♥
♥
♥
♥
♥
♥
♥
CuAlO2 (550°C,4h)
♥
♥
♥
♥
C N O Al Cu
CPS
♥
CN-Cu(II)-CuAlO2
♥
110 1010
♥
009
♥
107
006
002 101 111 012 104
♥
♦ ♦
CuAlO2 ∗100 g-C3N4
664
1800
(b)
♥
♥ CuAlO 2 ♦ CuO
003
Intensity (a.u.)
(a)
600
♥ ♥
♥
∗ 002
20
30
40 50 2θ (degree)
60
0 0.0
70
0.2
0.4
0.6 0.8 Distance (nm)
1.0
1.2
665 666
Figure 1. (a) XRD patterns of the prepared samples. (b) EDS linescan (from TEM)
667
and (c) elemental mappings (from TEM) of CN-Cu(II)-CuAlO2.
668
28 ACS Paragon Plus Environment
Page 29 of 33
Environmental Science & Technology
669
2.0x10
4
1.5x10
4
+
CuAlO2 Cu 2p +
1.0x10 5.0x10
3
(b) 1.8x10
4
1.5x10
4
1.2x10
4
9.0x10
3
6.0x10
3
3.0x10
3
Cu LMM-Auger
(Cu ) 932.1
+
Cu (1848.8)
1830
4
2+
Cu :Cu =1:0.61
CPS
CPS
(a) 2.5x104
1840 1850 1860 Auger Parameter (eV)
+
2+
Cu :Cu =1:1.79
CN-Cu(II)-CuAlO2 Cu 2p
Cu LMM-Auger
+
Cu (932.8)
+
Cu (1849.9) 2+
Cu (934.7)
1830
1840 1850 1860 Auger Parameter (eV) 2+
Cu satellite (941.6) (944.0)
2+
(Cu ) 934.6 2+
(Cu satellite)
0.0 925
0.0 930
670
935 940 945 Binding energy (eV)
950
4
(c) 1.2x10
925
950
CN-Cu(II)-CuAlO2 O 1s
Plattice/Psurface=2.68
4
4x10
284.7 C-C
9.0x10
CPS
4
CPS
935 940 945 Binding energy (eV)
4
(d) 5x10
CN-Cu(II)-CuAlO2 C 1s
3
3
6.0x10
3x10
530.2 O-Cu(I)
531.1 O-Al(III)
532.0 surface -OH
4
2x10
286.5 C-O-M
3
3.0x10
284
286
288
532.5 C-O-Cu
4
288.7 N-C=N
1x10
0.0
671
930
0 528
290
Binding energy (eV)
530
532 534 Binding energy (eV)
536
672
Figure 2. XPS spectra in Cu 2p for (a) CuAlO2 and (b) CN-Cu(II)-CuAlO2, (c) C 1s
673
and (d) O 1s for CN-Cu(II)-CuAlO2.
674
29 ACS Paragon Plus Environment
Environmental Science & Technology
(a) 1.5x10
5
g-C3N4
g=1.996
Intensity (a.u.)
Page 30 of 33
1.0x10
5
5.0x10
4
CuAlO2
g=2.084
0.0 2.6
2.4
2.2
675
CN-Cu(II)-CuAlO2
2.0 1.8 g Value
1.6
1.4
(b)
(c)
676 677
Figure 3. (a) EPR spectra of the fresh g-C3N4, CuAlO2 and CN-Cu(II)-CuAlO2
678
samples. (b) and (c) DFT calculations for the optimized structure (left) and the
679
corresponding two-dimensional valence-electron density color-filled maps (right) of
680
the CN-Cu(II)-CuAlO2 model in -Cu(II)-CN vision fragment (b) and -CN vision
681
fragment (c). Orange, red, gray, blue and white circles denote Cu, O, C, N, and H
682
atoms, respectively. The valence-electron density is given in e/Å3.
683
30 ACS Paragon Plus Environment
Page 31 of 33
Environmental Science & Technology
684
(b) 1.0
(a) 1.0
0.4
-1.5 -2.0 -2.5
g-C3N4
-3.0
CuAlO2
-3.5
g-C3N4
0.2
k=0.027
CN-Cu(II)-CuAlO2
0
30 60 90 Reaction time(min)
CuAlO2
120
0
g-C3N4
120
Phenol 2-Chlorphenol Ibuprofen Phenytoin
0.6 0.4
CuAlO2 CN-Cu(II)-CuAlO2
0
30
60 90 120 150 Reaction time (min)
(d) 1.0
180
pH=5 pH=6 pH=7 pH=8 pH=9
0.8
BPA C/C0
0.8
0.6 0.4 0.2
0.2
686
0.7
0.5
30 60 90 Reaction time (min)
(c) 1.0
0.0
0.8
0.6
CN-Cu(II)-CuAlO2
685
C/C0
TOC/TOC0
k=0.0011
-1.0
0.6
0.0
0.9
k=0.0006
0.0 -0.5
ln(C/C0)
BPA (C/C0)
0.8
0
30
60
90
0.0
120
0
30
Reaction time (min)
60 90 Reaction time (min)
120
687
Figure 4. (a) BPA degradation curves in various suspensions with H2O2 (Insert shows
688
the corresponding kinetic curves). (b) TOC removal curves during BPA degradation.
689
(c) Decomposition curves of different pollutants in CN-Cu(II)-CuAlO2 suspensions
690
with H2O2. (d) Effect of initial pH values for BPA degradation. Reaction conditions:
691
Natural initial pH (pH 6-7, except d), initial pollutants 25 mg/L, initial H2O2 10 mM,
692
catalyst 1.0 g L−1.
693
31 ACS Paragon Plus Environment
Environmental Science & Technology
Page 32 of 33
694
(a)
1.0
(b)
Decomposition of H2O2
2302
CN-Cu(II)-CuAlO2
Absorbance (a. u.)
C/C0
CuAlO2/D2O
2290
0.8 0.6 0.4 0.2
1193
2560 CuAlO2/BPA/D2O
1190 2285
2550
CN-Cu(II)-CuAlO2/D2O
2304 CN-Cu(II)-CuAlO2/BPA/D2O
1190
2560
0.0
CuAlO2
0
30
695
(d)
•
BMPO- OH (with H2O2)
0
CuAlO2/H2O2
2400
•
2100 1800 1500 -1 Wavenumber (cm )
•−
0
CuAlO2/H2O2 CN-Cu(II)-CuAlO2/H2O2
CN-Cu(II)-CuAlO2/BPA/H2O2
CN-Cu(II)-CuAlO2/BPA/H2O2
3500 3520 Magnetic field (G)
3540
1200
BMPO-HO2/O2 (with H2O2)
CN-Cu(II)-CuAlO2/H2O2
3480
696
2700
180
Intensity (a.u.)
Intensity (a.u.)
(c)
60 90 120 150 Reaction time (min)
3480
3500 3520 Magnetic field (G)
3540
697
Figure 5. (a) H2O2 decomposition curves during BPA degradation in CuAlO2 and
698
CN-Cu(II)-CuAlO2 suspensions. Reaction conditions: Natural initial pH, initial
699
pollutants 25 mg/L, initial H2O2 10 mM, catalyst 1.0 g L−1. (b) ATR-FTIR spectra of
700
CuAlO2 and CN-Cu(II)-CuAlO2 in D2O environment with/without BPA. BMPO
701
spin-trapping EPR spectra for (c) •OH and (d) HO2•/O2•− in various suspensions in the
702
presence of H2O2 with/without pollutants.
703
32 ACS Paragon Plus Environment
Page 33 of 33
Environmental Science & Technology
704
Table of Contents Art
705
33 ACS Paragon Plus Environment