Photocatalytic Simultaneous Removal of Nitrite and Ammonia via a

2 hours ago - Nitrite and ammonia often coexist in waters. Thus, it is very significant to develop a photocatalytic process for the simultaneous remov...
1 downloads 0 Views 2MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 6411−6420

http://pubs.acs.org/journal/acsodf

Photocatalytic Simultaneous Removal of Nitrite and Ammonia via a Zinc Ferrite/Activated Carbon Hybrid Catalyst under UV−Visible Irradiation Jia Ye, Shou-Qing Liu,* Wen-Xiao Liu, Ze-Da Meng, Li Luo,* Feng Chen, and Jing Zhou Jiangsu Key Laboratory of Environmental Functional Materials; School of Chemistry, Biology and Material Engineering, Suzhou University of Science and Technology, Suzhou 215009, China

ACS Omega 2019.4:6411-6420. Downloaded from pubs.acs.org by 5.62.154.39 on 04/08/19. For personal use only.

S Supporting Information *

ABSTRACT: Nitrite and ammonia often coexist in waters. Thus, it is very significant to develop a photocatalytic process for the simultaneous removal of nitrite and ammonia. Herein, zinc ferrite/activated carbon (ZnFe2O4/AC) was synthesized and characterized by X-ray diffraction spectroscopy, transmission electron microscopy, Raman spectroscopy, and ultraviolet− visible diffuse reflectance spectroscopy. The valence band level of ZnFe2O4 was measured by X-ray photoelectron spectroscopy−valence band spectroscopy, and first-principles calculation was performed to confirm the band structure of ZnFe2O4. The as-synthesized ZnFe2O4/AC species functioned as a photocatalyst to simultaneously remove nitrite and ammonia under anaerobic conditions upon UV−visible light irradiation at the first stage. The results indicated that an average removal ratio of 92.7% with ±0.2% error for nitrite degradation for three runs was achieved in 50.0 mg/L nitrite + 100.0 mg/L ammonia solution with pH 9.5 under anaerobic conditions for 3 h at this stage; simultaneously, the removal ratio of 64.0% with ±0.2% error for ammonia was also achieved. At the second stage, oxygen gas was bubbled in the reactor to photocatalytically eliminate residual ammonia under aerobic conditions upon continuous irradiation. The results demonstrated that the removal ratios for nitrite, ammonia, and total nitrogen reached to 92.0, 90.0, and 90.2% at 12th hour, respectively, and the product released during photocatalysis is N2 gas, detected by gas chromatography, fulfilling the simultaneous removal of nitrite and ammonia. The reaction mechanism was exploited. nitrogen molecules by the pathway of NO3− → NO2− → NO → N2O → N2 under anaerobic conditions.4 Both nitrification and denitrification pathways are complicated and time-consuming, in which a carbon source is needed and temperature, pH, and dissolved oxygen are controlled painstakingly and accurately to maintain the activity of bacteria and removal efficiency of total nitrogen.5,6 Therefore, the development of a simple, economical, time-saving, and easygoing process for the complete removal of nitrogen is very desired.7,8 In theory, photocatalysis can realize both oxidization and reduction processes in viewpoint of electrochemistry if the electrode potential is appropriate. The photogenerated electrons on the conduction band of a semiconductor material could reduce a substrate molecule if the conduction band level (Ec) is more negative than the standard electrode potential (E0) of the substrate molecule. On the other hand, the photogenerated holes on the valence band of the semiconductor material could oxidize the other substrate molecule if the valence band level (Ev) is more positive than E0 of the

1. INTRODUCTION Nitrite (NO2−) is very toxic to human health. It can combine with hemoglobin to form methemoglobin in the body, decreasing the ability of red blood cells to carry oxygen. Moreover, it can be converted into carcinogenic nitrosamine, which causes hypertension, leukemia, brain tumor, stomach, and bowel cancers.1,2 Ammonia (NH3 or NH4+) can be easily transferred to NO2− in natural waters because of nitrifying bacteria, resulting in the coexistence of NO2− and NH3 in natural waters. Owing to their hazardous nature, the World Health Organization recommends the limit concentration of 3.0 mg/L for nitrite and 1.5 mg/L for ammonia in drinking water to protect humans from health risks.3 Therefore, the simultaneous removal of NO2− and NH3 is very significant to human health. The conventional biological removal of nitrogen can be achieved by nitrification and denitrification processes. The nitrification process is the oxidation of NH3 or NH4+ to nitrate (NO3−) through the pathway of NH4+/NH3 → NO2− → NO3− in the presence of aerobic bacteria. Nitrite could be the ultimate product formed during the nitrification process when the dissolved oxygen is inadequate, for example, at the deep bottom of waters. Subsequently, anaerobic bacteria start the denitrification process, which reduce NO3− or NO2− to © 2019 American Chemical Society

Received: July 24, 2018 Accepted: November 14, 2018 Published: April 8, 2019 6411

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

substrate molecule. In the photocatalytic field, Butler and coworkers reported that nitrite was oxidized to nitrate using TiO2 as the photocatalyst.9 With the assistance of oxalic acid or sodium oxalate as a hole scavenger and Au, Ag, and Pd as cocatalysts on TiO2, nitrite was reduced to ammonia and nitrogen gas or oxidized to nitrate.10,11 Kominami et al. utilized a metal-free TiO2 suspension to reduce nitrite to N2 using ammonia as a hole scavenger in the atmosphere of argon and investigated the effect of temperature on the reaction of ammonia and nitrite.12 Other photocatalysts such as CuInS2, KTaO3, LiNbO3, and BaLa4Ti4O15 have been developed to photocatalytically reduce nitrate.13−16 The studies are very valuable and interesting. However, ZnFe2O4/activated carbon (AC) used as a photocatalyst for simultaneous removal of NO2− and NH3 has not been reported and a mimicking process for removal of nitrogen has not been developed neither. Among photocatalysts, zinc ferrite (ZnFe2O4) possesses a prominent feature with a direct band gap of 1.98 eV. Ec of ZnFe2O4 was −0.90 V versus normal hydrogen electrode (NHE), and Ev was 1.08 V.17,18 As a result, it is often utilized for photocatalytic oxidization of organic pollutants,19−25 reduction of carbon dioxide,26 nitroaromatic compounds,27 and hydrogen production.28−30 Therefore, it is reasonable that the photogenerated electrons on the conduction band of ZnF2O4 could reduce NO2− (Ec is less negative than the 0 − , which is equal to 1.52 V standard electrode potential E NO 2 /N2 vs NHE) and the photoholes on the valence band of ZnFe2O4 could oxidize NH3 in theory (Ev is more positive than the standard electrode potential E N0 2 /NH3, which is equal to 0.057 V vs NHE). According to the theoretical analysis and facts above, we designed the photocatalytic simultaneous removal of nitrite and ammonia via coupling zinc ferrite and AC to form a zinc ferrite/AC (ZnFe2O4/AC) hybrid catalyst, which can mimic the denitrification process under anaerobic conditions, simultaneously removing nitrite and ammonia under UV− visible light irradiation, and the nitrification process under aerobic conditions to eliminate residual ammonia. AC is utilized to enhance the photocatalytic activity because it can adsorb the substrate molecules and promote the separation of photogenerated electron−hole pairs.31−35 The mimicking process can eliminate two toxic species, without addition of a carbon source, in addition to being cost-effective.

Figure 1. XRD patterns of AC (a), ZnFe2O4 (b), and ZnFe2O4/AC (c) samples.

our measurement). The results displayed that the D values of ZnFe2O4/AC and ZnFe2O4 particles were 6.9 and 7.2 nm, respectively. The size was confirmed by transmission electron microscopy (TEM) observations. 2.2. Morphological Observation. The morphological and structural characteristics of AC, ZnFe2O4, and ZnFe2O4/ AC species were analyzed through TEM observation. Figure 2A,B shows AC images. Figure 2C displays the ZnFe2O4 particles clearly; the size is very uniform, ranging from 6 to 9 nm, also consistent with those measured by XRD. Compared with Figure 2B (there is only AC), black ZnFe2O4 particles can been seen in Figure 2D, and the indistinct interface between ZnFe2O4 and AC shows their compact bonding. It is reasonable that Fe(III) and Zn(II) ions were adsorbed to porous caves of AC material during continuous stirring. AC functioned as templates by incorporating ZnFe2O4 precursors, yielding ZnFe2O4 in porous caves of AC. The high-resolution transmission electron microscopy (HRTEM) image in Figure 2E shows the distinct lattice fringes; the interplanar spacing between the adjacent lattice fringes is 0.25 nm, which is assigned to plane (311) of ZnFe2O4.38−40 A 0.29 nm d-spacing is attributed to plane (220) of ZnFe2O4 particles in Figure 2F,;the particles were embedded in AC, showing that the ZnFe2O4 particles were grown in caves of AC. 2.3. Raman Spectroscopic Characterization. The Raman spectra of AC, ZnFe2O4, and ZnFe2O4/AC samples are presented in Figure 3. Well-resolved Raman scattering peaks of ZnFe2O4 and ZnFe2O4/AC were observed at 337, 478, and 650 cm−1. The Raman peak at 650 cm−1 is attributed to the movement (mode A1g) of oxygen in tetrahedral AO4 groups, while the two low-frequency peaks are ascribed to metal ion vibration (mode F2g) involved in octahedral sites (BO6).41−43 The peaks at around 1320 and 1600 cm−1 correspond to the D band and G band of AC, respectively, in which the D band was caused by disordered sp3-bonded carbon atoms in the graphitic structure and the G band was caused by sp2-bonded carbon atoms in the graphitic structure.44 The large intensity ratio of ID/IG (round 1.3:1) for AC verified the amorphous nature of carbon.45 2.4. Band Gap and Band Structure. The optical absorption property of a semiconductor material reflects the ability of harvesting photos, thereby playing a significant role in determining the photocatalytic activity. The diffuse reflectance spectroscopy (DRS) spectra of the ZnFe2O4/AC and ZnFe2O4 samples have been measured and are shown in Figure 4A. Compared with curve b in Figure 4A, an enhanced absorption portion appeared in the wavelength region of more than 530 nm as displayed in curve a, implying that AC on ZnFe2O4

2. RESULTS AND DISCUSSION 2.1. X-ray Diffraction Characterization. The X-ray diffraction (XRD) patterns of ZnFe2O4/AC, ZnFe2O4, and AC samples are shown in Figure 1. The diffraction peaks located at 2θ = 29.7°, 35.0°, 42.7°, 53.0°, 56.5°, 62.0°, 70.4°, and 73.3° corresponded to the Bragg reflections from the (220), (311), (400), (422), (511), (440), (620), and (533) crystal planes of spinel-structured ZnFe2O4 (JCPDS card no. 22-1012),36−38 respectively. The peak at 26.2° is attributed to the (002) plane of graphene in AC. Compared with curve a, a small peak appeared in curve c, indicating the presence of AC in the ZnFe2O4/AC sample. The average diameters (D) of the samples could be estimated using Scherrer equation. D = Kλ /(W cos θ )

(1)

Here, W is the breadth of the observed diffraction peak at its half height, K is the shape factor (usually approximately 0.89), and λ is the wavelength of the X-ray source used (0.154 nm by 6412

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

Figure 2. TEM images of AC (a,b), ZnFe2O4 (c), and ZnFe2O4/AC (d) and HRTEM images of ZnFe2O4 (e) and ZnFe2O4/AC (f).

enhanced the absorption efficiency of incident photons. It will improve the use of solar irradiation. The direct and indirect band gaps of a semiconductor material can be determined by Tauc relation (αhv)n = A(hv − Eg )

(2)

where A is a constant, hv is the photon energy, and α is the absorption coefficient, while n = 2 for direct and n = 1/2 for indirect transition. A Tauc plot for direct transition was obtained by transforming data in Figure 4A and is shown in Figure 4B. The as-obtained direct band gaps for ZnFe2O4 and the ZnFe2O4/AC composite are 2.1 eV and 2.0 eV, respectively. The band gap for ZnFe2O4 was also confirmed by the first-principles calculation as shown in Figure 5A. As seen in Figure 5A, the direct transition at G point is equal to 1.968 eV, which is close to the measured value of 2.1 eV. In

Figure 3. Raman spectra of AC (a), ZnFe2O4 (b), and ZnFe2O4/AC (c).

Figure 4. UV−vis diffuse reflectance absorption spectra (A) of ZnFe2O4/AC (a) and ZnFe2O4 (b) samples. Tauc plots for direct transition of ZnFe2O4/AC (B). Figure 5. Band structure (A) and density of states (DOS) (B) of ZnFe2O4. 6413

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

Scheme 1. Reactions for the Measurement of Nitrite

addition, the band gap is related to particle size when the diameter of particles falls in the nanometer scale. The larger the band gap is, the smaller the particle size is due to the quantum confinement effects, although the band gap is not inversely proportional to the size of particles.46,47 ZnFe2O4 particles with approximate 10 nm size showed 1.78 eV of band gap,24 while the present ZnFe2O4 with approximate 7 nm displayed 2.1 eV of band gap, which confirmed the existence of quantum confinement effects in the system. Also, as seen in Figure 5B and Figure S1 in Supporting Information, the valence band is majorly composed of O 2p orbitals and the conduction band is majorly composed of Fe 3d orbitals; thus, the electron transition takes place from O 2p orbitals to Fe 3d orbitals upon UV−visible light irradiation. 2.5. Photocatalytic Removal of Nitrite and Ammonia. The concentrations of nitrite and ammonia were measured using ultraviolet visible spectroscopy based on Schemes 1 and 2 during the photocatalytic process under anaerobic conditions.48

Figure 6. The as-recorded ultraviolet−visible absorption spectra of nitrite (A) and ammonia (B) under anaerobic conditions in the presence of the ZnFe2O4/AC photocatalyst at various irradiation times. Photocatalytic removal of NN (C) and AN (D) under anaerobic conditions using 1.5 g/L ZnFe2O4/AC (AC 7 wt %) as the catalyst in 250.0 mL of solution with pH 9.5. The initial concentrations of NN and AN are 50.0 and 100.0 mg/L, respectively. (a) NN + AN + ZnFe2O4/AC under light irradiation; (b) NN + AN under light irradiation without ZnFe2O4/AC; (c) NN + ZnFe2O4/AC under light irradiation; (d) NN + AN + ZnFe2O4/AC in dark; and (e) AN + ZnFe2O4/AC under light irradiation.

Scheme 2. Reaction for the Measurement of Ammonia

The as-recorded ultraviolet visible absorption spectra for nitrite and ammonia are presented in Figure 6A,B at 0.0, 0.5, 1.0, 1.5, 2.0, 2.5, and 3.0 h, respectively. Figure 6A,B displayed unambiguously that the concentrations of nitrite and ammonia declined simultaneously with irradiation time under anaerobic conditions when ZnFe2O4/AC was utilized as the photocatalyst. The control tests were also conducted in the presence or absence of the ZnFe2O4/AC catalyst upon irradiation or in dark or in the absence of one component; the removal ratios for nitrite and ammonia are presented in Figure 6C,D, respectively. Curve a in Figure 6C shows an average removal ratio of 92.7% with ±0.2% error (three times) for nitrite degradation using ZnFe2O4/AC as the photocatalyst in the presence of 50.0 mg/L nitrite−N (NO2−−N, NN) + 100.0 mg ammonia−N (NH3−N, AN) under irradiation for 3 h, whereas the removal ratio approaches only 20.5% (curve b) in the absence of the catalyst under similar conditions. Parallel to the case of nitrite, the average removal ratio for degradation of ammonia for three times achieved 64.3% in the presence of ZnFe2O4/AC upon irradiation, whereas the removal ratio was only 18.8% in the absence of ZnFe2O4/AC even upon irradiation as shown in curve a and b in Figure 6D. It indicates distinctly that ZnFe2O4/AC possessed a photocatalytic feature for the simultaneous removal of nitrite and ammonia. A control test displayed that the concentration of nitrite almost remained constant in the absence of ammonia even

under light irradiation and that only 4.5% of nitrite was adsorbed on the catalyst in 3 h (curve c in Figure 6C). Nitrite (4.5%, 3 h) and ammonia (16.5%, 3 h) can also be adsorbed simultaneously on the ZnFe2O4/AC catalyst but were not degraded in dark as shown in curve d in Figure 6C,D. Taking curve d to compare with curve a, light irradiation promotes the reaction of nitrite with ammonia in the presence of the ZnFe2O4/AC catalyst under the lack of dissolved oxygen. Curve e in Figure 6D shows that ammonia cannot almost be oxidized in the absence of nitrite or dissolved oxygen under irradiation even if the ZnFe2O4/AC catalyst is present in the system. By analyzing the data above, it can be concluded that ZnFe2O4/AC can remove nitrite and ammonia simultaneously under anaerobic conditions upon UV−visible light irradiation. 2.6. Effect of AC Dosage on Photocatalytic Efficiency. Generally, carbon materials can markedly improve the photocatalytic process, largely through the four following mechanisms: (i) minimization of the recombination of photogenerated electron−hole pairs; (ii) modification of the band gap of the photocatalyst to longer wavelengths; (iii) adsorption that accelerate contact between the pollutant and 6414

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

catalyst; and (iv) catalysis of the carboxyl group.49 RiveraUtrilla compared the enhanced effect of ordinary AC with that of AC oxidized with H2O2, HNO3, and O3 and concluded that AC with the greatest carboxyl group content showed the highest synergistic activity, whereas the enhanced effect of ordinary AC that had not been oxidized with oxidants was mainly attributed to its high adsorption capacity.50 In the present case, a series of ZnFe2O4/AC hybrid catalysts containing various weight ratios of AC from 0.0 to 9.0% were synthesized using the abovementioned method; the resulting products were utilized as photocatalysts for the removal of nitrite and ammonia under anaerobic conditions. The removal curves are shown in Figure 7. The results show

that the removal ratios increased as the weight ratio of AC (AC: ZnFe2O4) in the composite rose at the initial stage and that they reached the summit value of 92.7% (±0.2%) for nitrite and 64.3% (±0.2%) for ammonia when the AC content was equal to 7.0%. Subsequently, contrary to what was expected, the removal ratios decreased slightly as the AC content continued to increase; the removal ratio is 90.0% (±0.2%) for nitrite and 63.0% (±0.2%) for ammonia at AC 9.0 wt %. Thus, 7.0% is the optimal weight ratio. The control test indicated that pure ZnFe2O4 (0.1 g) displayed average adsorption percentages of 0.2 and 12.0% for NN and AN in a mixed solution containing 50.0 mg/L NN and 100.0 mg/L AN at 3 h, respectively, whereas the as-prepared ZnFe2O4/AC (0.1 g) displayed adsorption ratios of 4.5 and 16.5% for NN and AN in the same mixed solution at 3 h, respectively. Compared with the adsorption ratios of NN and AN on ZnFe2O4, the adsorption ratios of NN and AN on ZnFe2O4/ AC are high, showing the concentration role of AC for NN and AN. Moreover, the photocatalytic removal difference between ZnFe2O4/AC and pure ZnFe2O4 is 12.7% for NN and 14.4% for AN. The differences of photocatalytic removal are much larger than those of adsorption. Therefore, AC improved the adsorption of NN and AN and harvesting of incident light of λ > 530 nm wavelength, resulting in the enhanced photocatalytic activity.51,52 2.7. Effect of pH on the Removal of Nitrite and Ammonia. The hydrogen ion concentration (pH) exerts great impact on both reduction of nitrite and oxidization of ammonia because their reactions involved transfer of hydrogen ions. Thus, pH influence was examined. The results showed that different pHs caused various removal ratios of both NN and AN for three runs shown in Figure 8. For the removal of nitrite, 54.7% (±0.2%) of NN removal was achieved in pH 8.5 solutions during UV−visible light irradiation for 180 min. In pH 9.0 solutions, 75.2% (±0.2%) removal ratio was achieved. When pH 9.5 solutions were tested, 92.7% (±0.2%) removal ratio was achieved, which is a higher value of degradation. On the contrary, the removal ratio will decline if pH continues to rise. The removal ratio declined to 81.0% (±0.2%) at pH 10.0, even to 40.0% (±0.2%) at pH 10.5. For the removal of ammonia, a similar event took place as shown in Figure 8B. At the initial stage, the removal ratio of ammonia went upward as pH rose until a value of 64.0% (±0.2%) was achieved at pH 9.5. After that the removal ratio started to descend, even to 36.5% (±0.2%) at pH 10.5. According to the ultraviolet−visible spectroscopic measurements, nitrogen gas was formed during the photocatalytic

Figure 7. Effects of AC content on the removal of nitrite (A) and ammonia (B) under anaerobic conditions upon UV−visible light irradiation. A 250.0 mL solution was utilized containing 50.0 mg/L NN + 100.0 mg/L AN + 1.5 g/L ZnFe2O4/AC with pH 9.5.

Figure 8. Effect of pH on the removal of nitrite (A) and ammonia (B) under anaerobic conditions upon UV−visible light irradiation. A 250.0 mL solution was utilized containing 50.0 mg/L NN + 100 mg/L AN + 1.5 g/L ZnFe2O4/AC with various pHs of 8.5, 9.0, 9.5, 10.0, and 10.5. 6415

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

process. As it will be seen below, the oxidization of NH3 will release hydrogen ions out in solution (Scheme 3), so the high Scheme 3. Photogenerated Holes on Ev Oxidize Ammonia

pH (low concentration of H+ ions) in solution will promote oxidation of NH3 in a viewpoint of chemical equilibrium. On the other hand, hydrogen ions involve in the reduction of nitrite as shown in Scheme 4. Thus, the extortionate pH value resulted in the declined removal. In this case, the optimal pH value in solution turns out to be equal to 9.5.

Figure 9. Complete removal of ammonia under aerobic conditions by means of photocatalysis.

Scheme 4. Photogenerated Electrons on Ec Reduce Nitrite

removal of nitrite and ammonia in the sealed photocatalytic reaction system mentioned previously, in which a 100 mL aqueous solution containing 50.0 mg/L NO2−−N and 100.0 mg/L NH3−N was irradiated under UV−visible light, the mixed gas of oxygen and argon was cycled, and the produced N2 and others were detected by gas chromatography. The results are displayed in Figure 10. As seen, the peak of N2 gas

In addition, it seems that the high concentration of hydrogen ions favors the overall reaction (Scheme 5). However, the Scheme 5. Overall Photocatalytic Reaction for the Simultaneous Removal of Nitrite and Ammonia

increase of hydrogen ions favors only the reduction of nitrite, but the oxidization of ammonia will get deteriorated, which is similar to ammonia oxidized photocatalytically by graphene− manganese ferrite.34 The possible reason is related to NH3 adsorbed on the catalyst. NH3, but not NH4+, is the predominant form in pH > 9.3 solutions because pKa for NH4+ is equal to 9.3. NH3 molecules containing lone pairs of electrons in sp3 hybrid orbitals for donation are more easily adsorbed on the ZnFe2O4 catalyst than NH4+ ions without lone pairs of electrons. Therefore, NH3 adsorbed on the catalyst will sacrifice the photogenerated holes so that the overall reaction continues to occur. 2.8. Complete Removal of Total Nitrogen. At the first stage, nitrite was reduced to nitrogen gas in the presence of ammonia under anaerobic conditions upon irradiation; the average removal ratio of nitrite for three runs reached to 93.3% with ±0.2% error in 3 h; simultaneously, the average removal ratio of ammonia for three runs approached to 64.5% with ±0.2% error. That is, the remaining ammonia of 35.5 mg/L in the solution needs to be degraded to achieve the complete removal of nitrogen. Thus, at the second stage, the solution containing 35.5 mg/L AN was aerated for 20 min; then, the lamp was switched on to continue the removal of the remaining ammonia. The results are shown in Figure 9. At the 12th hour during irradiation, the average removal ratios of NN, total nitrogen (TN), and AN for three runs reached to 92.0, 90.2, and 90.0% with ±0.2% error, respectively, removing most of total nitrogen. In addition, the removal ratios of NN and AN at the end of the first stage are about 6.9 and 13.2%, respectively, and the removal of AN is about 20.4% at the end of the second stage, when P25 (TiO2) was used as the photocatalyst. It turns out that ZnFe2O4/AC is very effective in the removal of NN and AN under similar conditions, compared with P25. 2.9. Identification of Products. To identify the products of nitrite reaction with ammonia during photocatalysis, measurements were performed during the photocatalytic

Figure 10. Gas chromatograms during photocatalytic simultaneous removal of nitrite and ammonia via the zinc ferrite/AC hybrid catalyst under UV−visible light irradiation at t = 0.0, 1.0, 2.0, 3.0, 5.0, 7.0, 9.0, 11.0, and 13.0 h.

was boosting with the irradiation time, whereas the peak of O2 gas in the sealed reaction system was declining with the irradiation time; meanwhile, no other gas peaks such as N2O, NO, and NO2 were observed, indicating that the product N2 gas was formed during the photocatalysis of nitrite and ammonia. 2.10. Photocatalytic Reaction Mechanism. The valence band spectrum of ZnFe2O4 was measured by X-ray photoelectron spectroscopy (XPS) technique as shown in Figure 11. The measured results indicated that the valence band of ZnFe2O4 is equal to 1.33 eV, which is very close to 1.08 eV,17 so the conduction band is equal to −0.65 eV. It is well known that photogenerated holes may oxidize water to O2 when Ev is more positve than E(O2/H2O), whereas photogenerated electrons may reduce hydrogen ions to H2 when Ec is more negative than E(H2/H2O).53 For the oxidization of NH3 (Scheme 3), the driving power ΔE1 is found to be 1.27 V (ΔE = E V − E N0 2 /NH3) because of the standard electrode potential (E N0 2 /NH3) of 0.057 V versus NHE. For the reduction of NO2− (Scheme 4), the drive power 0 − − EC) because of the standard ΔE2 is 2.17 V (ΔE = E NO 2 /N2 0 − electrode potential (E NO ) being 1.52 V versus NHE. Here, 2 /N2

6416

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

obtained from Nanjing Chemical Reagent Co., Ltd. Sodium nitrate and sodium nitrite were purchased from Tianjin Baodi Chemical Industry Co., Ltd. and Wuxi Jingke Chemical Industry Co., Ltd, respectively. All reagents were of analytical grade and utilized without further purification. All solutions were prepared with 18.2 MΩ·cm deionized Milli-Q water. 4.2. Synthesis of ZnFe2O4/AC. A one-step hydrothermal process was adopted.24,57 First, 1.7850 g of Zn(NO3)2·6H2O (6.0 mmol) and 4.8480 g of Fe(NO3)3·9H2O (12.0 mmol) were dissolved in 20.0 mL of deionized water under vigorous stirring. Second, 0.101 g of AC (after dried at 120 °C for 2 h, 7 wt % of ZnFe2O4) was dispersed in 10.0 mL of deionized water by ultrasonic dispersion for 1 h. The as-prepared solution containing Zn(II) and Fe(III) was added to 10 mL of the AC aqueous suspension; then, it was vigorously stirred for 1 h. Finally, a 10.0 mL solution containing 2.40 g of NaOH was slowly added dropwise to the above brown suspension with continuous stirring for 1.5 h. Deionized water (20.0 mL) was also added to the suspension to obtain a total volume of 60 mL. The suspension was transferred into a 100 mL Teflonlined stainless steel autoclave, which was sealed and heated at 180 °C for 8 h. Then, it was cooled naturally to room temperature, taken out, and filtered to obtain ZnFe2O4/AC precipitates. The ZnFe2O4/AC precipitates were washed thrice with deionized water to remove excess NaOH and other electrolytes. The precipitates were collected and dried at 60 °C for 24 h in a vacuum chamber for characterization and photocatalytic tests. 4.3. Structure Characterization of ZnFe2O4/AC. XRD measurements were performed using an X’Pert-Pro MPD Xray diffractometer (Panalytical, Netherlands). The X-ray source was Cu Kα radiation with a wavelength of 0.154 nm, tube voltage of 40 kV, and tube current of 40 mA. ZnFe2O4/ AC, ZnFe2O4, and AC powder were dispersed in water by using an ultrasonic device, placed on carbon-coated copper grids, and dried under ambient conditions for morphological observations using a TEM system (Tecnai G220, FEI, USA). UV−vis DRS spectra were recorded on a double-beam TU1901 spectrophotometer (type T1901, Puxi Co., Beijing, China). Raman spectra were recorded on an HR800 microconfocal Raman spectrometer with a 633 nm laser. 4.4. Simultaneous Removal of Nitrite and Ammonia. The experiments for the removal of nitrite and ammonia were performed using a commercial glass photoreactor (Shanghai Binlong Instrument Co. Ltd.) as shown in Figure 13. A highpressure mercury lamp (300 W), surrounded by a flow of water to maintain 25 °C (the temperature in the reaction solution is

Figure 11. XPS valence band spectrum of the ZnFe2O4 sample.

the oxidization of NH3 to N2 by photogenerated holes is similar to the oxidization of H2O to O2 by photogenerated holes,54,55 whereas the reduction of NO2− to N2 by photogenerated electrons is similar to the reduction of H+ ions to H2.56 Reactions 3 and 4 can both occur spontaneously in energy because both ΔE1 and ΔE2 are more than zero. Thus, the overall reaction is denoted in Scheme 5; the overall photocatalytic process is presented in Figure 12.

Figure 12. Reduction of nitrite and oxidization of ammonia to release nitrogen gas by means of photocatalysis.

3. CONCLUSIONS Simultaneous removal of nitrite and ammonia was achieved successfully based on the zinc ferrite/AC catalyst via two steps under UV−visible light irradiation. The detection of products by means of gas chromatography indicated that nitrogen gas was released out during photocatalysis. The XPS measurements and analysis of the band structure for ZnFe2O4 showed that Ec (−0.65 V vs NHE) of ZnFe2O4 is lower than the reduced potential (1.52 V vs NHE) of nitrite and thereby the photogenerated electrons can reduce nitrite to N2 gas under irradiation; while Ev (1.33 V vs NHE) of ZnFe2O4 is higher than the potential (0.057 V vs NHE) of ammonia, as a result, the photogenerated holes can oxidize ammonia to N2 gas. AC caused the red shift of the band gap, separation of the photogenerated electron−hole pairs, and adsorption for nitrite and ammonia, promoting the photocatalytic activity. 4. EXPERIMENTAL SECTION 4.1. Chemicals. AC and sodium hydroxide (NaOH) were purchased from Sigma-Aldrich LLC (China). Ferric nitrate nonahydrate (Fe(NO3)3·9H2O) was bought from Tianjin Damao Chemical Factory, China. Zinc nitrate hexahydrate (Zn(NO3)2·6H2O) and ammonium chloride (NH4Cl) were

Figure 13. Experimental apparatus for the simultaneous removal of nitrite and ammonia. 6417

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

28 °C as measured by the thermometer), was inserted in a 250 mL suspension solution containing NN (50 mg/L), AN (100 mg/L), and 1.5 g/L ZnFe2O4/AC catalyst. The pH of the suspension was adjusted to 9.5 using 1.0 mol/L NaOH solution. The suspension was bubbled with nitrogen gas at a flow rate of 0.3 L min−1 for 30 min to remove dissolved oxygen for photocatalytic tests under magnetic stirring. Solution (3 mL) was taken out to determine the concentrations of nitrite, ammonia, or nitrate at the interval of 0.5 h. To achieve complete removal of nitrogen, after nitrite was reduced by ammonia photocatalytically, NaOH was added to the suspension solution to maintain pH 9.5 and oxygen gas was also pumped into the suspension solution to provide with dissolved oxygen for degrading the residual ammonia to achieve the complete removal of total nitrogen. 4.5. Measurements of Nitrite, Nitrate, and Ammonia. A double-beam TU-1901 spectrophotometer was used to probe the concentrations of nitrite and ammonia by using Griess reagents58,59 and Nessler reagents60 during photocatalysis, respectively. Griess reagents contain sulphanilamide and N-(1-naphthyl)ethylenediamine as the coupling agent. Five grams (5.0 g) of sulphanilamide was added to a 400 mL solution containing 36.5 wt % HCl of 50 mL; the solution was diluted to 500 mL. Then, 100.0 mg of N-(1-naphthyl)ethylenediamine was dissolved in 50 mL of deionized water; it was diluted to 100 mL. The measurement procedures are as follows: one milliliter (1.0 mL) of the sample solution containing nitrite was taken out to a 50 mL colorimetric cylinder. Then, 1.0 mL of sulphanilamide solution was added to the nitrite solution for 5 min; after that 1.0 mL of N-(1naphthyl)ethylenediamine solution was also added the solution to form a red azo dye with an absorption maximum at 540 nm. Finally, it was diluted to 50 mL for the measurement at 540 nm. The reactions are presented in Scheme 1. Ammonia reacts with Nessler reagent to give colored solutions as shown in Scheme 2. The absorbance at 382 nm approached the peak and was recorded at the wavelength of 382 nm.48 Nitrate was determined directly at 203 nm using ultraviolet visible spectroscopy.61 The removal ratios of NN, AN, and TN were calculated using eqs 3−5 NN % =

0 C NN − C NN t 0 C NN

photocatalysis was conducted under 300 W Xe-lamp irradiation for 2 h. After that a mixed gas with the initial ratio of oxygen to argon at 1:2.6 was pumped until the internal pressure reached to 80 kPa and then the second stage of photocatalysis continued to occur. The reaction system was coupled with a GC-7806 gas chromatography (Shiwei Puxin Instruments Co. Ltd., Beijing, China) system. A chromatographic column of 5 m length × 2 mm i.d., filled with a 5 Å molecular sieve was utilized as a separation column. Highpurity argon was utilized as the carrier gas. The flow rate of the carrier gas was set at 23 mL/min. A thermal conductivity tank was used to detect N2, O2, and others. The column temperature was set at 80 °C, inlet temperature at 120 °C, and detector temperature at 150 °C. 4.7. First-Principles Calculation. The calculations of band structure and DOS were performed by using the CASTEP package of Materials Studio 7.0. In the plane-wave calculations, a cutoff energy of 520 eV was applied. The generalized gradient approximation was adopted with the Perdew−Burke−Ernzerhof functional. The calculations were performed in a ferromagnetic spin-polarized configuration. The self-consistent field convergence criterion and energy tolerance were set as fine levels. A Monkhorst−Pack scheme with a 6 × 6 × 6 k-point grid was employed. Tkatchenko−Scheffler method was used for DFT-D correction. The ZnFe2O4 phase is of the cubic structure with space group Fd3̅m (no. 227) and the lattice parameters a = b = c = 8.4418 Å, and α = β = γ = 90°, in which the Zn, Fe, and O atoms occupy the fractional coordinates of Zn1 (0.00, 0.00, 0.00) with site occupation factor (SOF) 0.89, Fe1 (0.00, 0.00, 0.00) with SOF 0.11, Fe2 (0.625, 0.625, 0.625) with SOF 0.945, Zn2 (0.625, 0.625, 0.625) with SOF 0.055, and O (0.3854, 0.3854, 0.3854), respectively. These accurate data validated the applicability of the CASTEP package for calculating the band structures and DOS of ZnFe2O4.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b00677.

× 100%

ij C 0 − C t yz AN % = jjj AN 0 AN zzz × 100% j z CAN k {

ij C t + CAN t yzz z × 100% TN % = jjj1 − NN 0 0 z j z + CAN C NN k {

Three partial DOS related to iron, oxygen, and zinc atoms, calibration curve of nitrate, determination of nitrate, and stability and reuse of the ZnFe2O4/AC catalyst (PDF)

(3)



(4)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected], shouqing_liu@hotmail. com (S.-Q.L.). *E-mail: [email protected]. Phone: +86 51268415070 (L.L.).

(5)

where C0NN is the initial concentration of NN, C0AN is the initial concentration of AN, CNNt is the concentration of NN at t min during the photocatalytic process, and CANt is the concentration of AN at t min during the photocatalytic process. 4.6. Identification of Products. To detect the products of nitrite reaction with ammonia during photocatalysis, a sealed photocatalytic reaction system (Labsolar 6A photocatalytic system, Perfectlight Co. Ltd., Beijing, China) was applied, in which a 100 mL aqueous solution containing 50.0 mg/L NO2−−N + 100.0 mg/L NH3−N was irradiated under UV− visible light irradiation after the 465 mL reaction system was evacuated to 0.4 kPa. Subsequently, the first stage of

ORCID

Shou-Qing Liu: 0000-0002-3774-4180 Li Luo: 0000-0001-5296-7000 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work is financially supported by the National Natural Science Foundation of China (nos. 21576175, 51502187), the 6418

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

(15) Liu, G.; You, S.; Ma, M.; Huang, H.; Ren, N. Removal of nitrate by photocatalytic denitrification using nonlinear optical material. Environ. Sci. Technol. 2016, 50, 11218−11225. (16) Oka, M.; Miseki, Y.; Saito, K.; Kudo, A. Photocatalytic reduction of nitrate ions to dinitrogen over layered perovskite BaLa4Ti4O15, using water as an electron donor. Appl. Catal., B 2015, 179, 407−411. (17) Liu, S.-Q.; Zhu, X.-L.; Zhou, Y.; Meng, Z.-D.; Chen, Z.-G.; Liu, C.-B.; Chen, F.; Wu, Z.-Y.; Qian, J.-C. Smart photocatalytic removal of ammonia through molecular recognition of zinc ferrite/reduced graphene oxide hybrid catalyst under visible-light irradiation. Catal. Sci. Technol. 2017, 7, 3210−3219. (18) Zhao, W.; Liang, C.; Wang, B.; Xing, S. Enhanced Photocatalytic and Fenton-like Performance of CuOx-Decorated ZnFe2O4. ACS Appl. Mater. Interfaces 2017, 9, 41927−41936. (19) Fu, Y.; Wang, X. Magnetically separable ZnFe2O4-graphene catalyst and its high photocatalytic performance under visible light irradiation. Ind. Eng. Chem. Res. 2011, 50, 7210−7218. (20) Yao, Y.; Cai, Y.; Lu, F.; Qin, J.; Wei, F.; Xu, C.; Wang, S. Magnetic ZnFe2O4-C3N4 Hybrid for photocatalytic degradation of aqueous organic pollutants by visible light. Ind. Eng. Chem. Res. 2014, 53, 17294−17302. (21) Nada, A. A.; Nasr, M.; Viter, R.; Miele, P.; Roualdes, S.; Bechelany, M. Mesoporous ZnFe2O4@TiO2 nanofibers prepared by electrospinning coupled to PECVD as highly performing photocatalytic materials. J. Phys. Chem. C 2017, 121, 24669−24677. (22) Chen, X.; Dai, Y.; Guo, J.; Liu, T.; Wang, X. Novel magnetically separable reduced graphene oxide (rGO)/ZnFe2O4/Ag3PO4 nanocomposites for enhanced photocatalytic performance toward 2,4dichlorophenol under visible light. Ind. Eng. Chem. Res. 2016, 55, 568−578. (23) Abazari, R.; Mahjoub, A. R. Potential Applications of Magnetic β-AgVO3/ZnFe2O4 Nanocomposites in Dyes, Photocatalytic Degradation, and Catalytic Thermal Decomposition of Ammonium Perchlorate. Ind. Eng. Chem. Res. 2017, 56, 623−634. (24) Huang, Y.; Liang, Y.; Rao, Y.; Zhu, D.; Cao, J.-j.; Shen, Z.; Ho, W.; Lee, S. C. Environment-friendly carbon quantum dots/ZnFe2O4 photocatalysts: characterization, biocompatibility, and mechanisms for NO removal. Environ. Sci. Technol. 2017, 51, 2924−2933. (25) Zhang, F.; Li, X.; Zhao, Q.; Zhang, D. Rational design of ZnFe2O4/In2O3 nanoheterostructures: efficient photocatalyst for gaseous 1,2-dichlorobenzene degradation and mechanistic insight. ACS Sustainable Chem. Eng. 2016, 4, 4554−4562. (26) Guo, J.; Wang, K.; Wang, X. Photocatalytic reduction of CO2 with H2 O vapor under visible light over Ce doped ZnFe2O4. Catal. Sci. Technol. 2017, 7, 6013−6025. (27) Li, Y.; Shen, J.; Hu, Y.; Qiu, S.; Min, G.; Song, Z.; Sun, Z.; Li, C. General flame approach to chainlike MFe2O4 Spinel (M = Cu, Ni, Co, Zn) nanoaggregates for reduction of nitroaromatic compounds. Ind. Eng. Chem. Res. 2015, 54, 9750−9757. (28) Lv, H.; Ma, L.; Zeng, P.; Ke, D.; Peng, T. Synthesis of floriated ZnFe2O4 with porous nanorod structures and its photocatalytic hydrogen production under visible light. J. Mater. Chem. 2010, 20, 3665−3672. (29) Peeters, D.; Taffa, D. H.; Kerrigan, M. M.; Ney, A.; Jöns, N.; Rogalla, D.; Cwik, S.; Becker, H.-W.; Grafen, M.; Ostendorf, A.; Winter, C. H.; Chakraborty, S.; Wark, M.; Devi, A. Photoactive zinc ferrites fabricated via conventional CVD approach. ACS Sustainable Chem. Eng. 2017, 5, 2917−2926. (30) Dom, R.; Subasri, R.; Hebalkar, N. Y.; Chary, A. S.; Borse, P. H. Synthesis of a hydrogen producing nanocrystalline ZnFe2O4 visible light photocatalyst using a rapid microwave irradiation method. RSC Adv. 2012, 2, 12782−12791. (31) Adamu, H.; McCue, A. J.; Taylor, R. S. F.; Manyar, H. G.; Anderson, J. A. Simultaneous photocatalytic removal of nitrate and oxalic acid over Cu2O/TiO2 and Cu2O/TiO2-AC composites. Appl. Catal., B 2017, 217, 181−191. (32) Tian, M.-J.; Liao, F.; Ke, Q.-F.; Guo, Y.-J.; Guo, Y.-P. Synergetic effect of titanium dioxide ultralong nanofibers and

key industrial prospective program of Jiangsu Science and Technology Department (BE2015190), the Natural Science Foundation of Jiangsu Province of China (no. BK20141178), the Earmarked Nanotechnology Fund of the Bureau of Science and Technology of Suzhou City (no. ZXG201429), the Creative Project of Postgraduate of Suzhou University of Science and Technology (no. SKCX16_066), and Collaborative Innovation Center of Technology and Material of Water Treatment. The authors also appreciate support from the National Supercomputing Center in Shenzhen (Shenzhen Cloud Computing Center) for first-principles calculation using the CASTEP package of Materials Studio.



REFERENCES

(1) Sá, J.; Agüera, C. A.; Gross, S.; Anderson, J. A. Photocatalytic nitrate reduction over metal modified TiO2. Appl. Catal., B 2009, 85, 192−200. (2) Pandikumar, A.; Manonmani, S.; Ramaraj, R. TiO2−Au nanocomposite materials embedded in polymer matrices and their application in the photocatalytic reduction of nitrite to ammonia. Catal. Sci. Technol. 2012, 2, 345−353. (3) World Health Organization. Guidelines for Drinking-Water Quality, 4th ed., 2011. (4) Wang, Q.; Duan, H.; Wei, W.; Ni, B.-J.; Laloo, A.; Yuan, Z. Achieving stable mainstream nitrogen removal via the nitrite pathway by sludge treatment using free ammonia. Environ. Sci. Technol. 2017, 51, 9800−9807. (5) De Clippeleir, H.; Courtens, E.; Mosquera, M.; Vlaeminck, S. E.; Smets, B. F.; Boon, N.; Verstraete, W. Efficient total nitrogen removal in an ammonia gas biofilter through high-rate OLAND. Environ. Sci. Technol. 2012, 46, 8826−8833. (6) Vlaeminck, S. E.; Terada, A.; Smets, B. F.; Linden, D. V. D.; Verstraete, W.; Carballa, M. Nitrogen removal from digested black water by one-stage partial nitritation and anammox. Environ. Sci. Technol. 2009, 43, 5035−5041. (7) Aponte-Morales, V. E.; Payne, K. A.; Cunningham, J. A.; Ergas, S. J. Bioregeneration of chabazite during nitrification of centrate from anaerobically digested livestock waste: Experimental and modeling studies. Environ. Sci. Technol. 2018, 52, 4090−4098. (8) Chen, H.; Zhao, X.; Cheng, Y.; Jiang, M.; Li, X.; Xue, G. Iron robustly stimulates simultaneous nitrification and denitrification under aerobic conditions. Environ. Sci. Technol. 2018, 52, 1404−1412. (9) Zhu, X.; Castleberry, S. R.; Nanny, M. A.; Butler, E. C. Effects of pH and catalyst concentration on photocatalytic oxidation of aqueous ammonia and nitrite in titanium dioxide suspensions. Environ. Sci. Technol. 2005, 39, 3784−3791. (10) Zhang, F.; Pi, Y.; Cui, J.; Yang, Y.; Zhang, X.; Guan, N. Unexpected selective photocatalytic reduction of nitrite to nitrogen on silver-doped titanium dioxide. J. Phys. Chem. C 2007, 111, 3756− 3761. (11) Gekko, H.; Hashimoto, K.; Kominami, H. Photocatalytic reduction of nitrite to dinitrogen in aqueous suspensions of metalloaded titanium(iv) oxide in the presence of a hole scavenger: an ensemble effect of silver and palladium co-catalysts. Phys. Chem. Chem. Phys. 2012, 14, 7965−7970. (12) Kominami, H.; Kitsui, K.; Ishiyama, Y.; Hashimoto, K. Simultaneous removal of nitrite and ammonia as dinitrogen in aqueous suspensions of a titanium(iv) oxide photocatalyst under reagent-free and metal-free conditions at room temperature. RSC Adv. 2014, 4, 51576−51579. (13) Yue, M.; Wang, R.; Ma, B.; Cong, R.; Gao, W.; Yang, T. Superior performance of CuInS2 for photocatalytic water treatment: full conversion of highly stable nitrate ions into harmless N2 under visible light. Catal. Sci. Technol. 2016, 6, 8300−8308. (14) Kato, H.; Kudo, A. Photocatalytic reduction of nitrate ions over tantalate photocatalysts. Phys. Chem. Chem. Phys. 2002, 4, 2833− 2838. 6419

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420

ACS Omega

Article

activated carbon in the photocatalytic degradation of cytarabine. Appl. Catal., B 2011, 104, 177−184. (51) Hazarika, D.; Karak, N. Unprecedented influence of carbon dot@TiO2 nanohybrid on multifaceted attributes of waterborne hyperbranched polyester nanocomposite. ACS Omega 2018, 3, 1757− 1769. (52) Kumar, S.; Sahoo, P. K.; Satpati, A. K. Electrochemical and SECM investigation of MoS2/GO and MoS2/rGO nanocomposite materials for HER electrocatalysis. ACS Omega 2017, 2, 7532−7545. (53) Kato, H.; Kudo, A. Photocatalytic water splitting into H2 and O2 over various tantalate photocatalysts. Catal. Today 2003, 78, 561− 569. (54) Chattopadhyay, S.; Chen, L.-C.; Chen, K.-H. Energy production and conversion applications of one-dimensional semiconductor nanostructures. NPG Asia Mater. 2011, 3, 74−81. (55) Hu, S.; Shaner, M. R.; Beardslee, J. A.; Lichterman, M.; Brunschwig, B. S.; Lewis, N. S. Amorphous TiO2 coatings stabilize Si, GaAs, and GaP photoanodes for efficient water oxidation. Science 2014, 344, 1005−1009. (56) Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.-T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science 2015, 347, 970−974. (57) Feng, J.; Hou, Y.; Wang, Y.; Li, L. Synthesis of Hierarchical ZnFe2O4@SiO2@RGO core-shell microspheres for enhanced electromagnetic wave absorption. ACS Appl. Mater. Interfaces 2017, 9, 14103−14111. (58) Wang, B.; Lin, Z.; Wang, M. Fabrication of a paper-based microfluidic device to readily determine nitrite ion concentration by simple colorimetric assay. J. Chem. Educ. 2015, 92, 733−736. (59) Norwitz, G.; Keliher, P. N. Spectrophotometric determination of nitrite with composite reagents containing sulphanilamide, sulphanilic acid or 4-nitroaniline as the diazotisable aromatic amine and N-(1-naphthyl)ethylenediamine as the coupling agent. Analyst 1984, 109, 1281−1286. (60) Zhou, Y.; Xiao, B.; Liu, S.-Q.; Meng, Z.; Chen, Z.-G.; Zou, C.Y.; Liu, C.-B.; Chen, F.; Zhou, X. Photo-Fenton degradation of ammonia via a manganese-iron double-active component catalyst of graphene-manganese ferrite under visible light. Chem. Eng. J. 2016, 283, 266−275. (61) Miles, D. L.; Espejo, C. Comparison between an ultraviolet spectrophotometric procedure and the 2,4-xylenol method for the determination of nitrate in groundwaters of low salinity. Analyst 1977, 102, 104−109.

activated carbon fibers on adsorption and photodegradation of toluene. Chem. Eng. J. 2017, 328, 962−976. (33) Hakamizadeh, M.; Afshar, S.; Tadjarodi, A.; Khajavian, R.; Fadaie, M. R.; Bozorgi, B. Improving hydrogen production via water splitting over Pt/TiO2/activated carbon nanocomposite. Int. J. Hydrogen Energy 2014, 39, 7262−7269. (34) Basha, S.; Keane, D.; Morrissey, A.; Nolan, K.; Oelgemöller, M.; Tobin, J. Studies on the adsorption and kinetics of photodegradation of pharmaceutical compound, indomethacin using novel photocatalytic adsorbents (IPCAs). Ind. Eng. Chem. Res. 2010, 49, 11302−11309. (35) Chen, X.; Kuo, D.-H.; Lu, D. Nanonization of g-C3N4 with the assistance of activated carbon for improved visible light photocatalysis. RSC Adv. 2016, 6, 66814−66821. (36) Hu, X.-W.; Liu, S.; Qu, B.-T.; You, X.-Z. Starfish-shaped Co3O4/ZnFe2O4 hollow nanocomposite: synthesis, supercapacity, and magnetic properties. ACS Appl. Mater. Interfaces 2015, 7, 9972−9981. (37) Jiang, B.; Han, C.; Li, B.; He, Y.; Lin, Z. In-situ crafting of ZnFe2O4 nanoparticles impregnated within continuous carbon network as advanced anode materials. ACS Nano 2016, 10, 2728− 2735. (38) Fu, Y.; Wang, X. Magnetically separable ZnFe2O4 graphene catalyst and its high photocatalytic performance under visible light irradiation. Ind. Eng. Chem. Res. 2011, 50, 7210−7218. (39) Zhang, F.; Li, X.; Zhao, Q.; Zhang, D. Rational design of ZnFe2O4/In2O3 nanoheterostructures: Efficient photocatalyst for gaseous 1,2-dichlorobenzene degradation and mechanistic insight. ACS Sustainable Chem. Eng. 2016, 4, 4554−4562. (40) Xiong, P.; Zhu, J.; Wang, X. Cadmium Sulfide-Ferrite Nanocomposite as a Magnetically Recyclable Photocatalyst with Enhanced Visible-Light-Driven Photocatalytic Activity and Photostability. Ind. Eng. Chem. Res. 2013, 52, 17126−17133. (41) Maletin, M.; Moshopoulou, E. G.; Kontos, A. G.; Devlin, E.; Delimitis, A.; Zaspalis, V. T.; Nalbandian, L.; Srdic, V. V. Synthesis and structural characterization of In-doped ZnFe2O4, nanoparticles. J. Eur. Ceram. Soc. 2007, 27, 4391−4394. (42) Wang, Z.; Lazor, P.; Saxena, S. K.; Artioli, G. High-pressure Raman spectroscopic study of spinel (ZnCr2O4). J. Solid State Chem. 2002, 165, 165−170. (43) Wang, Z.; Schiferl, D.; Zhao, Y.; O’Neill, H. S. C. High pressure Raman spectroscopy of spinel-type ferrite ZnFe2O4. J. Phys. Chem. Solids 2003, 64, 2517−2523. (44) Bresser, D.; Paillard, E.; Kloepsch, R.; Krueger, S.; Fiedler, M.; Schmitz, R.; Baither, D.; Winter, M.; Passerini, S. Carbon coated ZnFe2O4, nanoparticles for advanced lithium-ion anodes. Adv. Energy Mater. 2013, 3, 513−523. (45) Xu, J.; Jeon, I.-Y.; Seo, J.-M.; Dou, S.; Dai, L.; Baek, J.-B. Edgeselectively halogenated graphene nanoplatelets (XgNps, X = Cl, Br, or I) prepared by ball-milling and used as anode materials for lithium-ion batteries. Adv. Mater. 2014, 26, 7317−7323. (46) Segets, D.; Lucas, J. M.; Klupp Taylor, R. N.; Scheele, M.; Zheng, H.; Alivisatos, A. P.; Peukert, W. Determination of the quantum dot band gap dependence on particle size from optical absorbance and transmission electron microscopy measurements. ACS Nano 2012, 6, 9021−9032. (47) Jacobsson, T. J.; Edvinsson, T. Photoelectrochemical determination of the absolute band edge positions as a function of particle size for ZnO quantum dots. J. Phys. Chem. C 2012, 116, 15692−15701. (48) Liu, W.-X.; Zhu, X.-L.; Liu, S.-Q.; Gu, Q.-Q.; Meng, Z.-D. Near-infrared-driven selective photocatalytic removal of ammonia based on valence band recognition of an α-MnO2/N-doped graphene hybrid catalyst. ACS Omega 2018, 3, 5537−5546. (49) Rivera-Utrilla, J.; Sánchez-Polo, M.; Abdel daiem, M. M.; Ocampo-Pérez, R. Role of activated carbon in the photocatalytic degradation of 2,4-dichlorophenoxyacetic acid by the UV/TiO2/ activated carbon system. Appl. Catal., B 2012, 126, 100−107. (50) Ocampo-Pérez, R.; Sánchez-Polo, M.; Rivera-Utrilla, J.; LeyvaRamos, R. Enhancement of the catalytic activity of TiO2 by using 6420

DOI: 10.1021/acsomega.8b00677 ACS Omega 2019, 4, 6411−6420