Journal Pre-proof All-solid-state Z-scheme WO3 nanorod/ZnIn2 S4 composite photocatalysts for the effective degradation of nitenpyram under visible light irradiation Mengling Tang (Methodology) (Software) (Writing - original draft), Yanhui Ao (Conceptualization) (Writing - review and editing) (Visualization), Peifang Wang (Writing - review and editing) (Project administration) (Funding acquisition), Chao Wang (Supervision) (Project administration)
PII:
S0304-3894(19)31667-X
DOI:
https://doi.org/10.1016/j.jhazmat.2019.121713
Reference:
HAZMAT 121713
To appear in:
Journal of Hazardous Materials
Received Date:
30 September 2019
Revised Date:
13 November 2019
Accepted Date:
17 November 2019
Please cite this article as: Tang M, Ao Y, Wang P, Wang C, All-solid-state Z-scheme WO3 nanorod/ZnIn2 S4 composite photocatalysts for the effective degradation of nitenpyram under visible light irradiation, Journal of Hazardous Materials (2019), doi: https://doi.org/10.1016/j.jhazmat.2019.121713
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier.
All-solid-state Z-scheme WO3 nanorod/ZnIn2S4 composite photocatalysts for the effective degradation of nitenpyram under visible light irradiation
Mengling Tang, Yanhui Ao*, Peifang Wang, Chao Wang
ro of
Key Laboratory of Integrated Regulation and Resource Development on Shallow Lakes of
Ministry of Education, College of Environment, Hohai University, No.1, Xikang road, Nanjing,
author. Tel./ fax: +86 25 83787330,
re
*Corresponding
-p
210098, China
Jo
ur
na
Graphical Abstract
lP
E-mail address:
[email protected] (Yanhui Ao)
1
Highlight
All-solid-state Z-scheme WO3 nanorod/ZnIn2S4 composite photocatalysts were synthesized.
The photocatalyst exhibited high performance for the degradation of nitenpyram.
ro of
The rate constant improved by a factor of 3.8 and 2.5 compared to WO3 and ZnIn2S4,.
-p
A deductive degradation pathway of nitenpyram was proposed.
re
Abstract: A Z-scheme WO3/ZnIn2S4 photocatalyst was synthesized via a simple solvothermal method. Compared with pure WO3 and ZnIn2S4, photocatalytic
lP
experiments showed that these Z-scheme photocatalysts exhibited enhanced activity
na
for the degradation of nitenpyram (NTP). The apparent rate constant (k) of NTP degradation on 50WZ (WO3/ 50 wt% Znln2S4) was 0.042 min-1 (~3.8 times higher
ur
than WO3 and ~2.5 times higher than ZnIn2S4). Photoluminescence (PL), photocurrent (PC), and electrochemical impedance spectroscopy (EIS) showed that
Jo
the separation and transfer efficiency of photogenerated carriers in 50WZ was markedly enhanced, which was favorable for improving its photocatalytic activity. Active
species
capture
experiments
and
electron
spin
resonance
(ESR)
measurements showed that superoxide radicals and holes were the main active species for NTP degradation, and they confirmed the formation of the Z-scheme 2
structure. Furthermore, a possible NTP degradation pathway was deduced based on the results of high-performance liquid chromatography mass spectrometry (HPLC-MS).
Key words: photocatalysis; Z-scheme; nitenpyram; neonicotinoids
ro of
1. Introduction
The degradation of nitenpyram (NTP), which is an effective and widely used
neonicotinoids pesticide in agriculture[1, 2], in the environment is of great concern
-p
because of its persistence and toxicity in water[3], resulting in a marked impact on
re
the ecology and human health[4]. There are numerous problems with current NTP degradation methods: the microbiological method has restricted reaction conditions
lP
and a low efficiency in general[5]; both the electrocatalytic and low-temperature
na
plasma methods have high energy consumption[6, 7]; in contrast, the photocatalytic method has attracted increasing attention because of its environmental friendliness 9]
. Photocatalysts that only respond to ultraviolet light have
ur
and efficiency[8,
difficulties degrading complex organic compounds, hence, visible light responsive
Jo
photocatalysts have become a significant area of research[10]. Semiconductor materials, such as oxides, nitrides and sulfides, have been
extensively studied[11-13]. Graphite phase C3N4 is a visible light responsive metal-free semiconductor, however, it has poor absorption and rapid photogenerated carrier recombination, which greatly limits its application as a photocatalyst[14-17], and the 3
heterojunctions related to this material have been widely researched. Sulfide semiconductors have an excellent photoelectric property and an appropriate bandgap size, hence, they have a wider optical response range and higher photocatalytic activity than g-C3N4[18-21]. This makes them suitable for the photocatalytic degradation of complex organic compounds. ZnIn2S4 is a ternary sulfide, a typical semiconductor member of the AB2X4 group, and a good option for
ro of
use as a visible light responsive photocatalyst because of its suitable bandgap energy (2.2–2.8 eV)[22, 23]. ZnIn2S4 has two common polymorphs based on the hexagonal and cubic phases, both of which have good charge storage capacity and
26]
, the cubic ZnIn2S4 displays stable
re
photoluminescence and photoconductivity[25,
-p
electrochemical performance[24]. It has been reported that hexagonal ZnIn2S4 shows
thermoelectric properties[27], while it is a high-pressure phase[28]. Chen et al,
lP
synthesized these two ZnIn2S4 phases with different precursors and investigated
na
their degradation properties: MO was almost completely degraded by the cubic phase, while the hexagonal phase had a degradation efficiency of only 52%, but it
ur
had a higher stability than the cubic phase[28]. Studies have considered the selection of suitable bandgap semiconductors to form heterojunctions to address
Jo
the issues of sulfide photocorrosion and the high recombination rate of photogenerated electron-hole pairs in a single photocatalyst[29, 30]. Photocatalysts forming traditional heterojunction with ZnIn2S4 have been reported, such as g-C3N4/ZnIn2S4[31], ZnIn2S4/CdIn2S4[32], TiO2/ZnIn2S4[33]. However, the reduction and oxidation ability of photo-induced carriers in conventional heterojunction are usually 4
lower than one of its components[34], therefore, the Z-scheme structure is a better option[35-37]. However, few studies have investigated Z-type photocatalyst containing ZnIn2S4. This work investigated semiconductors with a suitable bandgap for the formation of a Z-type heterojunction with ZnIn2S4. WO3 is considered as a promising material because of its electrochromic and photocatalytic properties[38, 39]. Moreover, because of its narrow band gap (2.4–2.8 eV)
ro of
and good stability, WO3 has been reported to have good photocatalytic activity for dye
degradation[40-42]. The ECB and EVB of ZnIn2S4 are at approximately -0.80 eV and
1.52 eV[43], while the ECB and EVB of WO3 are at approximately 0.74 eV and 3.44
-p
eV[44], respectively. Thus, photogenerated electrons can theoretically transfer from the
re
CB of WO3 to the VB of ZnIn2S4 to form a Z-scheme structure. A composite ZnIn2S4 nanosheets/ WO3 nanorods core-shell structure composite has been reported via a
lP
solvothermal method[45], and a 2D/2D WO3/ZnIn2S4 photocatalyst has been prepared
na
via a electrostatic self-assembly method[46]. However, these catalysts were only applied for photocatalytic hydrogen production. Furthermore, research on the
ur
photocatalytic degradation of neonicotinoids is minimal to date. In this paper, a composite photocatalyst was prepared via a simple one-step hydrothermal
Jo
method and applied to the photocatalytic degradation of neonicotinoid pesticides.
In this work, a new Z-scheme WO3/ZnIn2S4 photocatalyst was prepared by a simple solvothermal method. The crystal structure, element composition and optical properties of the hybrids were characterized using various characterization methods. 5
NTP was chosen as a target contaminant to assess the photocatalytic activities of the samples. In addition, a possible pathway for NTP degradation by 50WZ and a rational Z-scheme mechanism are proposed.
2. Results and discussion 2.1 Characterizations of WO3/Znln2S4 The morphologies of WO3, Znln2S4 and 50WZ samples were investigated via
ro of
TEM. As can be seen from Fig. 1a, the pure WO3 exhibited a nanorod morphology with a diameter of ~35 nm and a length of ~210 nm. HRTEM image (Fig. 1b) showed
that the lattice spacing of the WO3 crystal was 0.391 nm, which corresponds to the
-p
(001) plane of WO3. From Fig. 1c, we can see that pure ZnIn2S4 shows a flower-like
re
morphology with a diameter of ~3.5 μm. From the high-magnification local micrograph, we can see a folded and flaky structure on the surface of the nanoflower.
lP
The HRTEM (Fig. 1d) showed that the lattice fringe of this nanoflower was ~0.32 nm,
na
corresponding to the (102) plane of hexagonal ZnIn2S4. On observing the 50WZ hybrid (Fig. 1e), it could be determined that ZnIn2S4 did not present a flower-like
ur
structure in the mixture, but formed a sheet structure on the surface of the nanorods. Through HRTEM analysis (Fig. 1f), it was found that the lattice fringes of these two
Jo
substances were still present in 50WZ, which can be ascribed to the important role of WO3 groups for crystal growth and the ZnIn2S4 self-assembly[43]. WO3, ZnIn2S4 and 50WZ were investigated via XPS to further demonstrate the presence of each element in the mixture and the interaction between components. From the survey figure (Fig. 2a), the 50WZ hybrid showed all the peaks of WO3 and 6
ZnIn2S4, and all the expected elements, W, O, Zn, In and S, were detected. For the high-resolution W 4f spectrum (Fig. 2b), the two peaks at binding energies of 35.7 eV and 37.8 eV could be attributed to W4f7/2 and W4f5/2[47]. For the O 1s spectrum (Fig. 2c), the peak at a binding energy of 530.4 eV belongs to the oxygen atom that forms the strong W=O bond in WO3[48]. In Fig. 2d, two peaks at 1021.3 eV and 1044.4 eV can be attributed to Zn 2p3/2 and Zn 2p1/2, respectively[49]. Two binding
ro of
energy in the high-resolution In 3d spectrum (Fig. 2e) at 444.9 eV and 452.4 eV may
belong to In 3d5/2 and In 3d3/2[31]. In a high-resolution S 2p spectrum (Fig. 2f), the
two peaks at 161.5 eV and 162.7 eV can be assigned to S 2p3/2 and S 2p1/2,
-p
respectively[31]. The results indicate that the valence states of Zn, In and S are +2, +3
re
and -2, respectively[31, 43].
The XRD spectra of the prepared powders are shown in Fig. 3a. The peaks of
lP
the bare ZnIn2S4 at 2θ = 21.6°, 27.7°, 30.4°, 47.2°, 52.4° and 55.6° were in
na
agreement with those of the hexagonal phase ZnIn2S4 (JCPDS No. 65-2023). The typical diffraction peaks of bare WO3 at 2θ = 13.9°, 22.7°, 28.2° and 36.6° were
ur
indexed as the (100), (001), (200) and (201) planes of WO3 (JCPDS No. 33-1387). There were no impurity peaks in the pure sample patterns, and the main diffraction
Jo
peaks were consistent with the standard materials, which indicates that the prepared samples had a good crystallinity and purity. The WO3/ZnIn2S4 hybrids exhibited almost all diffraction peaks associated with the two pure samples. Because the diffraction peak at 2θ = 28.2° of WO3 was very close to the ZnIn2S4 peak at 2θ = 27.7°, a combined higher intensity peak can be observed in the hybrids. 7
In general, the PC and EIS mainly investigated the separation efficiency of the photogenerated electron-hole pairs[50,
29]
. It can be seen from Fig. 3b that WO3,
ZnIn2S4 and 50WZ all had an immediate response at the moment of turning on and off the light, but the photocurrent generated by 50WZ was much higher than WO3 and ZnIn2S4, which demonstrated that the photogenerated carriers in 50WZ had the longest lifetime. The EIS diagram (Fig. 3c) displayed the impedance changes for the
ro of
three samples. The arc of the Nyquist plot of 50WZ was the smallest, indicating that
the charge transfer resistance in 50WZ was relatively smaller than others. Therefore,
the separation efficiency of the photogenerated carriers in the hybrid 50WZ sample
-p
were much higher than the other two pure samples.
re
In addition, from the PL spectrum (Fig. 3d), we can see obvious emission peaks of all the tested samples appear at ~460 nm, and the peak intensity for 50WZ was
lP
significantly lower than that for the other two pure samples. This demonstrates that
na
the radiative recombination of photogenerated carriers in 50WZ was lower than that for the pure samples[51], which is in agreement with PC and EIS measurement
ur
results.
The UV-vis DRS spectrum shows the optical properties of WO3, ZnIn2S4 and
Jo
50WZ. From Fig. 4a, the absorption edge is at ~450 nm for bare WO3 and ~600 nm for bare ZnIn2S4, showing a relatively wide visible light absorption range. The composite 50WZ sample shows broadened light absorption compared with the bare WO3, for the interaction between WO3 and ZnIn2S4 in the composite. The band gap energy (Eg) of a semiconductor can be computed according to the following formula: 8
αhv = A(hv-Eg)n/2
(1)
In which, α, h, v and A are the absorption coefficient, Planck's constant, light frequency and a constant, respectively. n is related to the type of semiconductor (n = 1 or 4, for a direct or indirect transition, respectively). Both n values of WO3 and ZnIn2S4 are 1[52, 43], and the Eg for WO3 and ZnIn2S4 can be computed to be 2.75 eV and 2.4 eV (Fig. 4b), respectively[44,
28]
. The valence band potential (EVB) and
ro of
conduction band potential (ECB) can be calculated by the following formulas[53, 31]: EVB = χ - Ee + 0.5 Eg
(2)
ECB = EVB - Eg
(3)
-p
Wherein, χ is the absolute electronegativity of a compound, which is the geometric
re
mean of the electronegativity of the constituent atoms. For WO3 and ZnIn2S4, the values of χ are 6.49 eV and 4.86 eV, respectively[44, 31]. Ee is the energy of free
lP
electrons on the hydrogen scale (4.5 eV)[53, 52, 43]. Thus, the EVB values of WO3 and
na
ZnIn2S4 were computed as 3.36 eV and 1.56 eV, and the corresponding ECB are 0.61 eV and -0.84 eV, respectively, which are basically consistent with the values reported
ur
in other literatures[44, 43].
The N2 adsorption-desorption isotherms of the prepared samples are shown
Jo
in Fig. 5. These samples exhibit type IV isotherms with H3 hysteresis loops, indicating that these samples have mesoporous structures. The BET surface areas of WO3, ZnIn2S4 and 50WZ are 30.04, 17.31 and 37.69 m2 g-1, respectively. As we know, a higher specific surface area provides abundant active sites for contact and interaction with pollutant molecules, improving the photocatalytic 9
effect. Therefore, the enhanced specific surface area of 50WZ can improve its photocatalytic activity. 2.2 Photocatalytic activities of WO3/Znln2S4 Fig. 6a shows the photocatalytic activities for NTP degradation on the as-prepared samples under a same condition. Under dark conditions for 1 h, the NTP adsorption rates by WO3, ZnIn2S4, 25WZ, 50WZ, 75WZ, 100WZ and
ro of
PM-50WZ were 5.7%, 7.0%, 10.0%, 13.8%, 13.2%, 9.2% and 4.8%, respectively. The blank test demonstrated that there was minimal NTP degradation under
visible light irradiation without a photocatalyst, implying that NTP is relatively
-p
stable and difficult to hydrolyze. The photocatalytic activities of all the WZ mixtures
re
were higher than those of pure WO3 and ZnIn2S4 and 50WZ had the best photocatalytic effect. After 40 min of irradiation, 3 ppm NTP was degraded about
lP
74.1% in the presence of 50WZ, while the degradation efficiency was only 31.8%
na
with WO3 and 47.2% with ZnIn2S4. The reasons may be as follows, when the ZnIn2S4 content was relatively lower, a large amount of photogenerated electrons
ur
were present in WO3, which recombined with the photogenerated holes[31]. However, when the ZnIn2S4 content was relatively higher, the excess ZnIn2S4 flakes on the
Jo
surface of WO3 nanorods agglomerated, resulting in a lower contact area with the NTP molecules and reduced photocatalytic activity[54], and the 50WZ composition was the optimal one for photocatalytic degradation. In addition, a physical mixture with the same mass ratio as 50WZ had a lower effect. The apparent rate constant (k) for 50WZ was 0.042 min-1, which is 3.8 times higher than that of WO3 and 2.5 10
times higher than that of ZnIn2S4 (Fig. 6b). Therefore, the results show that 50WZ has an enhanced photocatalytic activity. To find out the main active substances on degradation NTP by 50WZ, an active species capture experiment were carried out. From the degradation map after adding capture agents for different active species (Fig. 6c), the addition of TBA had little effect on NTP degradation by 50WZ, while the addition of BZQ has a great effect,
ro of
with the degradation rate decreasing from 74.1% to 10.1%. After adding Na2-EDTA,
the degradation rates decreased from 74.1% to 40.1%. Thus, the effects of these three
investigated active species on NTP degradation by 50WZ are •O2− > h+ > •OH.
-p
Considering the CB of ZnIn2S4 is sufficient to produce •O2− to participate in the
re
reaction for the CB potential of ZnIn2S4 (-0.84 eV) is higher than the standard reduction potential of •O2−/O2 (-0.33 eV), the result is rational.
lP
Total organic carbon (TOC) analysis is an effective method to evaluate the
na
mineralization rate of pollutant. As can be seen from Fig. 6d, the TOC removal rate reached 66.5% after extending the photocatalytic time to 120 min.
ur
Accordingly, it can reasonably be hypothesized that as the photocatalytic degradation time continues to be increased, NTP can increasingly be mineralized
Jo
by the prepared catalysts. 2.3 The stability of 50WZ photocatalyst In order to explore the reusability of our samples, we conducted 4 times repeated experiments by recycle and reuse 50WZ. After each photocatalytic test, the sample was centrifuged and rinsed with ultrapure water several times, and then dried 11
for the next run. From Fig. 7a, the degradation efficiency of NTP for the photocatalyst on the first cycle was 74.1%, and the degradation rate was 68.2% after four cycles. This indicated that the reusability and stability of the material was relatively good. A structural comparison for the fresh and used 50WZ samples is shown in Fig. 7b. The positions of the characteristic peaks for the used 50WZ are basically
ro of
unchanged, indicating that the structure of the photocatalyst did not change after use.
In addition, XPS analysis of the 50WZ surface elements after four cycles
-p
was compared with the unused 50WZ. From Fig. ESI 2, it can be seen that the
re
positions of the W 4f, O 1s, Zn 2p, In 3d and S 2p diffraction peaks of the fresh and used 50WZ were nearly the same. However, the S 2p diffraction peak
lP
intensity at 162.7 eV decreased, which may be due to the depletion of the S
na
element in the 50WZ sample after 4 cycles. This change is relatively small compared with the severe photocorrosion of pure sulfide. Therefore, the results
ur
show that the prepared 50WZ photocatalyst had a good stability. 2.4 Possible photocatalytic mechanism in the WO3/Znln2S4 system
Jo
To further reveal the electron transport modes in the 50WZ hybrid, ESR was
used to detect •O2− and •OH in this sample (Fig. 8a-b). Under dark conditions, no •O2− or •OH radicals formed in 50WZ, while the signals for both •O2− and •OH appeared and increased with extending irradiation time, which indicates that the 50WZ could produce •O2− and •OH under light irradiation. 12
Based on this, a possible photocatalytic mechanism is proposed (Fig. 9). If the composite photocatalyst was a traditional heterojunction, the photogenerated electrons would transfer from the CB of ZnIn2S4 to the CB of WO3, and •O2− would not be generated, because the CB position of WO3 (0.61 eV) was lower than the potential of •O2−/O2 (-0.33 eV). Photogenerated holes would transfer from the VB of WO3 to the VB of ZnIn2S4, and •OH would not be generated, for the VB position of
ro of
ZnIn2S4 (1.56 eV) is higher than the potential of •OH/OH- (1.99 eV) and •OH/H2O (2.68 eV). However, this is not in agreement with our ESR results. In a Z-scheme
structure, photogenerated electrons can transition from the VB to the CB of the WO3
-p
semiconductor, then, they would transfer to the VB of ZnIn2S4 through a Z-scheme
re
electron transport channel, and finally transition to its conduction band. In this structure, the VB position of WO3 would be lower than the potential required to
lP
produce •OH, and the CB position of ZnIn2S4 would be higher than the potential
na
needed to generate •O2−, so that the 50WZ composite would produce both •O2− and •OH. This is consistent with the observed results of the above capture and ESR
ur
experiments. Thus, we believe that the 50WZ composite is more in line with a Z-scheme photocatalytic mechanism.
Jo
2.5 Degradation pathway of NTP by 50WZ Based on the record of high-performance liquid chromatography mass
spectrometry (HPLC-MS), we analyzed the main intermediate structures and proposed a possible degradation pathway for NTP (m/z = 270.72 g mol-1) using 50WZ (Table 1 and Fig. 10). The results showed a major peak with m/z = 270.8, 13
corresponding to the NTP molecule. With extending reaction time, NTP was oxidized and degraded into many other small molecules with m/z values of 257.5, 227.9, 171.9, 224.9, 191.8, 163.8 and 122.7. Usually, the oxidation of organic compounds is mainly carried out at the oxidation site of organic molecules, and the oxidation sites are mainly located on the functional groups. In Fig. 10, the methyl group on the side chain of the NTP molecule may be removed under •O2−
ro of
and h+ attacking to form P1 (m/z = 256.69 g mol-1). Then under the action of •O2− and •OH, the furthest nitro group from the pyridine ring would denitrify to form a
hydroxyl group, forming P2 (m/z = 227.69 g mol-1)[55]. Studies have indicated out
-p
that the intermediate P2 is unstable and readily converts to P3 (m/z = 227.69 g
re
mol-1) in solution. With continuous attack of •O2− and h+, the ethyl group that connects the amino group and a ether bond in P3 may break off, forming P4 (m/z =
lP
170.64 g mol-1)[56]. We also detected the presence of 2-chloropyridine P12 (m/z =
na
131.09 g mol-1) on the HPLC-MS spectrogram, which may be formed by the oxidation of the pyridine ring side chain groups in P4. In addition, the
ur
carbon-carbon double bond is susceptible to oxidation reaction, and NTP may also shed its methyl-linking nitro groups on side chain to form P5 (m/z = 225.72 g
Jo
mol-1) under the attack of •O2−, •OH and h+[56]. Then, chlorine atoms in the 2-chloropyridine ring at position 2 may be break by active species to form P6 (m/z = 191.27 g mol-1)[57]. Under sustaining attack of •O2− and h+, methyl and nitro groups are continuously oxidized to form product P7 (m/z = 163.22 g mol-1). We all know that the amino group of organic compounds is vulnerable to attack by 14
oxidizing molecules, hence, P7 is easily oxidized to P8 (m/z = 122.17 g mol-1). However, although the oxidation reaction will first take place at oxidation sites, different initial oxidation sites will also lead to different oxidation paths. P11 (m/z = 286.67 g mol-1), a substance with a larger molecular weight than NTP, was also detected, which may be formed via oxidation of methyl groups on the side chain of NTP molecules into carboxyl groups. In the degrading solution, the and P10
(m/z = 131.09 g
ro of
ring-opening product P9 (m/z = 155.09 g mol-1)
mol-1)[55] were also detected, which were the main intermediates and corresponded
with the strong signals observed in the HPLC-MS spectrum. Combined with the
-p
TOC degradation results, when the irradiation time was extended to 120 min, the
re
TOC degradation efficiency reached 30%. Therefore, it can reasonably be inferred that NTP might eventually be completely oxidized to carbon dioxide and
na
3. Conclusions
lP
water by extending the degradation time.
A series of WO3/ZnIn2S4 heterojunctions were synthesized via a facile solvent
ur
thermal method. These hybrids exhibit enhanced photocatalytic activity for the photocatalytic degradation of NTP, and the 50WZ sample exhibited the best
Jo
photocatalytic performance compared with the pure WO3 and ZnIn2S4 samples and their other compounds of different proportions. This enhanced performance could mainly be ascribed to the Z-scheme heterostructure of the 50WZ, which helped extend the life of the photogenerated carriers. We believe this work provides a strong base for further study of the degradation of neonicotinoids in the environment 15
using photocatalysts.
Author Contributions Section Mengling Tang: Methodology, Software, Writing - Original Draft Yanhui Ao: Conceptualization, Writing - Review & Editing, Visualization Peifang Wang: Writing - Review & Editing, Project administration, Funding acquisition Chao Wang: Supervision, Project administration
ro of
Declaration of interests The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
-p
Acknowledgements
We are grateful for grants from Major Science and Technology Program for Water
re
Pollution Control and Treatment (2017ZX07203002), National Science Funds for
lP
Creative Research Groups of China (No.51421006), Natural Science Foundation of China (51679063), The World-Class Universities (Disciplines) and Characteristic
na
Development Guidance Funds for the Central Universities, the Key Program of National Natural Science Foundation of China (No. 91647206), the National Key Plan
Jo
ur
for Research and Development of China (2016YFC0502203), and PAPD.
16
References [1] Peter J, Ralf N. Neonicotinoids-from zero to hero in insecticide chemistry [J]. Pest management science, 2010, 64 (11): 1084-1098. [2] Peter J, Ralf N, Michael S, et al. Overview of the status and global strategy for neonicotinoids [J]. Journal of Agricultural & Food Chemistry, 2011, 59 (7): 2897-2908.
ro of
[3] Sun C W, Jin J, Zhu J, et al. Discovery of bis-aromatic ring neonicotinoid analogues fixed as -configuration: Synthesis, insecticidal activities, and molecular docking studies [J]. Bioorganic & Medicinal Chemistry Letters, 2010, 20 (11):
-p
3301-3305.
re
[4] Elbert A, Haas M, Springer B, et al. Applied aspects of neonicotinoid uses in crop protection [J]. Pest management science, 2008, 64 (11): 1099-1105.
lP
[5] Rodriguez-Castillo G, Molina-Rodriguez M, Perez-Villanueva M, et al. Removal 14
C-Imidacloprid in
na
of Two Neonicotinoid Insecticides and Mineralization of
Biomixtures [J]. Bulletin of environmental contamination and toxicology, 2018, 101
ur
(1): 137-143.
[6] Li S P, Jiang Y Y, Cao X H, et al. Degradation of nitenpyram pesticide in aqueous
Jo
solution by low-temperature plasma [J]. Environmental Technology, 2013, 34 (12): 1609-1616.
[7] Li S P, Ma X R, Cao X H, et al. Electro-Catalytic Degradation of Nitenpyram Wastewater Using C/PTFE Gas Diffusion Electrode [J]. Advanced Materials Research, 2013, 699: 747-752. 17
[8] Ahmed S, Rasul M G, Brown R, et al. Influence of parameters on the heterogeneous photocatalytic degradation of pesticides and phenolic contaminants in wastewater: A short review [J]. Journal of Environmental Management, 2011, 92 (3): 311-330. [9] Romina Z, Tilen K, Jure F, et al. Photocatalytic degradation with immobilised TiO2 of three selected neonicotinoid insecticides: imidacloprid, thiamethoxam and
ro of
clothianidin [J]. Chemosphere, 2012, 89 (3): 293-301.
[10] Zhang Y L, Han C, Nadagouda M N, et al. The fabrication of innovative single
crystal N, F-codoped titanium dioxide nanowires with enhanced photocatalytic
-p
activity for degradation of atrazine [J]. Applied Catalysis B: Environmental, 2015,
re
168: 550-558.
[11] Wang C, Ao Y H, Wang P F, et al. Controlled synthesis in large-scale of CdS
lP
mesospheres and photocatalytic activity [J]. Materials Letters, 2010, 64 (3): 439-441.
na
[12] Wang C, Wu D, Wang P F, et al. Effect of oxygen vacancy on enhanced photocatalytic activity of reduced ZnO nanorod arrays [J]. Applied Surface Science,
ur
2015, 325: 112-116.
[13] Mu R H, Ao Y H, Wu T F, Wang C, Wang P F, Synthesis of novel ternary
Jo
heterogeneous anatase-TiO2 (B) biphase nanowires/Bi4O5I2 composite photocatalysts for the highly efficient degradation of acetaminophen under visible light irradiation, Journal of Hazardous Materials, 2020, 382: 121083. [14] Ren
M
L,
Ao
Y
H,
Wang
P
F,
Wang
C,
Construction
of
silver/graphitic-C3N4/bismuth tantalate Z-scheme photocatalyst with enhanced 18
visible-light-driven performance for sulfamethoxazole degradation, Chemical Engineering Journal, 2019, 378: 122122. [15] Yang X F, Tang H, Xu J S, et al. Silver Phosphate/Graphitic Carbon Nitride as an Efficient Photocatalytic Tandem System for Oxygen Evolution [J]. Chemsuschem, 2015, 8 (8): 1350-1358. [16] Yu H G, Xiao P, Wang P, et al. Amorphous molybdenum sulfide as highly
ro of
efficient electron-cocatalyst for enhanced photocatalytic H2 evolution [J]. Applied Catalysis B: Environmental, 2016, 193: 217-225.
[17] Zou J P, Wang L C, Luo J M, et al. Synthesis and efficient visible light
-p
photocatalytic H2 evolution of a metal-free g-C3N4/graphene quantum dots hybrid
re
photocatalyst [J]. Applied Catalysis B: Environmental, 2016, 193: 103-109. [18] Dong Y Z, Xue Y S, Yang W W, et al. Visible light driven CdS/WO3 inverse opals
lP
with enhanced RhB degradation activity [J]. Colloids and Surfaces
A:
na
Physicochemical and Engineering Aspects, 2019, 561: 381-387. [19] Guo Y, Ao Y H, Wang P F, Wang C, Mediator-free direct dual-Z-scheme
ur
Bi2S3/BiVO4/MgIn2S4 composite photocatalysts with enhanced visible-light-driven performance
towards
carbamazepine
degradation,
Applied
Catalysis
B:
Jo
Environmental, 2019, 264: 479-490. [20] Yu H G, Huang X, Wang P, et al. Enhanced Photoinduced-Stability and Photocatalytic Activity of CdS by Dual Amorphous Cocatalysts: Synergistic Effect of Ti(IV)-Hole Cocatalyst and Ni(II)-Electron Cocatalyst [J]. Journal of Physical Chemistry C, 2016, 120 (7): 3722-3730. 19
[21] Yu Z, Yin B S, Qu F Y, et al. Synthesis of self-assembled CdS nanospheres and their photocatalytic activities by photodegradation of organic dye molecules [J]. Chemical Engineering Journal, 2014, 258: 203-209. [22] Chen Z X, Li D Z, Zhang W J, et al. ChemInform Abstract: Low-Temperature and Template-Free Synthesis of ZnIn2S4 Microspheres [J]. Inorganic Chemistry, 2010, 40 (8): 9766-9772.
ro of
[23] Fang F, Chen L, Chen Y B, et al. Synthesis and Photocatalysis of ZnIn2S4 Nano/Micropeony [J]. Journal of Physical Chemistry C, 2010, 114 (6): 2393-2397.
[24] Shang L, Zhou C, Bian T, et al. Facile synthesis of hierarchical ZnIn2S4
-p
submicrospheres composed of ultrathin mesoporous nanosheets as a highly efficient
Chemistry A, 2013, 1 (14): 4552-4558.
re
visible-light-driven photocatalyst for H2 production [J]. Journal of Materials
lP
[25] Georgobiani A, Ilyukhina Z, Tiginyanu I. Photoluminescence spectra of ZnIn2S4
na
single crystals after irradiation with neon ions [J]. Soviet Physics-Semiconductors, 1982, 16 (2): 231-232.
ur
[26] Romeo N, Dallaturca A, Braglia R, et al. Charge storage in ZnIn2S4 single crystals [J]. Applied Physics Letters, 1973, 22 (1): 21-22.
Jo
[27] Seo W S, Otsuka R, Okuno H, et al. Thermoelectric properties of sintered polycrystalline ZnIn2S4 [J]. Journal of Materials Research, 1999, 14 (11): 4176-4181. [28] Chen Y J, Hu S W, Liu W J, et al. Controlled syntheses of cubic and hexagonal ZnIn2S4 nanostructures with different visible-light photocatalytic performance [J]. Dalton Transactions, 2011, 40 (11): 2607-2613. 20
[29] Ren M L, Chen J, Wang P F, et al. Construction of silver iodide/silver/bismuth tantalate Z-scheme photocatalyst for effective visible light degradation of organic pollutants [J]. Journal of Colloid and Interface Science, 2018, 532: 190-200. [30] Wang M Y, Cai L J, Wang Y, et al. Graphene-draped semiconductors for enhanced photocorrosion resistance and photocatalytic properties [J]. Journal of the American Chemical Society, 2017, 139 (11): 4144-4151.
ro of
[31] Guo F, Cai Y, Guan W S, et al. Graphite carbon nitride/ZnIn2S4 heterojunction photocatalyst with enhanced photocatalytic performance for degradation of
tetracycline under visible light irradiation [J]. Journal of Physics and Chemistry of
-p
Solids, 2017, 110: 370-378.
re
[32] Yu Y G, Chen G, Wang G, et al. Visible-light-driven ZnIn2S4/CdIn2S4 composite photocatalyst with enhanced performance for photocatalytic H2 evolution [J].
lP
International Journal of Hydrogen Energy, 2013, 38 (3): 1278-1285.
na
[33] Xia Y, Li Q, Lv K L, et al. Heterojunction construction between TiO2 hollowsphere and ZnIn2S4 flower for photocatalysis application [J]. Applied Surface
ur
Science, 2017, 398: 81-88.
[34] Cui L F, Ding X, Wang Y G, et al. Facile preparation of Z-scheme WO3/g-C3N4
Jo
composite photocatalyst with enhanced photocatalytic performance under visible light [J]. Applied Surface Science, 2017, 391: 202-210. [35] Li B S, Lai C, Xu P, et al. Facile synthesis of bismuth oxyhalogen-based Z-scheme photocatalyst for visible-light-driven pollutant removal: Kinetics, degradation pathways and mechanism [J]. Journal of Cleaner Production, 2019, 21
225: 898-912. [36] Zhang M M, Lai C, Li B S, et al. Rational design 2D/2D BiOBr/CDs/g-C3N4 Z-scheme heterojunction photocatalyst with carbon dots as solid-state electron mediators for enhanced visible and NIR photocatalytic activity: Kinetics, intermediates, and mechanism insight [J]. Journal of Catalysis, 2019, 369: 469-481.
ro of
[37] Zhou P, Yu J G, Jaroniec M. All-Solid-State Z-Scheme Photocatalytic Systems [J]. Advanced Materials, 2014, 26 (29): 4920-4935.
[38] Nakajima T. Visible light photocatalytic activity enhancement for water
-p
purification in Cu(II)-grafted WO3 thin films grown by photoreaction of nanoparticles
re
[J]. Applied Catalysis B: Environmental, 2011, 108: 47-53.
[39] Srinivasan A, Miyauchi M. Chemically Stable WO3 Based Thin-Film for
lP
Visible-Light Induced Oxidation and Superhydrophilicity [J]. Journal of Physical
na
Chemistry C, 2012, 116 (29): 15421–15426.
[40] Chen D, Ye J H. Hierarchical WO3 hollow shells: dendrite, sphere, dumbbell, and
ur
their photocatalytic properties [J]. Advanced Functional Materials, 2008, 18 (13): 1922-1928.
Jo
[41] Guo Y F, Quan X, Lu N, et al. High photocatalytic capability of self-assembled nanoporous WO3 with preferential orientation of (002) planes [J]. Environmental science & technology, 2007, 41 (12): 4422-4427. [42] Shukla S, Chaudhary S, Umar A, et al. Surfactant functionalized tungsten oxide nanoparticles with enhanced photocatalytic activity [J]. Chemical Engineering Journal, 22
2016, 288: 423-431. [43] Liu H, Jin Z T, Xu Z Z, et al. Fabrication of ZnIn2S4–g-C3N4 sheet-on-sheet nanocomposites for efficient visible-light photocatalytic H2-evolution and degradation of organic pollutants [J]. RSC Advances, 2015, 5 (119): 97951-97961. [44] Chen S F, Hu Y F, Meng S G, et al. Study on the separation mechanisms of photogenerated electrons and holes for composite photocatalysts g-C3N4-WO3 [J].
ro of
Applied Catalysis B: Environmental, 2014, 150-151: 564-573.
[45] Ye L, Wen Z. ZnIn2S4 nanosheets decorating WO3 nanorods core-shell hybrids for boosting visible-light photocatalysis hydrogen generation [J]. International Journal of
-p
Hydrogen Energy, 2019, 44: 3751-3759.
re
[46] Tan P, Zhu A, Qiao L, et al. Constructing a direct Z-scheme photocatalytic system based on 2D/2D WO3/ZnIn2S4 nanocomposite for efficient hydrogen evolution under
lP
visible light [J]. Inorganic Chemistry Frontiers, 2019, 6 (4): 929-939.
na
[47] Zheng F, Guo M, Zhang M. Hydrothermal preparation and optical properties of orientation-controlled WO3 nanorod arrays on ITO substrates [J]. CrystEngComm,
ur
2013, 15 (2): 277-284.
[48] Li J, Zhu J W, Liu X H. Synthesis, characterization and enhanced gas sensing
Jo
performance of WO3 nanotube bundles [J]. New Journal of Chemistry, 2013, 37 (12): 4241.
[49] Ye L, Fu J L, Xu Z, et al. Facile one-pot solvothermal method to synthesize sheet-on-sheet reduced graphene oxide (RGO)/ZnIn2S4 nanocomposites with superior photocatalytic performance [J]. ACS Applied Material & Interfaces, 2014, 6 (5): 23
3483-3490. [50] Guo Y, Wang P F, Qian J, et al. Phosphate group grafted twinned BiPO4 with significantly enhanced photocatalytic activity: Synergistic effect of improved charge separation efficiency and redox ability [J]. Applied Catalysis B: Environmental, 2018, 234: 90-99. [51] Lu J S, Wang Y J, Fei L, et al. Fabrication of a direct Z-scheme type WO3
ro of
/Ag3PO4 composite photocatalyst with enhanced visible-light photocatalytic performances [J]. Applied Surface Science, 2017, 393: 180-190.
[52] Jiang L B, Yuan X Z, Zeng G M, et al. In-situ synthesis of direct solid-state dual
-p
Z-scheme WO3/g-C3N4/Bi2O3 photocatalyst for the degradation of refractory pollutant
re
[J]. Applied Catalysis B: Environmental, 2018, 227: 376-385.
[53] Chen Z H, Wang W L, Zhang Z G, et al. High-Efficiency Visible-Light-Driven
lP
Ag3PO4/AgI Photocatalysts: Z-Scheme Photocatalytic Mechanism for Their Enhanced
19346-19352.
na
Photocatalytic Activity [J]. The Journal of Physical Chemistry C, 2013, 117 (38):
ur
[54] He Y M, Wang Y, Zhang L H, et al. High-efficiency conversion of CO2 to fuel over ZnO/g-C3N4 photocatalyst [J]. Applied Catalysis B: Environmental, 2015, 168:
Jo
1-8.
[55] Li S P, Wang W R, Zeng X Y, et al. Electro-catalytic degradation mechanism of nitenpyram in synthetic wastewater using Ti-based SnO2–Sb with rare earth-doped anode [J]. Desalination and Water Treatment, 2015, 54 (7): 1925-1938. [56] Noestheden M, Roberts S, Hao C Y. Nitenpyram degradation in finished drinking 24
water [J]. Rapid Communications in Mass Spectrometry, 2016, 30 (13): 1653-1661. [57] Gonzalez-Marino I, Rodriguez I, Rojo L, et al. Photodegradation of nitenpyram under UV and solar radiation: Kinetics, transformation products identification and
Jo
ur
na
lP
re
-p
ro of
toxicity prediction [J]. Science of the Total Environment, 2018, 644: 995-1005.
25
Figure Captions: Fig. 1 TEM and HRTEM images of WO3 (a, b), ZnIn2S4 (c, d) and 50WZ (e, f). Fig. 2
XPS survery spectra of WO3, ZnIn2S4 and 50WZ (a), high-resolution spectra of W 4f
(b), O 1s (c), Zn 2p (d), In 3d (e), and S 2p (f). Fig. 3 XRD patterns of as-prepared photocatalysts (a), transient photocurrent response (b), EIS changes (c) and PL spectra (d) of WO3, ZnIn2S4 and 50WZ.
ro of
Fig. 4 UV-vis DRS spectra of WO3, ZnIn2S4 and 50WZ (a), and the plots of (ahv)1/2 versus hv for the band gap energies of WO3 and ZnIn2S4.
Fig. 5 N2 adsorption-desorption isotherms of WO3, ZnIn2S4 and 50WZ.
-p
Fig. 6 Photocatalytic activities of as-prepared samples for NTP degradation (a), the apparent
re
rate constants (k) for NTP degradation (b), capture experiment diagram of 50WZ for NTP
lP
degradation (c) and TOC removal by 50WZ (d).
Fig.7 Photocatalytic stability of 50WZ for NTP degradation after 4 cycles (a), XRD patterns
na
of freshand used 50WZ catalyst (b).
Fig. 8 ESR spectra of 50WZ photocatalyst under dark and visible light irradiation for
ur
DMPO-•O2- (a) and DMPO-•OH (b).
Fig. 9 Proposed possible photocatalytic mechanism of 50WZ composite.
Jo
Fig. 10 Proposed possible degradation pathway of NTP in the presence of 50WZ. Table 1 HPLC-MS data for the intermediate products during NTP degradation.
26
ro of -p re lP na
Jo
ur
Fig. 1 TEM and HRTEM images of WO3 (a, b), ZnIn2S4 (c, d) and 50WZ (e, f).
27
ro of -p re lP na
Fig. 2
XPS survery spectra of WO3, ZnIn2S4 and 50WZ (a), high-resolution spectra of W 4f
Jo
ur
(b), O 1s (c), Zn 2p (d), In 3d (e), and S 2p (f).
28
ro of -p re
lP
Fig. 3 XRD patterns of as-prepared photocatalysts (a), transient photocurrent response (b),
Jo
ur
na
EIS changes (c) and PL spectra (d) of WO3, ZnIn2S4 and 50WZ.
29
ro of
Fig. 4 UV-vis DRS spectra of WO3, ZnIn2S4 and 50WZ (a), and the plots of (ahv)1/2 versus hv
Jo
ur
na
lP
re
-p
for the band gap energies of WO3 and ZnIn2S4.
30
ro of
Jo
ur
na
lP
re
-p
Fig. 5 N2 adsorption-desorption isotherms of WO3, ZnIn2S4 and 50WZ.
31
ro of
-p
Fig. 6 Photocatalytic activities of as-prepared samples for NTP degradation (a), the apparent
re
rate constants (k) for NTP degradation (b), capture experiment diagram of 50WZ for NTP
Jo
ur
na
lP
degradation (c) and TOC removal by 50WZ (d).
32
Fig.7 Photocatalytic stability of 50WZ for NTP degradation after 4 cycles (a), XRD patterns
Jo
ur
na
lP
re
-p
ro of
of freshand used 50WZ catalyst (b).
33
Fig. 8 ESR spectra of 50WZ photocatalyst under dark and visible light irradiation for
Jo
ur
na
lP
re
-p
ro of
DMPO-•O2- (a) and DMPO-•OH (b).
34
ro of
Jo
ur
na
lP
re
-p
Fig. 9 Proposed possible photocatalytic mechanism of 50WZ composite.
35
ro of
Jo
ur
na
lP
re
-p
Fig. 10 Proposed possible degradation pathway of NTP in the presence of 50WZ.
36
Table 1 HPLC-MS data for the intermediate products during NTP degradation. Empirical formula
NTP
C11H15ClN4O2
Molecular structure
NH
P1
NH2
C10H13ClN4O2+H
Cl
NH2
C10H14ClN3O
Cl
N
C10H14ClN3O
256.69
227.9
227.69
227.9
NH2 O N N
-p
Cl
C8H11ClN2+H
170.64
C11H16ClN3
224.9
225.72
NH
191.8
191.27
NH
163.8
163.22
122.7
122.17
154.9
155.09
131.2
131.09
re
N
227.69
171.9
NH Cl
P5
257.5
N OH
P4
270.72
NO2
N
P3
270.8
N
N
P2
Exact m/z value (g mol-1)
NO2
N Cl
Observed m/z value (g mol-1)
ro of
Name
NH
lP
N
Cl
C11H17N3
N
na
P6
N
N
P9
N H N
C7H10N2
Jo
P8
C9H13N3
ur
P7
C8H13NO2
N H N OHC
N O
P10
C6H13NO2
N OH
O
37
P11
C10H11ClN4O4+H
NH2 N Cl
P12
286.67
113.9
113.54
N COOH
C5H4ClN N
Jo
ur
na
lP
re
-p
ro of
Cl
287.9 NO2
38