Near-Infrared-Driven Selective Photocatalytic Removal of Ammonia

47 mins ago - ACS Earth and Space Chemistry · ACS Energy Letters · ACS Infectious .... Near-Infrared-Driven Selective Photocatalytic Removal of Ammoni...
0 downloads 0 Views 3MB 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 2018, 3, 5537−5546

Near-Infrared-Driven Selective Photocatalytic Removal of Ammonia Based on Valence Band Recognition of an α‑MnO2/N-Doped Graphene Hybrid Catalyst Wen-Xiao Liu,† Xiao-Lei Zhu,† Shou-Qing Liu,* Qin-Qin Gu, and Ze-Da Meng Jiangsu Key Laboratory of Environmental Functional Materials, School of Chemistry, Biology and Material Engineering, Suzhou University of Science and Technology, Suzhou 215009, China S Supporting Information *

ABSTRACT: Near-infrared (NIR)-response photocatalysts are desired to make use of 44% NIR solar irradiation. A flower-like α-MnO2/N-doped graphene (NG) hybrid catalyst was synthesized and characterized by X-ray diffraction spectroscopy, transmission electron microscopy, Raman spectroscopy, UV−vis−NIR diffuse reflectance spectroscopy, and X-ray photoelectron spectroscopy. The flower-like material of α-MnO2/NG was oval-shaped with the semi major axis of 140 nm and semi minor axis of 95 nm and the petal thickness of 3.5−8.0 nm. The indirect band gap was measured to be 1.16 eV, which is very close to 0.909 eV estimated by the first-principles calculation. The band gap can harvest NIR irradiation to 1069 nm. The coupling of α-MnO2 with NG sheets to form α-MnO2/NG can significantly extend the spectrum response up to 1722 nm, improving dramatically the photocatalytic activity. The experimental results displayed that the α-MnO2/NG hybrid catalyst can recognize ammonia in methyl orange (MO)−ammonia, rhodamine B (RHB)−ammonia, and humic acid−ammonia mixed solutions and selectively degrade ammonia. The degradation ratio of ammonia reached over 93.0% upon NIR light irradiation in the mixed solutions, while those of MO, RHB, and humic acid were only 9.7, 9.4, and 15.7%, respectively. The products formed during the photocatalytic process were followed with ion chromatography, gas chromatography, and electrochemistry. The formed nitrogen gas has been identified during the photocatalytic process. A valence band recognition model was suggested based on the selective degradation of ammonia via α-MnO2/NG. particles.16 Obviously, the band gap of α-MnO2 can been tailored by the size of the material to fulfill the goal of wide spectrum absorption and full use of solar irradiation. In addition, the surface modification such as copper phosphide decorated on g-C3 N 4 nanosheets,19 Ag on Bi4Ti3O12 and CuBi2O4, improved markedly the photocatalytic activity.20,21 CdSe decorated N-doped TiO2 to boom the generation of H2.22 N-doped graphene (NG) can enhance the photocatalytic activity of photocatalysts because of the absorption of incident light, fast electron transportation from semiconductors to NG, and the reduced recombination of the photogenerated electron−hole pairs23,24 when it was coupled with semiconductors. Therefore, the as-coupled product, αMnO2/NG, was expected to be more active to NIR irradiation. Selective photocatalysis is a very interesting research from organic synthesis to environmental applications.25−30 It can synthesize specific target organic molecules or degrade specific target pollutants. These selective photocatalytic reactions are based on the interaction forces such as van der Waals and coordination between hosts and guests. However, the potential

1. INTRODUCTION Near-infrared (NIR) light (44%) and visible light (48%) account for 92% of solar irradiation spectrum. Therefore, it is urgently desired to develop NIR-response photocatalysts to make full use of solar spectrum. Efforts have been paid to find out NIR-response photocatalysts, in which up-conversion materials composed of rare earths,1−6 carbon quantum dots,7−10 and plasmon composites composed of Au and Ag6,11−13 have been synthesized. The efforts inspired scientists to develop novel NIR-response photocatalysts for thorough usefulness of solar irradiation. It is known that α-MnO2 is a semiconductor material with various band gaps with the size of particles or sheet thickness. The band gap depends on the size and structure of this material. The bulk MnO2 materials are usually regarded as indirect semiconductors with very small band-gap energies of about 0.20−0.37 eV,14,15 while an ultralong nanowire α-MnO2 with the length of more than 10 mm and width of 30 nm produced a band gap of 0.98 eV.16 One of α-MnO2 nanofibers with typical diameters of 20−60 nm and lengths of 1−6 μm formed a band gap of 1.32 eV;17 another ultrathin MnO2 of nanosheets with the thickness of about 0.5 nm achieved a band gap of about 2.23 eV.18 The shift of band gap to higher energies is attributed to the carrier confinement in small semiconductor © 2018 American Chemical Society

Received: January 25, 2018 Accepted: April 9, 2018 Published: May 23, 2018 5537

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega

2.4. Structure Characterization and Mechanism Investigation. X-ray diffraction (XRD) measurements were performed using an X-ray diffractometer (D/Max 2500 PC). The X-ray source was Cu Kα radiation with a wavelength of 0.154 nm, a tube voltage of 40 kV, and a tube current of 40 mA. α-MnO2/NG and α-MnO2/rGO powders were dispersed in water using an ultrasonic device, placed on carbon-coated copper grids, and dried under ambient conditions for morphological observations using a transmission electron microscopy (TEM) system (Tecnai G220, FEI, USA). UV− vis−NIR diffuse reflectance spectra (UV−vis−NIR DRS) were recorded on a Shimadzu UV−vis−NIR spectrometer (UV-3600 plus, Japan), and the wavelength scope is 190−3400 nm. An Xray photoelectron spectrometer (ESCALAB 50 XI) was utilized to characterize ammonia adsorbed on α-MnO2 samples and chemical surrounding of Mn in α-MnO2 samples before and after ammonia was adsorbed to elucidate the molecular recognition mechanism. The binding energy (Eb) scale was calibrated with respect to the carbon 1s peak at 284.60 eV. The aluminum Kα 1.2 line (hν 1486.60 eV) was used as the X-ray excitation source, and each powder sample was dispersed on a gold-plated copper surface. The spectra were recorded in a fixed analyzer transmission mode to achieve maximum instrumental resolution. The instrument was operated under a vacuum of 1 × 10−9 Torr in the analysis chamber. Wide and high-resolution spectra were recorded at a constant pass energy of 50 eV and channel widths of 1.0 and 0.1 eV. The α-MnO2 samples were collected before and after adsorbing ammonia to detect the Eb of N and Mn atoms on the catalyst. Cyclic voltammetry was performed using a CHI660C electrochemical workstation (CH Instrument Company, TX, USA). The electrochemical experiments were conducted in a conventional three-electrode system. A ring-disk platinum electrode was used as the working electrode, a platinum foil as the counter electrode, and a saturated calomel electrode as the reference electrode. In order to obtain the stable cyclic voltammograms of oxygen dissolved, the working electrode was scanned between 0.0 and −1.4 V for 10 cycles before the measurement. The voltammetric measurements were carried out in a sealed container to prevent O2 from air. A Dionex (Dionex, Sunnyvale, CA, USA) model ICS-900 ion chromatograph was used for separation and detection of nitrite and nitrate. The analytical column was Dionex IonPac AS14A (250 mm × 4 mm, i.d.), and the guard column was IonPac AG14A (50 mm × 4 mm, i.d.). The eluent was a mixture of 4.5 mmol/L sodium carbonate and 0.8 mmol/L sodium bicarbonate, the eluent flow rate was set at 1.2 mL/min, the column temperature was 25 °C, and the injection volume was 2 mL. A conductivity detector CD produced by the company Dionex (USA) and conductivity suppressor ASRS-300 4 mm (Dionex, USA) were used for the detection of anions. In order to investigate the mechanism of photocatalysis, a sealed photocatalytic reaction system (Labsolar 6A photocatalytic system, Perfect light Co. Ltd., Beijing, China) was used, in which a 100 mL aqueous solution containing 100.0 mg/L NH3−N was irradiated under NIR light and a 465 mL gas with an initial ratio of oxygen to argon at 1:3 was cycled and the initial internal pressure was maintained at 26 kPa. The sealed system was vacuumed before the mixed gas of oxygen and argon was injected. The reaction system was matched with a GC-7806 gas chromatograph (Shiwei Puxin Instruments Co. Ltd., Beijing, China). The chromatographic column of 5 m length × 2 mm i.d., filled with 5 Å molecular sieve, was utilized

recognition based on the valence band (Ev) level has not been reported. Herein, we reported the NIR-driven selective degradation of ammonia via the α-MnO2/NG nanophotocatalyst, and a valence band recognition model was achieved. It is expected to highlight NIR photocatalysis.

2. EXPERIMENTAL AND CALCULATIONS 2.1. Chemicals. Potassium permanganate (KMnO4) was purchased from Shanghai General Chemical Reagent Plant, and manganese sulfate monohydrate (MnSO4·H2O) was purchased from Sinopharm Chemical Reagent Co., Ltd. Graphite powder (average particle size = 30 μm) was purchased from Shanghai Colloid Chemical Plant, and sodium hydroxide (NaOH) was purchased from Tianjin Damao Chemical Factory. Ammonium sulfate ((NH4)2SO4), ammonium chloride (NH4Cl), rhodamine B (RHB), and methyl orange (MO) were obtained from Nanjing Chemical Reagent Co., Ltd. All reagents were of analytical grade and used without further purification. All solutions were prepared with 18.2 MΩ·cm deionized Milli-Q water. 2.2. Synthesis of NG. NG sheets were prepared via a onepot hydrothermal reaction of graphene oxide (GO) with urea at 170 °C for 12 h.31 GO was obtained by a modified Hummers method as we reported previously.32,33 The general procedures are as follows. First, natural graphite powders were oxidized with concentrated KMnO4 in the presence of H2SO4; then, the excess of KMnO4 was reduced with H2O2 and exfoliated by sonication. After GO was obtained, nitrogen insertion into graphene frameworks using urea was achieved by a facile hydrothermal reaction at 170 °C for 12 h. The nitrogen content could be controlled by the mass ratio of GO to urea. A typical mass ratio (GO/urea) used in our case was 1:300 to obtain an optimal nitrogen content. Eighty milligrams (80 mg) of GO was dispersed into 80 mL of water with sonication, and then 24.0 g of urea was introduced in the 80 mL suspension solution for continuing sonication for 2 h. The mixture was transferred to a Teflon-lined stainless steel autoclave with a volume of 100 mL, sealed, and heated to 170 °C for 12 h to form NG (nitrogen content: ca. 7%).31 Finally, the resulting product was filtered, washed, and dried at 60 °C in a vacuum chamber. 2.3. Synthesis of α-MnO2/NG. The α-MnO2/NG catalyst with a loading of 1.0 wt % NG was synthesized as follows. First, 6 mg of NG was dispersed into 20 mL of water with sonication, and 0.2672 g (1.35 mmol) of MnCl2·4H2O was added into the suspension solution under sonication for 6 h. Then, the suspension solution was transferred into a 50 mL three-necked flask, and then it was heated to 85 °C under magnetic stirring. Finally, 0.1422 g (0.9 mmol) of KMnO4 was dissolved in 5 mL of deionized water, and the KMnO4 solution was titrated into a three-necked flask at 85 °C under continuous stirring. This temperature was maintained for 12 h to form α-MnO2/NG. After that, it was cooled to room temperature, and the precipitates were collected, washed at least for three times, and dried at 60 °C in a vacuum chamber for 8 h. The resulting product was denoted as α-MnO2/NG. Similarly, α-MnO2/ reduced graphene oxide (rGO) and pure α-MnO2 samples were synthesized for comparison. The reaction formula was denoted as follows: 2KMnO4 + 3MnCl 2 + 2H 2O → 5MnO2 + 2KCl + 4HCl

(1) 5538

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega

0.17, 0.00), O1 (0.16, 0.205, 0.00), and O2 (0.16, 0.458, 0.00). These accurate data validated the applicability of the CASTEP package for calculating the α-MnO2 structures.

as a separating column. High-purity argon was utilized as the carrier gas. The flow rate of the carrier gas was set at 23 mL/ min, and the thermal conductivity tank was used to detect N2 and O2. The column temperature was set at 80 °C, the inlet temperature at 120 °C, and the detector temperature at 150 °C. 2.5. Selective Degradation of Ammonia. Photocatalytic experiments for degradation of ammonia were conducted under NIR light irradiation (λ > 780 nm). A 300 W UV−visible lamp (OSRAM, Germany) was used as the light source. Briefly, 50 mL of ammonia solution was subjected to photocatalytic degradation in a 100 mL beaker at room temperature (25 ± 2 °C). The distance between the lamp and the test solution was approximately 10 cm. The wall of the beaker was protected from the surrounding light with an aluminum foil. NIR light was obtained with a cutoff filter of λ > 780 nm (see the picture of the cutoff filter in Figure S1 in the Supporting Information). The filter covered the window of the beaker to absorb UV−vis light and allow NIR light of λ > 780 nm to pass through. Typically, 50 mL of test solutions was used in the degradation experiments. A specific concentration of ammonia−organic compound solution was prepared according to the desired concentration. Briefly, 0.10 g of the catalysts was used for photocatalysis. NaHCO3−Na2CO3 buffer was used to control the pH of the solutions tested. A double-beam TU-1901 spectrophotometer was used to determine the ammonia concentration by using the Nessler reagent during photocatalysis. The Nessler reagent [alkaline solution of dipotassium tetraiodomercurate(II)] was prepared according to the following procedures. Briefly, 10 g of HgI2 and 7 g of KI were dissolved in water, NaOH solution (16 g of NaOH in 50 mL of water) was added to the mixture, and the mixture was diluted to 100 mL by adding deionized water. The Nessler reagent was stored in a dark bottle and diluted properly before analysis. Ammonia reacts with the Nessler reagent to obtain colored solutions by the reaction given in the following. The absorbance at 392 nm approached the peak and was recorded at the wavelength of 392 nm. The decline of the peak indicated the degradation of ammonia during the photocatalytic process.

3. RESULTS AND DISCUSSION 3.1. XRD Characterization. The as-synthesized samples were determined by XRD spectroscopy, and the resulting patterns are shown in Figure 1A. It can be seen from Figure 1

Figure 1. (A) XRD patterns of the samples: (a) α-MnO2/NG; (b) αMnO2/rGO; (c) NG; (d) rGO; and (e) α-MnO2. (B) Crystal structure of α-MnO2 using the [MnO6] model demonstrating [1 × 1] and [2 × 2] tunnels. The purple balls denote Mn atoms and red balls O atoms.

that the diffraction peaks appeared at 2θ = 12.8°, 18.0°, 28.8°, 37.4°, 50.0°, 56.4°, 60.2°, 65.0°, and 69.7°, which are indexed to a pure tetragonal phase of α-MnO2 (JCPDS, no. 41-0141) with I4/m space group and lattice constants a = 9.7980 and c = 2.8513 Å, corresponding to the Bragg reflections of (110), (220), (310), (211), (301), (411), (600), (521), (002), and (541) planes, respectively.34−37 There are one-dimensional 2 × 2 (4.6 Å × 4.6 Å) and 1 × 1 (1.9 Å × 1.9 Å) tunnels intergrowth along the c axis in the lattice shown in Figure 1B.37−40 No characteristic peaks are observed for the impurities, indicating high purity and crystallinity of the assynthesized samples. In addition, the diffraction peak at 26.31° was indexed to the plane (102) of rGO and NG, indicating that rGO and NG were formed during the hydrothermal process. As aforementioned, the 2 × 2 tunnels constructed of octahedral [MnO6] have a 4.6 Å × 4.6 Å space, which enables to accommodate cations. The tunnel structure makes it attractive to be used as photocatalysts because the photocatalytic process is a redox one involving electron transfer and associated cation shift to maintain the electrical neutrality equilibrium. 3.2. Raman Spectrum Characterization. To investigate the vibrational properties of the as-synthesized samples, the characterization of Raman spectroscopy has been conducted. Figure 2 presents the Raman spectra of α-MnO2/NG (a), αMnO2/rGO (b) composite, and α-MnO2 (c), rGO (d), and NG (e) components. The vibration peaks at 646 cm−1 were

NH4 + + 2[HgI4]2 − (yellow) + 4HO− → HgO·Hg(NH 2)I(brown, 392 nm) + 7I− + 3H 2O (2)

The products degraded were identified by ion chromatography, gas chromatography, electrochemistry, and UV−vis spectroscopy during the photocatalytic process (see Figures S5−S7 in the Supporting Information). 2.6. First-Principles Calculation. The calculations of band structure and density of states (DOS) were performed by using the CASTEP package of Material Studio 7.0. In the plane-wave calculations, a cutoff energy of 340 eV was applied. The generalized gradient approximation was adopted with Perdew− Burke−Ernzerhof functional. The calculations were performed in a ferromagnetic spin-polarized configuration. The selfconsistent field convergence criterion and energy tolerance were set as fine levels. A Monkhorst−Pack scheme with a 1 × 1 × 5 k-point grid was employed. The Taylor series method was used for dispersion-corrected density functional theory correction. The α-MnO2 phase is of a tetragonal structure with the space group 14/m (no. 87) and the lattice parameters a = 9.7980, c = 2.8513 Å, and α = β = γ = 90°, in which the Mn and O atoms occupy the fractional coordinates of Mn (0.35, 5539

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega

as the d-spacing value of the plane (002) in pure NG sheets in Figure 3D. It also shows the compact alignment between αMnO2 and NG. 3.4. XPS) Characterization. XPS measurements of the assynthesized α-MnO2/NG samples have been carried out to further confirm the structure and composition. The asmeasured results are presented in Figure 4. The total survey revealed the presence of Mn, O, and N atoms in the assynthesized α-MnO2/NG sample, as shown in Figure 4A. Highresolution XPS pattern indicated two peaks of 642.27 and 654.07 eV (Figure 4B), which are attributed to the binding energy of Mn 2p3/2 and Mn 2p1/2, respectively. The spin energy separation of 11.80 eV is in very good accordance with the reported datum on MnO2, confirming the presence of αMnO2,40,45 which is also consistent with the XRD result as reported previously. The XPS spectrum of O 1s binding energy is shown in Figure 4C, and the two peaks centered at 530.1 and 531.29 eV corresponded to Mn−O−Mn and Mn−O−H, respectively, in the α-MnO2/NG composite. The presence of Mn−O−H indicates that oxygen atoms linked to some of Mn atoms in the MnO6 octahedra may interact with hydrogen atoms via a hydrogen bond or a covalent coordination bond because water and related hydrides (H3O+) can also be accommodated in the tunnels.40,46 Three different types of nitrogen configurations (Figure 4D), including pyridinic N (397.13 eV), pyrrolic N (399.75 eV), and quaternary N (graphitic N, 401.68 eV) revealed that nitrogen atoms were bonded in the graphene framework.47 3.5. UV−Visible−NIR Reflection Spectra. Figure 5 shows the UV−visible−NIR reflection spectra of the as-synthesized αMnO2 (a), α-MnO2/rGO (b), and α-MnO2/NG (c) species. From the spectra, the absorption peaks were observed at 380 nm in the UV region, and an absorption band for the pure αMnO2 species was observed in the visible and NIR region as shown in curve (a). The absorption peak in the NIR region is assigned to the electron transition derived from the mixing of metal d−d and oxo-to-metal charge transfer,48−52 which is also confirmed by our calculation using the CASTEP method below. On the basis of the UV−visible−NIR reflection spectra, the direct and indirect optical band gaps of the as-synthesized αMnO2 (a), α-MnO2/rGO (b), and α-MnO2/NG (c) species can be determined by Tauc’s relation

Figure 2. Raman spectra of the as-synthesized samples: (a) α-MnO2/ NG; (b) α-MnO2/rGO; (c) α-MnO2; (d) rGO; and (e) NG.

assigned to the Ag mode originating from breathing vibrations of the MnO6 octahedra, indicating the α-MnO2 crystallite.41 This observation is in well-agreement with the XRD results. The G band at 1580 cm−1 in the rGO sample is characteristic of sp2-hybridized C−C bonds in a two-dimensional hexagonal lattice, and the band shifted to 1600 cm−1 after nitrogen atoms were doped in the lattice as shown in the α-MnO2/NG (a) and NG (e) samples. The D band located at 1335 cm −1 corresponds to the defects and disordered carbon in the graphitic layers.42 In addition, the broader 2D peaks appeared at around 2645 and 2905 cm−1, which is consistent with that of the few-layer NG reported by Yu.43 3.3. TEM Observation. Figure 3A,B shows the representative TEM images of the obtained α-MnO2/NG samples.

(αhν)n = A(hν − Eg )

(3)

where A is the constant, hν is the photon energy, and α is the absorption coefficient, while n = 2 for direct and n = 1/2 for indirect transition. It has been known that α-MnO2 is an indirect semiconductor.46 Thereby, the indirect band-gap values are given, as shown in Figure 6, which are equal to 1.16, 1.00, and 0.72 eV for α-MnO2, α-MnO2/rGO, and α-MnO2/NG, respectively. The indirect band gap for pure α-MnO2 is 1.16 eV, which is very close to 0.98 eV reported in the literature.16 The transition corresponds to the NIR wavelength of 1069 nm, which provides a profound basis for using NIR irradiation as power for photocatalytic applications. The band gap is further reduced to 0.72 eV after it was combined with NG as shown in curve (c) in Figure 6, implying that the α-MnO2/NG species can absorb more NIR irradiation as a photocatalyst for a more effective photocatalysis. As predicted, the combination of rGO with α-MnO2 led to a narrow indirect band gap, which is equal to 1.00 eV. These facts prove the strong interaction between αMnO2 and NG or rGO.

Figure 3. TEM images of the as-synthesized samples: (A,B) α-MnO2/ NG; (C) high-resolution TEM image; and (D) pure NG.

Dispersive nanoflowers can be observed throughout the samples. Most of the flowers are oval-shaped with the semi major axis of 140 nm and semi minor axis of 95 nm, as shown in Figure 3B. The thickness of prevailing petals is among 3.5− 8.0 nm, and this nanostructure favors to adsorb the randomly approaching molecules to decline the surface tension. Highresolution TEM image (Figure 3C) indicates the polycrystalline feature of α-MnO2 in the as-synthesized α-MnO2/NG sample. A finger distance of 0.50 nm corresponds to the d-spacing value of the plane (200) in α-MnO2,44 which is in agreement with the above XRD analysis. The other finger distance adjacent to αMnO2 was found to be 0.34 nm in Figure 3C, which is the same 5540

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega

Figure 4. XPS spectra of the as-synthesized samples: (A) wide-scan XPS spectra of α-MnO2/NG (a) and α-MnO2 (b); (B) Mn 2p XPS spectra in αMnO2/NG (a) and α-MnO2 (b); (C) O 1s XPS spectra in α-MnO2/NG (a) and α-MnO2 (b); and (D) N 1s XPS spectra in α-MnO2/NG.

3.6. Band Structure and DOS. The band structure (Figure 7A) and total DOS and partial DOS (PDOS) of α-MnO2 (Figure 7B) were calculated by the first-principles method, resulting in an indirect band gap of 0.909 eV, which is close to 1.160 eV determined by our experimental measurements, which is also in very good agreement with that calculated previously.16,46,53 As seen from Figure 7B, the total DOS are provided by s, p, and d orbitals. Obviously, 3d orbitals in Mn atoms (Figure 7C) were split into the low energetic level of t2g and high energetic level of eg in the MnO6 octahedral field46,48−52 and contributed to the valence band and conduction band (Ec), together with p orbitals provided by O atoms (Figure 7D). This band structure of α-MnO2 favors the absorption of incident NIR irradiation. 3.7. Selective Degradation. As seen in Figure S2 in the Supporting Information, the α-MnO2/NG catalyst can degrade ammonia under NIR light irradiation, and Figure S3 shows 3.0% of the NG content reaching the maximum degradation ratio (Figure S3). Given the coexistence of ammonia together with organic pollutant in practical waters, RHB and MO were chosen as organic model pollutants to examine the selectivity of the α-MnO2/NG catalyst. The results indicated that the degradation ratio of ammonia reached over 93.0% under NIR light irradiation in NH3−MO mixed solution or NH3−RHB mixed one as shown in Figure 8, while the ratios were only 9.7 and 9.4% for MO and RHB, respectively. The degradation ratio of ammonia is much higher than that of MO or RHB under the same conditions, indicating that the α-MnO2/NG photocatalyst can recognize and degrade ammonia selectively from MO or RHB in their coexistence. Under the similar conditions, humic acid was used to examine the selectivity of the α-MnO2/NG photocatalyst, and the results indicated that 15.7% of humic acid was degraded for 7.0 h (Figure S4), while over 93.0% of

Figure 5. UV−vis−NIR diffuse reflection spectra of the as-synthesized samples: (a) α-MnO2; (b) α-MnO2/rGO; and (c) α-MnO2/NG.

Figure 6. Tauc plot of (αhν)1/2 vs photo energy for the indirect band gaps of the samples: (a) α-MnO2; (b) α-MnO2/rGO; and (c) αMnO2/NG.

5541

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega

Figure 7. Band structures of α-MnO2 along the Z−Γ−X−P−N−Γ lines (A). Orange red and blue curves denote up- and down-spin energy levels at each k-point, respectively. Total DOS and PDOS of α-MnO2 (B). PDOS of Mn atoms (C). PDOS of O atoms (D). The energies are aligned to the Fermi level marked by a vertical dashed line.

Figure 9. Adsorption of NH3, MO, and RHB on the α-MnO2/NG catalyst in 50 mL solutions in the coexistence of 50 mg/L NH3−N (a) plus 50 mg/L MO (b) and in the coexistence of 50 mg/L NH3−N (c) plus 50 mg/L RHB (d) for 8 h in dark. The mass of the α-MnO2/NG catalyst was 0.10 g, and pHs were adjusted to 10.50 by adding 0.1 mol/ L NaHCO3−Na2CO3 buffer.

Figure 8. Selective degradation of ammonia in the presence of 50 mg/ L NH3−N (a) plus 50 mg/L MO (b) in 50 mL solution under NIR light irradiation and in the presence of 50 mg/L NH3−N (c) plus 50 mg/L RHB (d) in 50 mL solution under NIR light irradiation.

ammonia was eliminated in 50.0 mg/L humic acid and 50.0 mg/L NH3−N mixed solution. It confirmed further the selectivity of the photocatalyst for ammonia. The subsequent investigations indicated the adsorption of ammonia, MO, and RHB on the catalyst; the adsorption ratios were 17.6 and 2.1% for ammonia and MO, respectively, in NH3−MO mixed solution in dark for 8 h and 17.1 and 3.1% for ammonia and RHB, respectively, in NH3−RHB mixed solution (Figure 9). XPS measurements showed the adsorption of NH3, MO, and RHB molecules on the catalyst with the same shift of 0.2 eV at Mn 2p3/2 (Figure 10), implying a nonspecific adsorption. That is, the adsorption is a physical one but not a chemical one. In other words, the selective degradation of ammonia using α-

MnO2/NG as the photocatalyst in NH3−MO or NH3−RHB mixed solution under NIR irradiation could be attributed to another reason. 3.8. Insights of Selectivity and Reaction Mechanism. As estimated by experiments, the indirect band gap of αMnO2/NG is 0.72 eV, which corresponds to the energetic difference between valance band and conduction band. Figure 11 presents the valence band XPS spectrum of the α-MnO2 species. The top of valance band is the highest binding energy of the filled electrons and is defined as the intercept at 0.00 eV on the background. Therefore, the valence band of α-MnO2/ NG can be obtained as 1.44 eV based on the XPS measurement 5542

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega

at the so-called Dirac point. For the NG, the Fermi level is shifted up into the conduction band, which is due to the electron donation from the doped-nitrogen atom to the antibonding π* band.57,58 Graphitic nitrogen atoms in NG resulted in shifting up with the Fermi energy level of 0.66 V versus NHE,59 that is, the conduction band level of NG is 0.66 V versus NHE. The level (0.66 V) of conduction band is much negative than that of 1.23 V versus NHE, which shows the potential of oxygen being reduced to water. Therefore, oxygen can accept photoelectrons from the conduction band of NG to form water. Furthermore, the band gap of NG is in the range of 0.045−0.550 V, depending on the nitrogen-doped content from 0.5 to 8.0%.60 In the case here (ca. 7.0%), the band gap is about 0.48 eV, so the valence band level of NG is 1.14 V. The valence band level of NG could accept the photogenerated electrons transferred from the conduction band of α-MnO2. The band structure of NG was combined with that of α-MnO2, establishing the Z-scheme photocatalytic system for the removal of ammonia.61,62 Therefore, the overall mechanism for selective oxidization of ammonia can be presented in Figure 12. The product has been identified as N2 gas by means of

Figure 10. Binding energies of Mn 2p in α-MnO2 samples: (a) pure αMnO2; (b) after immersing in 50 mg/L NH3−N solution for 8 h; (c) after immersing in 50 mg/L MO solution for 8 h; and (d) after immersing in 50 mg/L RHB solution for 8 h.

Figure 11. Valence band measurement of α-MnO2 by XPS spectrum.

because the valence band is mainly composed of O 2p and Mn 3d orbitals as seen in Figure 7B. The conduction band can be estimated as 0.72 eV because the band gap is 0.72 eV. As far as photocatalytic reactions are concerned, they involve the oxidization process driven by photogenerated holes and the reduction process driven by photogenerated electrons. In a viewpoint of electrochemistry, the photogenerated holes can be regarded as oxidizing agents for the oxidization process, while the photogenerated electrons can be regarded as reducing agents for the reduction process. The photogenerated holes could oxidize the species collided only when the holes on top of the valence band are more positive than the redox potential of the species collided (reducing agents). It has been known that the electrode potential for E0(N2/NH3) is equal to 0.057 V versus normal hydrogen electrode (NHE), whereas the anodic peak potential is equal to about 1.18 V versus NHE for the oxidization of RHB54,55 and 1.497 V versus NHE for the oxidization of MO.56 Under consideration of over-potentials, the oxidization of ammonia to dinitrogen is much easier than that of MO or RHB to their oxidized states. Therefore, the preference of the valence band of α-MnO2 resulted in the selective oxidization of ammonia from MO or RHB in solutions. For the reduction process, the photogenerated electrons could reduce the species collided (oxidizing agents) only when the electrons excited on the bottom of the conduction band are more negative than the reduced potential of the species collided. The Fermi level for the perfect graphene is just located

Figure 12. NIR-driven selective photocatalytic degradation of ammonia via the α-MnO2/NG catalyst.

UV−vis spectroscopy, electrochemistry, ion chromatography, and gas chromatography (see Figures S5−S7 in the Supporting Information). It is concluded that the matched band structure of α-MnO2/NG oxidized ammonia selectively.

4. CONCLUSIONS The α-MnO2/NG hybrid catalyst is synthesized via a one-pot method. The theoretic band gap (0.909 eV) based on the firstprinciples calculation is very close to that (1.16 eV) measured by UV−vis−NIR diffuse reflectance spectroscopy, indicating that the theoretic simulation approaches to the real band structure of α-MnO2. The analysis based on first-principles calculation reveals that the valence band is composed of hybrid O 2p and Mn 3d orbitals and that the conduction band is composed of Mn 3d and O 2p orbitals. The band structure of α-MnO2 results in the response to NIR irradiation, which makes it as a NIR photocatalyst. The combination of α-MnO2 with NG to form α-MnO2/NG significantly extends the spectrum response to 1722 nm. The results show that αMnO2/NG as a photocatalyst can make full use of NIR irradiation. α-MnO2/NG can degrade ammonia selectively in a mixed solution of ammonia and MO or RHB or humic acid under NIR light irradiation to release nitrogen gas, which is 5543

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega attributed to the matched band structure of α-MnO2/NG and the valence band level of α-MnO2.



spheres Based on Lanthanide-Doped NaYF4, TiO2, and Au. Adv. Funct. Mater. 2015, 25, 2950−2960. (7) Ortega-Liebana, M. C.; Hueso, J. L.; Larrea, A.; Sebastian, V.; Santamaria, J. Feroxyhyte Nanoflakes Coupled to Up-converting Carbon Nanodots: a Highly Active, Magnetically Recoverable, Fentonlike Photocatalyst in the Visible-nir Range. Chem. Commun. 2015, 51, 16625−16628. (8) Xia, X.; Deng, N.; Cui, G.; Xie, J.; Shi, X.; Zhao, Y.; Wang, Q.; Wang, W.; Tang, B. NIR Light Induced H2 Evolution by a Metal-free Photocatalyst. Chem. Commun. 2015, 51, 10899−10902. (9) Hao, Y.; Dong, X.; Zhai, S.; Wang, X.; Ma, H.; Zhang, X. Controllable Self-assembly of a Novel Bi2MoO6-based Hybrid Photocatalyst: Excellent Photocatalytic Activity Under UV, Visible and Near-infrared Irradiation. Chem. Commun. 2016, 52, 6525−6528. (10) Zhang, X.; Yu, L.; Zhuang, C.; Peng, T.; Li, R.; Li, X. Highly Asymmetric Phthalocyanine as a Sensitizer of Graphitic Carbon Nitride for Extremely Efficient Photocatalytic H2 Production under Near-infrared light. ACS Catal. 2014, 4, 162−170. (11) Cui, J.; Li, Y.; Liu, L.; Chen, L.; Xu, J.; Ma, J.; Fang, G.; Zhu, E.; Wu, H.; Zhao, L.; Wang, L.; Huang, Y. Near-infrared plasmonicenhanced solar energy harvest for highly efficient photocatalytic reactions. Nano Lett. 2015, 15, 6295−6301. (12) Wang, J.-H.; Chen, M.; Luo, Z.-J.; Ma, L.; Zhang, Y.-F.; Chen, K.; Zhou, L.; Wang, Q.-Q. Ceria-Coated Gold Nanorods for PlasmonEnhanced Near-Infrared Photocatalytic and Photoelectrochemical Performances. J. Phys. Chem. C 2016, 120, 14805−14812. (13) Chen, K.-H.; Pu, Y.-C.; Chang, K.-D.; Liang, Y.-F.; Liu, C.-M.; Yeh, J.-W.; Shih, H.-C.; Hsu, Y.-J. Ag-Nanoparticle-Decorated SiO2 Nanospheres Exhibiting Remarkable Plasmon-Mediated Photocatalytic Properties. J. Phys. Chem. C 2012, 116, 19039−19045. (14) Lume-Pereira, C.; Baral, S.; Henglein, A.; Janata, E. Chemistry of Colloidal Manganese Dioxide. Part 1. Mechanism of Reduction by an Organic Radical (A Radiation Chemical Study). J. Phys. Chem. 1985, 89, 5772−5778. (15) Islam, A. K. M. F. U.; Islam, R.; Khan, K. A. Effects of Deposition Variables on Spray-deposited MnO2 Thin Films Prepared from Mn(C2H3O2)2·4H2O. Renewable Energy 2005, 30, 2289−2302. (16) Li, W.; Cui, X.; Zeng, R.; Du, G.; Sun, Z.; Zheng, R.; Ringer, S. P.; Dou, S. X. Performance Modulation of α-MnO2 Nanowires by Crystal Facet Engineering. Sci. Rep. 2015, 5, 8987. (17) Gao, T.; Glerup, M.; Krumeich, F.; Nesper, R.; Fjellvåg, H.; Norby, P. Microstructures and Spectroscopic Properties of Cryptomelane-type Manganese Dioxide Nanofibers. J. Phys. Chem. A 2008, 112, 13134−13140. (18) Sakai, N.; Ebina, Y.; Takada, K.; Sasaki, T. Photocurrent Generation from Semiconducting Manganese Oxide Nanosheets in Response to Visible Light. J. Phys. Chem. B 2005, 109, 9651−9655. (19) Shen, R.; Xie, J.; Lu, X.; Chen, X.; Li, X. Bifunctional Cu3P Decorated g-C3N4 Nanosheets as a Highly Active and Robust VisibleLight Photocatalyst for H2 Production. ACS Sustainable Chem. Eng. 2018, 6, 4026−4036. (20) Zhao, X.; Yang, H.; Cui, Z.; Li, R.; Feng, W. Enhanced photocatalytic performance of A−Bi4Ti3O12 nanocomposites prepared by a photocatalytic reduction method. Mater. Technol. 2017, 32, 870− 880. (21) Wang, F.; Yang, H.; Zhang, Y. Enhanced photocatalytic performance of CuBi2O4 particles decorated with Ag nanowires. Mater. Sci. Semicond. Process. 2018, 73, 58−66. (22) Song, Y.; Li, N.; Chen, D.; Xu, Q.; Li, H.; He, J.; Lu, J. N-Doped and CdSe-Sensitized 3D-Ordered TiO2 Inverse Opal Films for Synergistically Enhanced Photocatalytic Performance. ACS Sustainable Chem. Eng. 2018, 6, 4000−4007. (23) Dai, L.; Xue, Y.; Qu, L.; Choi, H.-J.; Baek, J.-B. Metal-free Catalysts for Oxygen Reduction Reaction. Chem. Rev. 2015, 115, 4823−4892. (24) Wang, H.; Maiyalagan, T.; Wang, X. Review on Recent Progress in Nitrogen-doped Graphene: Synthesis, Characterization, and Its Potential Applications. ACS Catal. 2012, 2, 781−794.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b00161. Picture of a black cutoff filter, comparison of removal of ammonia under different conditions, effects of NG contents on photocatalytic performance, selective degradation of ammonia in the presence of humic acid, identification of products yielded by means of UV−vis spectroscopy, electrochemistry, and gas chromatography (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected], shouqing_liu@hotmail. com. Phone: +86-512-69379159. Fax: +86-512-69379159. ORCID

Shou-Qing Liu: 0000-0002-3774-4180 Author Contributions †

W.-X.L. and X.-L.Z. made equal contribution to this work.

Notes

The authors declare the following competing financial interest(s): apply for patents.



ACKNOWLEDGMENTS This work is financially supported by the National Natural Science Foundation of China (nos. 21576175 and 51502187), the key industrial prospective program of Jiangsu Science and Technology Department (BE2015190), the Natural Science Foundation of Jiangsu Province of China (no. BK20141178), the Opening Project of Key Laboratory of Green Chemistry of Sichuan Institutes of Higher Education (no. LZJ1304), the Earmarked Nanotechnology Fund of the Bureau of Science and Technology of Suzhou City (no. ZXG201429), and Collaborative Innovation Center of Technology and Material of Water Treatment. The authors also appreciate supports from National Supercomputing Center in Shenzhen (Shenzhen Cloud Computing Center).



REFERENCES

(1) Qin, W.; Zhang, D.; Zhao, D.; Wang, L.; Zheng, K. Near-infrared Photocatalysis Based on YF3:Yb3+,Tm3+/TiO2 Core/shell Nanoparticles. Chem. Commun. 2010, 46, 2304−2306. (2) Tang, Y.; Di, W.; Zhai, X.; Yang, R.; Qin, W. NIR-Responsive Photocatalytic Activity and Mechanism of NaYF4:Yb,Tm@TiO2 Core−Shell Nanoparticles. ACS Catal. 2013, 3, 405−412. (3) Li, Z.-X.; Shi, F.-B.; Zhang, T.; Wu, H.-S.; Sun, L.-D.; Yan, C.-H. Ytterbium Stabilized Ordered Mesoporous Titania for Near-infrared Photocatalysis. Chem. Commun. 2011, 47, 8109−8111. (4) Chen, C. K.; Chen, H. M.; Chen, C.-J.; Liu, R.-S. Plasmonenhanced Near-infrared-active Materials in Photoelectrochemical Water Splitting. Chem. Commun. 2013, 49, 7917−7919. (5) Wang, W.; Huang, W.; Ni, Y.; Lu, C.; Xu, Z. Different Upconversion Properties of β-NaYF4:Yb3+,Tm3+/Er3+ in Affecting the Near-Infrared-Driven Photocatalytic Activity of High-Reactive TiO2. ACS Appl. Mater. Interfaces 2014, 6, 340−348. (6) Xu, Z.; Quintanilla, M.; Vetrone, F.; Govorov, A. O.; Chaker, M.; Ma, D. Harvesting Lost Photons: Plasmon and Upconversion Enhanced Broadband Photocatalytic Activity in Core@Shell Micro5544

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega

(42) Mou, Z.; Wu, Y.; Sun, J.; Yang, P.; Du, Y.; Lu, C. TiO2 Nanoparticles-functionalized N-doped Graphene with Superior Interfacial Contact and Enhanced Charge Separation for Photocatalytic Hydrogen Generation. ACS Appl. Mater. Interfaces 2014, 6, 13798−13806. (43) Chen, P.; Xiao, T.-Y.; Li, H.-H.; Yang, J.-J.; Wang, Z.; Yao, H.B.; Yu, S.-H. Nitrogen-doped Graphene/Znse Nanocomposites: Hydrothermal Synthesis and Their Enhanced Electrochemical and Photocatalytic Activities. ACS Nano 2012, 6, 712−719. (44) Zhou, M.; Zhang, X.; Wei, J.; Zhao, S.; Wang, L.; Feng, B. Morphology-controlled Synthesis and Novel Microwave Absorption Properties of Hollow Urchinlike α-MnO2 Nanostructures. J. Phys. Chem. C 2011, 115, 1398−1402. (45) Lei, Z.; Shi, F.; Lu, L. Incorporation of MnO2-coated Carbon Nanotubes Between Graphene Sheets as Supercapacitor Electrode. ACS Appl. Mater. Interfaces 2012, 4, 1058−1064. (46) Cockayne, E.; Li, L. First-principles DFT+U Studies of the Atomic, Electronic, and Magnetic Structure of α-MnO2 (cryptomelane). Chem. Phys. Lett. 2012, 544, 53−58. (47) Shin, S. R.; Park, J. H.; Kim, K.-H.; Choi, K. M.; Kang, J. K. Network of Heterogeneous Catalyst Arrays on the Nitrogen-doped Graphene for Synergistic Solar Energy Harvesting of Hydrogen from Water. Chem. Mater. 2016, 28, 7725−7730. (48) Gamelin, D. R.; Kirk, M. L.; Stemmler, T. L.; Pal, S.; Armstrong, W. H.; Penner-hahn, J. E.; Solomon, E. I. Electronic Structure and Spectroscopy of Manganese Catalase and Di-.mu.-oxo [MnIII MnIV] Model Complexes. J. Am. Chem. Soc. 1994, 116, 2392−2399. (49) Reed, J.; Ceder, G. Role of Electronic Structure in the Susceptibility of Metastable Transition-metal Oxide Structures to Transformation. Chem. Rev. 2004, 104, 4513−4534. (50) Jana, S.; Pande, S.; Sinha, A. K.; Sarkar, S.; Pradhan, M.; Basu, M.; Saha, S.; Pal, T. A Green Chemistry Approach for the Synthesis of Flower-like Ag-doped MnO2 Nanostructures Probed by Surfaceenhanced Raman Spectroscopy. J. Phys. Chem. C 2009, 113, 1386− 1392. (51) Pinaud, B. A.; Chen, Z.; Abram, D. N.; Jaramillo, T. F. Thin Films of Sodium Birnessite-type MnO2: Optical Properties, Electronic Band Structure, and Solar Photoelectrochemistry. J. Phys. Chem. C 2011, 115, 11830−11838. (52) Kwon, K. D.; Refson, K.; Sposito, G. On the role of Mn(IV) Vacancies in the Photoreductive Dissolution of Hexagonal Birnessite. Geochim. Cosmochim. Acta 2009, 73, 4142−4150. (53) Zhou, G.; Liu, H.; Ma, Z.; Li, H.; Pei, Y. Spatially Confined Li− Oxygen Interaction in the Tunnel of α-MnO2 Catalyst for Li−Air Battery: A First-Principles Study. J. Phys. Chem. C 2017, 121, 16193− 16200. (54) Austin, J. M.; Harrison, I. R.; Quickenden, T. I. Electrochemical and Photoelectrochemical properties of Rhodamine B. J. Phys. Chem. 1986, 90, 1839−1843. (55) Yu, L.; Mao, Y.; Qu, L. Simple Voltammetric Determination of Rhodamine B by Using the Glassy Carbon Electrode in Fruit Juice and Preserved Fruit. Food Anal. Method. 2013, 6, 1665−1670. (56) Li, X.; Li, X.; Yang, W.; Chen, X.; Li, W.; Luo, B.; Wang, K. Preparation of 3D PbO2, Nanospheres@SnO2 Nanowires/Ti Electrode and Its Application in Methyl Orange Degradation. Electrochim. Acta 2014, 146, 15−22. (57) Umeki, T.; Akaishi, A.; Ichikawa, A.; Nakamura, J. Anomalous Stabilization in Nitrogen-Doped Graphene. J. Phys. Chem. C 2015, 119, 6288−6292. (58) Robertson, J.; Davis, C. A. Nitrogen Doping of Tetrahedral Amorphous Carbon. Diamond Relat. Mater. 1995, 4, 441−444. (59) Jing, Y.; Zhou, Z. Computational Insights into Oxygen Reduction Reaction and Initial Li2O2 Nucleation on Pristine and NDoped Graphene in Li−O2 Batteries. ACS Catal. 2015, 5, 4309−4317. (60) Lherbier, A.; Botello-Méndez, A. R.; Charlier, J.-C. Electronic and Transport Properties of Unbalanced Sublattice N-Doping in Graphene. Nano Lett. 2013, 13, 1446−1450. (61) Li, X.; Xie, K.; Song, L.; Zhao, M.; Zhang, Z. Enhanced Photocarrier Separation in Hierarchical Graphitic-C3N4-Supported

(25) Tanaka, A.; Hashimoto, K.; Kominami, H. Selective Photocatalytic Oxidation of Aromatic Alcohols to Aldehydes in An Aqueous Suspension of Gold Nanoparticles Supported on Cerium(IV) Oxide Under Irradiation of Green Light. Chem. Commun. 2011, 47, 10446− 10448. (26) Zhou, Y.; Yi, Q.; Xing, M.; Shang, L.; Zhang, T.; Zhang, J. Graphene Modified Mesoporous Titania Single Crystals with Controlled and Selective Photoredox Surfaces. Chem. Commun. 2016, 52, 1689−1692. (27) Song, L.-N.; Chen, L.; He, J.; Chen, P.; Zeng, H.-K.; Au, C.-T.; Yin, S.-F. The First Synthesis of Bi self-doped Bi2MoO6−Bi2Mo3O12 Composites and Their Excellent Photocatalytic Performance for Selective Oxidation of Aromatic Alkanes Under Visible Light Irradiation. Chem. Commun. 2017, 53, 6480−6483. (28) Liu, S.-Q.; Zhu, X.-L.; Zhou, Y.; Meng, Z.-D.; Chen, Z.-G.; Liu, C.-B. 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. (29) Ghosh-Mukerji, S.; Haick, H.; Schvartzman, M.; Paz, Y. Selective Photocatalysis by Means Of Molecular Recognition. J. Am. Chem. Soc. 2001, 123, 10776−10777. (30) Shen, X.; Zhu, L.; Liu, G.; Yu, H.; Tang, H. Enhanced Photocatalytic Degradation and Selective Removal of Nitrophenols by Using Surface Molecular Imprinted Titania. Environ. Sci. Technol. 2008, 42, 1687−1692. (31) Wu, J.; Zhang, D.; Wang, Y.; Hou, B. Electrocatalytic Activity of Nitrogen-doped Graphene Synthesized Via A One-pot Hydrothermal Process Towards Oxygen Reduction Reaction. J. Power Sources 2013, 227, 185−190. (32) Liu, S.-Q.; Xiao, B.; Feng, L.-R.; Zhou, S.-S.; Chen, Z.-G.; Liu, C.-B.; Chen, F.; Wu, Z.-Y.; Xu, N.; Oh, W.-C.; Meng, Z.-D. Graphene Oxide Enhances the Fenton-like Photocatalytic Activity of Nickel Ferrite For Degradation of Dyes Under Visible Light Irradiation. Carbon 2013, 64, 197−206. (33) 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. (34) Xu, M.; Kong, L.; Zhou, W.; Li, H. Hydrothermal Synthesis and Pseudocapacitance Properties of α-MnO2 Hollow Spheres and Hollow Urchins. J. Phys. Chem. C 2007, 111, 19141−19147. (35) Zhang, Y.; Yang, Y.; Zhang, Y.; Zhang, T.; Ye, M. Heterogeneous Oxidation of Naproxen in the Presence of α-MnO2, Nanostructures with Different Morphologies. Appl. Catal., B 2012, 127, 182−189. (36) Jang, H.; Zahoor, A.; Kim, Y.; Christy, M.; Oh, M. Y.; Aravindan, V.; Lee, Y. S.; Nahm, K. S. Tailoring Three Dimensional α−MnO2/RuO2, Hybrid Nanostructure as Prospective Bifunctional Catalyst for Li−O2 Batteries. Electrochim. Acta 2016, 212, 701−709. (37) Yuan, Y.; Nie, A.; Odegard, G. M.; Xu, R.; Zhou, D.; Santhanagopalan, S.; He, K.; Asayesh-Ardakani, H.; Meng, D. D.; Klie, R. F.; Johnson, C.; Lu, J.; Shahbazian-Yassar, R. Asynchronous Crystal Cell Expansion During Lithiation of K+-Stabilized α-MnO2. Nano Lett. 2015, 15, 2998−3007. (38) Devaraj, S.; Munichandraiah, N. Effect of Crystallographic Structure of MnO2 on Its Electrochemical Capacitance Properties. J. Phys. Chem. C 2008, 112, 4406−4417. (39) Li, L.; King, D. L. Synthesis and Characterization of Silver Hollandite and Its Application in Emission Control. Chem. Mater. 2005, 17, 4335−4343. (40) Boppana, V. B. R.; Jiao, F. Nanostructured MnO2: An Efficient and Robust Water Oxidation Catalyst. Chem. Commun. 2011, 47, 8973−8975. (41) Qu, J.; Shi, L.; He, C.; Gao, F.; Li, B.; Zhou, Q.; Hu, H.; Shao, G.; Wang, X.; Qiu, J. Highly Efficient Synthesis of Graphene/MnO2 Hybrids and Their Application for Ultrafast Oxidative Decomposition of Methylene Blue. Carbon 2014, 66, 485−492. 5545

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546

Article

ACS Omega CuInS2 for Noble-Metal-Free Z-Scheme Photocatalytic Water Splitting. ACS Appl. Mater. Interfaces 2017, 9, 24577−24583. (62) Meng, Q.; Lv, H.; Yuan, M.; Chen, Z.; Chen, Z.; Wang, X. In Situ Hydrothermal Construction of Direct Solid-State Nano-Z-Scheme BiVO4/Pyridine-Doped g-C3N4 Photocatalyst with Efficient VisibleLight-Induced Photocatalytic Degradation of Phenol and Dyes. ACS Omega 2017, 2, 2728−2739.

5546

DOI: 10.1021/acsomega.8b00161 ACS Omega 2018, 3, 5537−5546