Structure–performance relationships of MnO2 nanocatalyst for the low-temperature SCR removal of NOX under ammonia

Yi Li*a, Yanping Lia, Yuan Wana, Sihui Zhan*b, Qingxin Guanb and Yang Tianc
aDepartment of Chemistry, Tianjin University, Tianjin 300072, P. R. China. E-mail: liyi@tju.edu.cn
bCollege of Environmental Science and Engineering, Nankai University, Tianjin 300071, P. R. China. E-mail: sihuizhan@nankai.edu.cn; Fax: +86-22-23502756; Tel: +86-22-23502756
cDepartment of Chemistry, Capital Normal University, Beijing 100048, P. R. China

Received 2nd February 2016 , Accepted 23rd May 2016

First published on 24th May 2016


Abstract

To investigate the corresponding relationship between catalytic efficiency and structure, MnO2 nanomaterials (nanospheres, nanosheets, nanorods) have been prepared successfully, and were thoroughly characterized by SEM and TEM. Furthermore, the selective catalytic reduction (SCR) performance of NOX under ammonia was used as an indicative reaction. Among the MnO2 nanomaterials with different morphologies, it was found that their SCR activities showed an interesting variation tendency: nanospheres > nanosheets > nanorods of MnO2. The NO conversion ratio of the MnO2 nanospheres could reach 100% from 200 to 350 °C. Moreover, in order to study the probable mechanism for the best removal efficiency of the nanospheres, XRD, H2-TPR, NH3-TPD, BET, XPS and in situ DRIFTS were performed in detail. It is found that surface chemisorbed oxygen, specific surface area, reducibility and acid sites have great influence on the NO removal efficiency in the SCR reaction. In addition, how several process parameters affect the NOX removal efficiency was carried out, such as time, H2O and SO2.


1. Introduction

Emission of nitrogen oxides (NOX) from the industrial combustion of coal and fossil fuels has been one of the major causes of environmental pollution. NOX contributes to the formation of photo-chemical smog and acid rain, and ozone layer depletion, and have drawn wide concern due to their high toxicity to human health.1–6 Many approaches have been developed to remove NOX, among which SCR degradation of NOX with NH3 has been proven to be an effective technology.

As typical commercial catalysts, vanadium-based catalysts exhibit higher catalytic activity and selectivity in the SCR degradation of NOX with NH3. However, some inherent drawbacks still exist, such as the poor low temperature SCR activity, the toxicity of VOX, and the narrow operating temperature window.7–9 Therefore, it is necessary to develop several new types of catalysts that contain various metal oxides (Mn, Fe, Ni, Cu, Ce oxides) on different commercial supports. For example, Liu10 found that the addition of Ce not only enhanced the activity and alkali resistance of V2O5/TiO2 but also reduced its doped vanadium content. Chen11 found that novel Fe–Mn mixed-oxide catalysts showed the highest removal performance and 100% selectivity of N2. Among the many transition metal oxides, manganese oxides have attracted much attention due to their high catalytic activity, good redox properties, relatively low toxicities, good sintering resistance and low cost. As we all know, MnO2 has a redox cycle process between Mn2+ and Mn4+, providing oxygen mobility in the oxide lattice. However, Mn4+ ions are expected to exhibit higher activity compared to Mn3+ and Mn2+ in catalytic reactions.12,13 Manganese oxide-based catalysts (MnOX) exhibited a relatively high activity for the SCR reaction, such as MnOX/CeO2,14,15 MnOX/CuO,16,17 MnOX/ZrO2–CeO2,18 MnOX/TiO2,19,20 MnO2/Fe2O3,11,21 MnO2/NiCo2O4,22 and so on.

As we all know, SCR degradation of NOX with NH3 can be affected by many factors, for instance, the oxidation state of the transition metal element, the shape, the crystallinity and the specific surface.23 However, few studies have focused on the corresponding relation between catalytic efficiency and morphology or dimensions. For example, Bai24 et al. studied MnO2 catalyst with different dimensions. They found that 3D-MnO2 obtained the best catalytic properties, which is due to abundant surface adsorbed oxygen species, better reducibility and more Mn4+ ions. Zhao23 found that Co3O4 nanorods showed much higher efficiency in SCR degradation of NOX with NH3 than Co3O4 nanoparticles, which is due to more Co3+ adsorption on the Co3O4 nanorods. Gao et al.18 studied the morphology-dependent properties of MnOX/ZrO2–CeO2 nanostructures for NOX reduction with NH3. They found that the MnOX/ZrO2–CeO2 nanorods achieved significantly higher NOX removal efficiency than the nanocubes and nanopolyhedra, which is attributed to more Mn4+ species, adsorbed surface oxygen and oxygen vacancies.

In this paper, pure MnO2 nanocrystals with different morphologies, including bulk particles, nanorods, nanosheets and nanospheres were synthesized successfully and their catalytic activities in the NH3-SCR reaction were investigated, indicating that the MnO2 nanospheres exhibited the highest NO conversion ratio in the NH3-SCR reaction. Furthermore, its reaction mechanism was proposed using the analysis of SEM, TEM, BET, XRD, NH3-TPD, H2-TPR, XPS and in situ DRIFTS.

2. Experimental

2.1. Preparation of materials

All of the chemicals were analytical grade without further purification. Manganese(II) chloride tetrahydrate (MnCl2·4H2O) and urea were purchased from Sigma-Aldrich. Manganese sulfate (MnSO4) and ammonium persulfate (NH4)2S2O8 were purchased from Aladdin Industrial Inc. PEG 400, potassium permanganate (KMnO4) and sodium dodecyl benzene sulfonate (SDBS) were purchased from Jiangtian Chemical Technology Co. Ltd. (Tianjin, China).

The preparation methods of the three precursors were as follows: the β-MnO2 nanorods25 were synthesized by a hydrothermal method. In a typical process, 0.2 g KMnO4 and 4 mL PEG (400) were added into a 100 mL Teflon-sealed autoclave. Then 80 mL deionized water was added into the autoclave and kept under rapid magnetic agitation at room temperature for 10 min. The Teflon-sealed autoclave was heated to 160 °C for 5 h.

The synthesis process of γ-MnO2 nanosheets26 was similar to the above steps. 1.3 g SDBS and 0.4 g MnCl2·4H2O were added into 45 mL deionized water, then 0.2 g urea was dropped into this homogeneous solution under stirring. After continuous stirring for 30 min, the mixture was transferred into a Teflon-sealed autoclave and maintained at 110 °C for 12 h.

In a similar process, 1.5 g MnSO4 and 23 g (NH4)2S2O8 were prepared to synthesize γ-MnO2 nanospheres,27 which were dissolved in 50 mL deionized water and maintained at 110 °C for 10 h in a Teflon-sealed autoclave.

Then the obtained three precursors were filtered and washed with deionized water and absolute ethanol several times, respectively. After drying at 60 °C in an oven, the resulting powders were finally calcined at 400 °C for 5 h in air.

As a comparison, the MnO2 bulk nanoparticles were prepared directly by calcining Mn(NO3)2·4H2O (97%) at 400 °C for 5 h.

2.2. Characterization

MnO2 nanomaterials with different morphologies were investigated by field-emission scanning electron microscopy (FESEM) (JSM-6700F), and the size and morphology of the samples were observed by transmission electron microscopy (TEM) using a JEOL Model JEM-1200EX instrument with an accelerating voltage of 200 kV. The X-ray diffraction (XRD) measurements of the MnO2 materials were carried out on an X-ray diffractometer (Rigaku D/Max 2200 PC) over a 2θ range between 10° and 80° with the electric current being fixed at 40 mA and the voltage being operated at 40 kV, with a graphite monochrometer and Cu Kα radiation (λ = 0.15418 nm) at room temperature. Temperature-programmed reduction by hydrogen (H2-TPR) was conducted on a chemisorption analyzer (Micromeritics Autochem 2920 II instrument) containing a thermal conductivity detector (TCD). The experimental pretreatment was as follows. MnO2 material was outgassed at 100 °C for 1 h in flowing Ar with about 30–40 mg different catalysts. These samples were heated up to 700 °C under a 10% H2/Ar gas flow (50 mL min−1) at a rate of 10 °C min−1 after cooling to room temperature. Temperature-programmed desorption by ammonia (NH3-TPD) was performed on a chemisorption analyzer (Quantachrome Autosorb iQ-C-TCD-MS). Approximately 100 mg MnO2 samples were saturated with highly purified NH3 at 50 °C for 1 h and then flushed at 50 °C for 1 h to remove physically adsorbed NH3 after degassing at 300 °C for 2 h in flowing N2. In the end, these samples were heated from 200 °C to 700 °C at a heating rate of 10 °C min−1. The X-ray photoelectron spectroscopy (XPS) of the three different MnO2 samples was carried out on a spectrometer (Thermal ESCALAB 250, UK) with Al Kα radiation, and the binding energies were calibrated by using C 1s (C 1s = 284.6 eV). The specific surface areas and the pore size distributions of the three different shapes of MnO2 nanocrystals were observed respectively by the Brunauer–Emmett–Teller (BET) method, a cylindrical pore model (BJH method) and the adsorption branches. The in situ diffuse reflectance infrared Fourier transform spectroscopy (in situ DRIFTS) was carried out on a spectrometer (Thermo Nicolet iS50 FTIR). Prior to each test, the MnO2 catalyst was disposed at 350 °C in flowing argon for 60 min and then cooled to 150 °C. The sample spectra were recorded using the KBr pellet method by collecting 32 scans at a resolution of 4 cm−1 in the range of 400–4000 cm−1.

2.3. Catalytic performance tests

The catalytic activity tests for the three different MnO2 samples were conducted in a fixed-bed quartz reactor (inner diameter of 10 mm) with 500 mg of samples measuring 40–60 mesh and ranging from 50–400 °C under atmospheric pressure. The SCR reaction activity measurement by ammonia was as follows: 500 mg catalyst was put in the reactor. The total flow rate of the mixed gases (500 ppm NH3, 500 ppm NO, 5% H2O (when used), 3% O2, all balanced by nitrogen) was 200 mL min−1, amounting to a gas hourly space velocity (GHSV) of 2.8 × 104 h−1. The concentrations of O2, NO, NOX and NO2 were continually monitored by an FTIR spectrometer (GASMET FTIR DX4000, Finland). All the concentrations of the outlet gases were recorded when the SCR process reached a steady-state after half an hour at each temperature. To further count the activity of the three different shaped MnO2 samples, the NO conversion was obtained from the following equation, where NOout is the concentration of the outlet NO and NOin is the concentration of the inlet NO.
image file: c6ra03108k-t1.tif

NO oxidation to NO2 and N2 selectivity were obtained by using the following equations, where NOX is the sum of the NO and NO2 concentrations.

image file: c6ra03108k-t2.tif

image file: c6ra03108k-t3.tif

In order to parallelly compare the activities of the three different MnO2 samples, we performed turnover frequency (TOF) experiments. The total flow rate of the mixed gases (500 ppm NH3, 500 ppm NO, 3% O2, all balanced by nitrogen) was 3000 mL min−1, which is equivalent to a gas hourly space velocity (GHSV) of 4.2 × 105 h−1. By assuming that the reaction components were free of diffusion limitations, TOF values can be calculated according to the following equation:

image file: c6ra03108k-t4.tif

image file: c6ra03108k-t5.tif
where FNO is the concentration of inlet NO (L min−1), XNO is the NO conversion, W is the catalyst weight (g), and nMn4+ is the mole number of Mn4+ ions calculated by H2-TPR.

3. Results and discussion

3.1. Structural properties

The SEM images of the obtained MnO2 nanostructures are shown in Fig. 1. A large number of MnO2 nanorods are presented in Fig. 1a, where the nanorods are shown to have a smooth surface and typical rod-like shape. These high-purity MnO2 nanorods are about 100 nm in diameter and their lengths are in the range of 2–10 μm. Meanwhile, from the SEM image in Fig. 1b, the product is presented as a uniform hexagonal nanosheet with a side length of 1.2 μm and a thickness of approximately 200 nm. The surfaces of these uniform hexagonal nanosheets are smooth. Fig. 1c shows urchin-like nanospheres which are composed of numerous randomly oriented nanorods, and the diameters are about 4 μm. These nanorods are 400–800 nm in length and 100–150 nm in width.
image file: c6ra03108k-f1.tif
Fig. 1 SEM and HRTEM images of MnO2 nanorods (a, d), MnO2 nanosheets (b, e) and MnO2 nanospheres (c, f).

The detailed morphology structures of the MnO2 nanorods, nanosheets and nanospheres are further analyzed by HRTEM in Fig. 1. As shown in Fig. 1d, it can be seen that the MnO2 nanorods are well crystallized and the interplanar distances are 3.098 nm and 2.35 nm, corresponding to the (110) and (101) crystal planes, respectively. A pretty morphology is displayed in Fig. 1e. The interplanar distance of the MnO2 nanosheets is 2.25 nm, which corresponds to the (022) crystal plane. In Fig. 1f, the MnO2 nanospheres are composed of numerous randomly oriented nanorods, and the interplanar distance is 3.398 nm, corresponding to the (101) crystal plane.

Fig. 2 shows the XRD patterns of the MnO2 materials with different morphologies. The diffraction peaks of MnO2 bulk nanoparticles can be well identified as β-MnO2 (JCPDS 24-0735).28 The MnO2 nanorods can be determined as tetragonal-phase β-MnO2 (JCPDS 24-0735), which are produced by the pyrolysis of MnOOH nanorods, as illustrated in Fig. 2. The lattice constants are calculated, and a is 4.38 Å and c is 2.88 Å, which are very consistent with the reported values (a = 4.40 Å and c = 2.87 Å).29 The intensity ratios of (110) and the other reflections are different, a result of the preferential growth of the materials. The mechanism has been proved using many one-dimensional nanocrystals.30 The MnO2 hexagonal nanosheets could be identified as γ-MnO2 (JCPDS 30-0820), which is from the calcination of MnCO3 nanosheets and can be easily indexed to the lattice constants (a is 2.80 Å, b is 2.80 Å, c is 4.45 Å). The XRD patterns of MnO2 nanospheres show that the calcined products were γ-MnO2 (JCPDS 30-0820). From the full width at half maximum and intensity of the peaks, the diffraction peaks of the MnO2 nanospheres are higher than that of the MnO2 nanosheets, suggesting that the MnO2 nanospheres obtained high crystallinity.


image file: c6ra03108k-f2.tif
Fig. 2 XRD patterns of the MnO2 nanomaterials with different morphologies, such as nanorods, nanosheets, nanospheres and bulk particles.

The N2 adsorption–desorption isotherms and the pore size distribution, and the specific surface area and pore volume of the MnO2 bulk nanoparticles, MnO2 nanorods, MnO2 nanosheets and MnO2 nanospheres are shown in Fig. 3 and Table 1, respectively. In Fig. 3a, it can be seen that the MnO2 nanospheres and nanosheets all have a typical type IV isotherm with type-H3 hysteresis loops according to the definition of IUPAC. The hysteresis loops indicate the presence of mesopores.31,32 However, both the MnO2 bulk nanoparticles and MnO2 nanorods did not show the typical IV-type isotherm, indicating that the MnO2 bulk nanoparticles and MnO2 nanorods have no mesoporous channels. It can be further observed over the N2 adsorption–desorption isotherms that the turning point of the nitrogen-adsorbed volume for MnO2 nanorods and nanospheres is 0.9, and the N2 adsorbed volume further increases sharply under higher relative pressures, which is ascribed to the large mesopores (>10 nm) in the samples.33 As illustrated in Fig. 3b, the MnO2 nanospheres show a wide peak at 40–60 nm, which might correspond to the staggered arrangement of numerous randomly oriented MnO2 nanorods. The consequence is consistent with the results of SEM and TEM analysis. The MnO2 nanosheets exhibit a sharp peak at around 8 nm, which might result from the packing of the nanosheets. The pore distribution of MnO2 bulk nanoparticles and MnO2 nanorods shows that a few large mesopores exist. As shown in Table 1, the specific areas of the MnO2 bulk nanoparticles and MnO2 nanorods are lower than the MnO2 nanosheets and nanospheres, which is in good conformity with the results of NO removal efficiency in the SCR reaction. Furthermore, it can be seen that the specific surface area of the MnO2 nanosheets is almost equal to the nanospheres at about 16 m2 g−1. However, the pore volume of MnO2 nanospheres (0.14 cm3 g−1) is significantly higher than that of the nanosheets (0.07 cm3 g−1), which is due to them being composed of numerous randomly oriented nanorods of urchin-like MnO2 nanospheres.24 The similar specific surface area and large differences in the total pore volume of the MnO2 nanosheets and nanospheres are due to the pore diameters of the nanosheets being less than for the nanospheres.34 It is reported that the larger specific surface area and porous structures can offer more active sites and increase the adsorption of reactants in the SCR reaction.35,36 The specific surface area and pore volume of the MnO2 nanospheres are higher than the MnO2 nanosheets and nanorods, which is beneficial for the exposure of more active sites and adsorption of the reactants. From what has been discussed above, we may draw the conclusion that the MnO2 nanospheres obtained the highest catalytic activity in the SCR reaction and it might be due to the large specific surface area and pore volume.37 This is consistent with the above discussion.


image file: c6ra03108k-f3.tif
Fig. 3 N2 adsorption–desorption isotherms (a) and pore diameter distribution (b) of the MnO2 nanomaterials with different morphologies.
Table 1 Specific area, pore volume and pore diameter distribution of the MnO2 samples
Materials Specific area (m2 g−1) Pore volume (cm3 g−1) Pore diameter (nm)
MnO2 bulk nanoparticles 5.6 0.01 3.83
MnO2 nanorods 9.6 0.02 24.78
MnO2 nanosheets 16.0 0.07 12.58
MnO2 nanospheres 15.7 0.14 68.74


The chemical states of the elements in the near-surface region were investigated by XPS. The survey spectra of three different samples (Fig. 4a) show the existence of Mn and O, with C from the reference. Moreover, the atom molar rate of Mn and O is about 1[thin space (1/6-em)]:[thin space (1/6-em)]2 in the survey spectra, demonstrating the formation of MnO2. As illustrated in Fig. 4b, two XPS binding energies of Mn 2p1/2 and Mn 2p3/2 at 654.1 eV and 642.4 eV were observed for MnOX. It can be seen that the binding energy separation is 11.7 eV between the Mn 2p1/2 and Mn 2p3/2 states.38 As shown in Fig. 4b, Mn 2p1/2 can be divided into two characteristic peaks which can be ascribed to Mn4+ and Mn3+ ions displayed with binding energies of 654.8 eV and 654.1 eV, respectively. Moreover, Mn 2p3/2 can be divided into two characteristic peaks which can be ascribed to Mn4+ and Mn3+ ions displayed with binding energies of 642.9 eV and 641.8 eV, respectively. The relative atomic percentages of Mn3+ and Mn4+ over the surface of the MnO2 catalysts are shown in Table 2, respectively. We find that the MnO2 nanospheres display the highest Mn4+/Mnn+ molar ratios. As we all know, higher oxidation state manganese species are more active for the SCR reaction over manganese-based catalysts. It is well known that Mn4+ ions play a major role in low-temperature SCR reactions, which could promote the oxidation of nitric oxide to nitrogen dioxide.24,39 Therefore a high Mn4+/Mnn+ ratio will be conducive to the low-temperature SCR reaction. The Mn4+ content in the MnO2 nanospheres (55.6%) was much higher than that in the MnO2 nanosheets (49.1%) and nanorods (37.2%), which is consistent with the results of SCR activity.


image file: c6ra03108k-f4.tif
Fig. 4 XPS spectra of the MnO2 nanorods, MnO2 nanosheets and MnO2 nanospheres.
Table 2 XPS results of the MnO2 samples
Surface Atomic Concentration (%)
MnO2 materials 2p Oα (531.2) Oβ (529.5)
Mn4+/Mnn+ Mn3+/Mnn+
Nanorods 37.2 62.8 36.73 63.27
Nanosheets 49.2 50.8 47.90 52.10
Nanospheres 55.6 44.4 43.72 56.28


The O 1s XPS spectra of the three samples in Fig. 4c were deconvoluted using a curve-fitting procedure, and contain two kinds of distinguished surface oxygen species. The lower binding energy peak at 529.0–530.0 eV might be assigned to the surface lattice oxygen (Oβ), such as defective oxides or surface oxygen ions bonded to manganese in a coordinatively saturated environment, and the higher binding energy peak at 531.0–532.0 eV could belong to the surface chemisorbed oxygen (Oα).39 Generally, the surface chemisorbed oxygen (Oα) is more reactive than the surface lattice oxygen (Oβ), which is due to the higher mobility of Oα. Hence a high Oα/(Oα + Oβ) ratio will be conducive to the SCR reaction.40 The Oα ratio between the MnO2 nanospheres (43.72%) and nanosheets (47.9%) has no significant difference. The Oα rate of the MnO2 nanospheres and nanosheets were much higher than that of the MnO2 nanorods (36.73%), which corresponds to the lowest NO conversion efficiency of the MnO2 nanorods in the SCR reaction.

To investigate the reducibility of the MnO2 samples with different morphologies in the SCR reaction, including bulk nanoparticles, nanorods, nanosheets and nanospheres, the H2-TPR spectra of the MnO2 catalysts are shown in Fig. 5a. It can be seen that the three reduction peaks of the MnO2 bulk nanoparticles ranging from 350 to 600 °C are attributed to the reduction of MnO2 to Mn2O3, Mn2O3 to Mn3O4 and Mn3O4 to MnO, respectively.34,41 Moreover, the TPR spectra of the MnO2 nanorods, nanosheets and nanospheres all show two overlapping strong reduction peaks. The first reduction peak may be the conversion of MnO2 to Mn2O3, and the second reduction peak may be attributed to the reduction of Mn2O3 to MnO. These features correspond to the reduction of MnO2 with low crystallinity.41 However, some differences in the four spectra can be seen. The first reduction peak of MnO2 nanospheres appeared at lower temperature than the two other MnO2 catalysts, which might imply that the MnO2 nanospheres have better low temperature reducibility than the two other MnO2 catalysts. Furthermore, the reduction peak region of the MnO2 nanospheres is also wider than those of the two other MnO2 catalysts, which might imply that the MnO2 nanospheres have a wider operating temperature window. All the features correspond to the SCR activity results. The reduction peaks 1 and 2 of the MnO2 nanorods correspond to H2 consumptions of 8.35 and 3.58 mmol g−1, respectively. The reduction peaks 1 and 2 of the MnO2 nanosheets correspond to H2 consumptions of 6.01 and 4.55 mmol g−1, respectively. Meanwhile, we find that the reduction peaks 1 and 2 of MnO2 nanospheres correspond to H2 consumptions of 7.22 and 4.01 mmol g−1 and are located at 277 °C and 422 °C, respectively. From the above we can come to the conclusion that the total hydrogen consumption of the samples approaches the theoretical hydrogen consumption (11.49 mmol g−1). The MnO2 samples show the first reduction step of MnO2 to Mn2O3 (peak 1), which corresponds to the reduction of Mn4+ to Mn3+.42 Therefore, we can calculate the Mn4+ total active sites of the MnO2 nanospheres, nanosheets and nanorods to be 0.835, 0.601 and 0.702 mmol, respectively. The H2 consumption of the reduction peaks of the MnO2 nanosheets are less than that of the MnO2 nanospheres, and the MnO2 nanospheres obtained a higher NO conversion efficiency than the MnO2 nanosheets (Table 3).


image file: c6ra03108k-f5.tif
Fig. 5 H2-TPR (a) and NH3-TPD (b) patterns of the MnO2 bulk nanoparticles, MnO2 nanorods, MnO2 nanosheets and MnO2 nanospheres.
Table 3 H2-TPR results of the MnO2 samples
Samples Temperature (°C) H2 consumption (mmol g−1) Mn4+ total active sites (mmol)
Peak 1 Peak 2 Peak 1 Peak 2
Nanorods 299 409 8.05 3.58 0.805
Nanosheets 309 412 6.01 4.55 0.601
Nanospheres 277 422 7.22 4.01 0.722


Temperature-programmed desorption by ammonia (NH3-TPD) experiments were performed in order to investigate the surface acidity of the catalyst. As we all know, ammonia storage capacity plays an important role in the SCR reaction system. Thus, it is necessary to research the adsorption–desorption behavior of the catalyst by NH3.43 Based on analysis in previous reports, the thermal stability of the NH4+ associated with the Brønsted acid sites was lower than the NH3 associated with the Lewis acid sites. The low-temperature desorption peak below 350 °C corresponds to the NH4+ desorbed from Brønsted acid sites and the high-temperature desorption peak above 400 °C corresponds to NH3 desorbed from Lewis acid sites.44 As shown in Fig. 5b, two obvious desorption peaks were observed for the MnO2 nanospheres. The lower temperature desorption peak of the MnO2 nanospheres at 310 °C is associated with NH4+ adsorbed on the Brønsted acid sites and the higher temperature desorption peak at 550 °C is associated with NH3 adsorbed on the Lewis acid sites. Compared with the MnO2 nanospheres, no Brønsted acid sites were seen for the MnO2 bulk nanoparticles, MnO2 nanorods and MnO2 nanosheets. Moreover, the intensity of the higher temperature desorption peak of the MnO2 nanospheres is also higher than the three other samples. As we all know, Lewis acid sites exert a major influence in SCR systems, which may be a very significant reason for the SCR reaction activity of the MnO2 nanospheres being higher than those of the MnO2 bulk nanoparticles, MnO2 nanorods and MnO2 nanosheets. All the features correspond to the SCR activity results.

As we all know, Mn4+ ions are most likely the active sites of NO oxidation to NO2 because the oxidation process often occurs from redox cycling of high and low valence state cations.45,46 Therefore, the TOF was defined as the number of NOX converted per Mn4+ ion. The catalytic activity can be seen from TOF analysis (Fig. 6a), and different TOF values of the three samples at six different temperature points was obtained, indicating that the TOF values increase with increasing temperature. Moreover, the TOF value of the MnO2 nanospheres is 0.07 h−1 at 100 °C, and 0.15 h−1 at 140 °C. Furthermore, it can be seen that the TOF values varied with the different morphology. The MnO2 nanospheres obtained the highest TOF values from 100 to 200 °C over the MnO2 nanosheets and MnO2 nanorods. Therefore, it can be concluded that abundant Mn4+ ions in the MnO2 nanospheres promoted their catalytic performance in the SCR reaction.45,46 The Arrhenius plots of NO oxidation of the three different samples are shown in Fig. 6b. Furthermore, the reaction activation energy can be calculated based on the slope of the plot. It can be seen that the activation energy of the MnO2 nanospheres (20.5 kJ mol−1) is lower than the MnO2 nanosheets (30.7 kJ mol−1) and nanorods (44.5 kJ mol−1), indicating that surface adsorbed oxygen species are easily activated and favour enhanced SCR performance, which also corresponds to the SCR results.


image file: c6ra03108k-f6.tif
Fig. 6 TOF values (a) and Arrhenius plots of the NO oxidation rates (b) of the MnO2 nanorods, MnO2 nanosheets and MnO2 nanospheres.

3.2. NH3-SCR activity

Stability tests and H2O resistance studies of the three different catalysts in the SCR system were performed, and the results are shown in Fig. 7. Fig. 7a showed the SCR by ammonia performance of the MnO2 with different structures and the traditional V2O5–WO3/TiO2 catalyst. It is seen that the NO conversion of all of the catalysts increases at first and finally decreases as the temperature is elevated. However, some inherent differences in the catalyst performance are discovered. It can be seen that the MnO2 nanospheres showed about 100% NOX conversion at 200–300 °C and obtained over 90% NOX conversion at 150–350 °C with a wide operating temperature window. The MnO2 nanosheets obtained an almost 100% conversion efficiency only at 200 °C, with a narrow operating temperature window. The MnO2 nanorods were also observed to show very low activity and a best NO conversion of only 80% at 300 °C in the investigated temperature window in the SCR by NH3. In contrast, the MnO2 bulk nanoparticles obtained the lowest activity in the investigated temperature window, and the traditional V2O5–WO3/TiO2 catalyst displays poor low temperature SCR activity. Comparing all of the above catalysts, MnO2 nanosphere catalysts display a wider operating temperature window and the highest conversion efficiency over the other catalysts. As shown in Fig. 7b, the NO conversion of the three different catalysts showed little change during a 72 h continuous running duration, indicating that all these catalysts have excellent stability. Compared to the samples before the SCR reaction, there is almost no change of the morphology, surface area and crystallinity of the reacted samples, which was proved by TEM, SEM, XRD and BET (Fig. S1–S3). Therefore, after the reaction the three MnO2 materials still maintain the original structure, which proves that the three MnO2 morphologies are still stable after the SCR reaction.
image file: c6ra03108k-f7.tif
Fig. 7 NO conversion (a), stability test (b), H2O resistance (c), effect of SO2 (d) and co-effect of SO2 and H2O (e) on NO conversion for the MnO2 nanorods, nanosheets and nanospheres at the optimum temperature. Reaction conditions: 500 ppm NO, 500 ppm NH3, 3% O2, 5% H2O (when used), 5% SO2 (when used) and N2 balance gas.

Fig. 7c indicates the effects of H2O on NH3-SCR activity. After adding water, it can be seen that the NO conversion efficiency of the MnO2 nanospheres decreased to 93% and then remained at 97% for the next 18 h. The NO conversion efficiency returned to 100% quickly after the removal of H2O. Moreover, the NO conversion efficiency of the MnO2 nanorods and nanosheets decreased quickly after the introduction of H2O. However the NO conversion efficiency returned quickly after cutting off the H2O. All these features indicate the recoverable activity of the samples in the NH3-SCR system. Hence, it can be trusted that the inhibition effect of H2O was possibly attributed to the competitive adsorption of H2O and NH3 onto the catalyst surface.47 Meanwhile, Fig. 7d shows the effects of SO2 on the NH3-SCR activity. The NO conversion of the MnO2 nanospheres decreased to 87% with the introduction of SO2 and was maintained at 87%. After removing the SO2 gas, the NO conversion returned to 95%, suggesting that the MnO2 nanospheres have a good SO2 resistance in the NH3-SCR reaction. The NO conversion of the MnO2 nanorods and nanosheets decreased rapidly after the introduction of SO2. On removing the SO2 gas, the NO conversion of the MnO2 nanosheet does not fully recover, suggesting that manganese sulfate formed on the catalyst surface.48 Compared with the MnO2 nanorods and nanosheets, the MnO2 nanospheres obtained the highest SO2 resistance in the NH3-SCR reaction. Meanwhile, the co-effect of SO2 and H2O was also studied. When SO2 and H2O were introduced into the feed gases at the same time, the NOX conversion ratios of the three MnO2 samples all decreased, but the NOX conversion of the MnO2 nanospheres was still 70%. The above results suggest that the MnO2 nanospheres had good SO2/H2O durability. Additionally, N2 selectivity tests are shown in Fig. S5. As can be seen, the N2 selectivity decreases and N2O concentration increases for all the catalyst with the elevating temperature. Moreover, the MnO2 samples can obtain higher N2 selectivity at low temperature (<200 °C). However, the N2 selectivity became lower in the high temperature SCR reaction.

It is now widely accepted that the enhancement of NO to NO2 over catalysts in the SCR reaction could significantly promote the catalytic activity, and can be attributed to the faster reaction: NO + NO2 + 2NH3 → 3H2O + 2N2.49,50 Therefore, the experiments of NO to NO2 in the absence of NH3 on the catalyst are studied. As shown in Fig. 8, we can see that the oxidation of NO to NO2 increases at first and finally decreases with the elevating temperature. Moreover, the oxidation of NO to NO2 reaches the maximum conversion at 350 °C, and then decreases above 350 °C, which is attributed to the thermal equilibrium compositions. Meanwhile, the MnO2 nanospheres obtained a higher activity of oxidation of NO to NO2 than the MnO2 nanosheets and nanorods. The formation of NH4NO2 from NO2 and NH3 in the SCR reaction has been reported.50,51 Therefore, a high NO2 concentration, or a high activity of oxidation of NO to NO2, enhance the SCR activity.


image file: c6ra03108k-f8.tif
Fig. 8 Oxidation of NO to NO2 over the MnO2 samples at different temperatures.

3.3. In situ DRIFTS

The acid site distribution and the gas molecule adsorption behavior on the surface were studied by in situ DRIFTS experiments. At first the catalysts were purged with N2 for 1 h and cooled down to 50 °C. NH3 was then introduced into the IR cell and spectra were recorded as a function of temperature. The in situ FTIR spectra of adsorbed NH3 on the three different shaped MnO2 nanomaterials at different temperatures are shown in Fig. 9. With the increase of temperature, the intensities of all the bands decreased. Several bands at 1387, 1300, 1143, 1106 and 993 cm−1 were observed on the MnO2 nanorods catalyst in Fig. 9a. The bands at 1300, 1143 and 1106 cm−1 could be ascribed to coordinated NH3 Lewis acid sites.52–54 The weak band at 1387 cm−1 may be due to the oxidization species of adsorbed ammonia species.35,55 The band at 933 cm−1 was assigned to gas or weakly adsorbed NH3.35 As shown in Fig. 9b, the NH3 adsorption peaks of the MnO2 nanosheets are at 1450, 1141, 1110 and 1082 cm−1, of which the weak peak at 1450 cm−1 can be assigned to the NH4+ bound to the Brønsted acid sites, and the adsorption peaks at 1141, 1110 and 1082 cm−1 can be ascribed to the NH3 bound to the Lewis acid sites.56,57 Several bands at 1642, 1395, 1137 and 1113 cm−1 can be found for the MnO2 nanospheres in Fig. 9c. The adsorption peaks at 1642 and 1395 cm−1 were ascribed to the NH4+ bound to the Brønsted acid sites. Two bands at 1113 and 1137 cm−1 were also found, which could be ascribed to coordinated NH3 Lewis acid sites. It is evident that the band intensity of Lewis acid sites over the MnO2 nanospheres is higher than that over the MnO2 nanorods and nanosheets. All the characteristics corresponded to the above analysis in NH3-TPD.
image file: c6ra03108k-f9.tif
Fig. 9 In situ DRIFTS of (a) NH3 adsorption on the MnO2 nanorods; (b) NH3 adsorption on the MnO2 nanosheets; (c) NH3 adsorption on the MnO2 nanospheres.

The in situ DRIFTS of NO + O2 adsorption over the three different shaped MnO2 catalysts at different temperatures are shown in Fig. 10. First of all the catalysts were purged with N2 for 1 h and cooled down to 50 °C. NO + O2 was then introduced into the IR cell, and spectra were recorded as a function of temperature. As shown in Fig. 10a, the band at 1379 cm−1 could be ascribed to the adsorbed NO3, which is due to the disproportionation of chemisorbed NO2 on the sample surface.58 The bands at 1559, 1312 and 1209 cm−1 be assigned to bridged nitrate, bidentate nitrates and bridging nitrites, respectively.57 As shown in Fig. 10b, three adsorption bands were observed over the MnO2 nanosheets catalyst. Besides the bands assigned to monodentate nitrate (1499 cm−1), trans-N2O22− (1461 cm−1) and M–NO2 nitro compounds (1375 cm−1) were also observed.53–55,59–61 Several bands at 1568, 1541, 1499, 1385, and 1215 cm−1 were observed on the MnO2 nanospheres catalyst shown in Fig. 10c. The bands at 1541 cm−1 and 1499 cm−1 could be attributed to the monodentate nitrate species and the bands at 1215 cm−1 could be assigned to the disproportionation of NO bridging nitrates.54,55,62 The band at 1568 cm−1 is in accordance with the literature and can be assigned to the bidentate nitrate.52,63,64 The band at 1385 cm−1 is close to the value in the previous study and could be attributed to monodentate nitrite.57 Remarkably, the bands of the MnO2 nanospheres are much more than those of the MnO2 nanorods and nanosheets. The above in situ DRIFTS analyses show that the larger pore volume and hierarchical porous structure have enhanced the chemisorption and activation of reactant molecules on the MnO2 nanospheres, resulting in the best NO conversion efficiency in NH3-SCR.


image file: c6ra03108k-f10.tif
Fig. 10 In situ DRIFTS of NO + O2 adsorption on the MnO2 nanorods (a), NO + O2 adsorption on the MnO2 nanosheets (b); NO + O2 adsorption on the MnO2 nanospheres (c).

3.4. Reaction mechanism analysis

Herein, γ-MnO2 nanospheres, γ-MnO2 nanosheets, and β-MnO2 nanorods were found to display different catalytic effects. Furthermore, it is found that the γ-MnO2 nanospheres and γ-MnO2 nanosheets obtained significantly higher NO conversions than the β-MnO2 nanorods and β-MnO2 bulk nanoparticles. As described in the above sections, the XPS results show that γ-MnO2 has more surface chemisorbed oxygen than β-MnO2. According to the literature,24,38 the surface chemisorbed oxygen on γ-MnO2 participates easier in the reaction than β-MnO2, because the surface chemisorbed oxygen on γ-MnO2 is more active. All the above analysis implies that surface chemisorbed oxygen is the main factor in promoting the activation of NO and NH3. Based on the above analysis and discussion of the XPS measurements, the MnO2 nanorods have a relatively high Mn4+ ion content (61.59%) but the lowest surface chemisorbed oxygen (36.73%) on the surface, and the MnO2 nanosheets have abundant surface chemisorbed oxygen (47.9%) but the lowest Mn4+ ion content (35.18%) on the surface. Generally, the MnO2 nanospheres have the highest Mn4+ ion content (78.35%) and higher surface chemisorbed oxygen (43.72%). As we all know, Mn4+ ions and surface chemisorbed oxygen play important roles in the low-temperature SCR reaction. Therefore a high Mn4+/Mnn+ ratio and Oα/(Oα + Oβ) ratio will be conducive to the low-temperature SCR reaction. As shown in Fig. 11, the first oxidation process of the MnO2 nanocatalysts is NO to NO2, which is consistent with the reduction of Mn4+ to Mn2+. Furthermore, the TPD and in situ DRIFTS results show that the MnO2 nanospheres have more acid sites on the outside surface than the MnO2 nanosheets and MnO2 nanorods. Meanwhile, the MnO2 nanospheres show a higher catalytic effect than the MnO2 nanosheets. The MnO2 nanospheres have more exposed Mn4+ and acid sites due to their three dimensional structure. However, the MnO2 nanosheets with a two dimensional structure are easy to pile up into pieces, which may result in fewer Mn4+ and acid sites on the outside surface. All the features correspond to the above results of in situ DRIFTS, TPD and XPS. Moreover, the MnO2 nanospheres obtained higher TOF values from 100 to 200 °C than both the MnO2 nanosheets and MnO2 nanorods. Therefore, it can be concluded that the abundant Mn4+ ions in the MnO2 nanospheres promote the catalytic performance in the SCR reaction.45,46 Meanwhile, it can be found that activation energy of the MnO2 nanospheres (20.5 kJ mol−1) is lower than the MnO2 nanosheets (30.7 kJ mol−1) and nanorods (44.5 kJ mol−1), indicating that surface adsorbed oxygen species are easily activated and in favour of SCR performance. This result is consistent with the SCR results. Therefore, surface chemisorbed oxygen, Mn4+ and acid sites on the outside surface play important roles in the SCR reaction.
image file: c6ra03108k-f11.tif
Fig. 11 Reaction mechanism of NH3-SCR on MnO2 with different morphologies.

4. Conclusions

In this paper, MnO2 nanomaterials with different morphologies were prepared by the redox hydrothermal method, such as nanorods, nanosheets, nanospheres and bulk particles. Moreover, their NO conversion efficiency was studied and compared through the SCR reaction. The γ-MnO2 nanospheres showed significantly higher NO conversion and a significantly higher rate constant with respect to NO conversion than the bulk particles, MnO2 nanorods and MnO2 nanospheres in the NH3-SCR of NO. The NO conversion efficiency of the MnO2 nanospheres reached 100% over a wide temperature window from 200 to 350 °C. Moreover, the MnO2 nanospheres show excellent stability, H2O resistance and SO2 tolerance. The high NO conversion efficiency of the MnO2 nanospheres is due to their large specific area, a large number of surface chemisorbed oxygen species, more Mn4+ on the outside surface, higher reducibility and more acid sites, based on XPS, H2-TPR, NH3-TPD and in situ DRIFTS.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 21377061, 81270041), Independent Innovation fund of Tianjin University (2015XRG0020), Key Laboratory of Colloid and Interface Chemistry (Shandong University, Ministry of Education) (201401), and by the Natural Science Foundation of Tianjin (Grant No. 15JCYBJC48400 and 15JCZDJC41200).

Notes and references

  1. L. Zhang, L. L. Li, Y. Cao, X. J. Yao, C. Y. Ge, F. Gao, Y. Deng, C. J. Tang and L. Dong, Appl. Catal., B, 2015, 165, 589–598 CrossRef CAS.
  2. S. J. Yang, C. X. Liu, H. Z. Chang, L. Ma, Z. Qu, N. Q. Yan, C. Z. Wang and J. H. Li, Ind. Eng. Chem. Res., 2013, 52, 5601–5610 CrossRef CAS.
  3. L. Ma, Y. S. Cheng, G. Cavataio, R. W. McCabe, L. X. Fu and J. H. Li, Appl. Catal., B, 2014, 156–157, 428–437 CrossRef CAS.
  4. S. Brandenberger, O. Kröcher and A. T. R. Althoff, Catal. Rev. Sci. Eng., 2008, 50(4), 492–531 CAS.
  5. R. D. Zhang, N. Liu, Z. G. Lei and B. H. Chen, Chem. Rev., 2016, 116(6), 3658–3721 CrossRef CAS PubMed.
  6. T. C. Bruggemann and F. J. Keil, J. Phys. Chem. C, 2011, 115, 23854–23870 Search PubMed.
  7. P. Maitarad, D. S. Zhang, R. H. Gao, L. Y. Shi, H. R. Li, L. Huang, T. Rungrotmongkol and J. P. Zhang, J. Phys. Chem. C, 2013, 117, 9999–10006 CAS.
  8. S. X. Cai, D. S. Zhang, L. Y. Shi, J. Xu, L. Zhang, L. Huang, H. R. Li and J. P. Zhang, Environ. Sci. Technol., 2014, 48, 10354–10362 CrossRef PubMed.
  9. S. J. Yang, S. C. Xiong, Y. Liao, X. Xiao, F. H. Qi, Y. Peng, Y. W. Fu, W. P. Shan and J. H. Li, Nanoscale, 2014, 6, 7346 RSC.
  10. Z. M. Liu, S. X. Zhang, J. H. Li, J. Z. Zhu and L. L. Ma, Appl. Catal., B, 2014, 158–159, 11–19 CrossRef CAS.
  11. Z. H. Chen, F. R. Wang, H. Li, Q. Yang, L. F. Wang and X. H. Li, Ind. Eng. Chem. Res., 2012, 51, 202–212 CrossRef CAS.
  12. B. Thirupathi and P. G. Smirniotis, J. Catal., 2012, 288, 74–83 CrossRef CAS.
  13. E. Saputra, S. Muhammad, H. Q. Sun, H. Ming, M. O. Tadé and S. B. Wang, Appl. Catal., B, 2014, 154–155, 246–251 CrossRef CAS.
  14. Y. J. Wei, Y. Sun, W. Sub and J. Liu, RSC Adv., 2015, 5, 26231–26235 RSC.
  15. S. J. Yang, Y. Liao, S. C. Xiong, F. H. Qi, H. Dang, X. Xiao and J. H. Li, J. Phys. Chem. C, 2014, 118, 21500–21508 CAS.
  16. A. Patel, P. Shukla, J. L. Chen, T. Rufford, V. Rudolph and Z. H. Zhu, Catal. Today, 2012, 212, 38–44 CrossRef.
  17. D. Fang, J. L. Xie, D. Mei, Y. M. Zhang, F. He, X. Q. Liu and Y. M. Li, RSC Adv., 2014, 4, 25540–25551 RSC.
  18. R. H. Gao, D. S. Zhang, P. Maitarad, L. Y. Shi, T. Rungro, H. R. Li, J. P. Zhang and W. G. Cao, J. Phys. Chem. C, 2013, 117, 10502–10511 CAS.
  19. K. Zhuang, J. Qiu, F. S. Tang, B. L. Xu and Y. N. Fan, J. Phys. Chem. C, 2011, 13, 4463 CAS.
  20. S. J. Yang, Y. W. Fu, Y. Liao, S. C. Xiong, Z. Qu, N. Q. Yan and J. H. Li, Catal. Sci. Technol., 2014, 4, 224–232 CAS.
  21. S. H. Zhan, M. Y. Qiu, S. S. Yang, D. D. Zhu, H. B. Yu and Y. Li, J. Mater. Chem. A, 2014, 2, 20486–20493 CAS.
  22. Y. Liu, J. Xu, H. R. Li, S. X. Cai, H. Hu, C. Fang, L. Y. Shi and D. S. Zhang, J. Mater. Chem. A, 2015, 3, 11543–11553 CAS.
  23. B. Meng, Z. B. Zhao, X. Z. Wang, J. J. Liang and J. S. Qiu, Appl. Catal., B, 2013, 129, 491–500 CrossRef CAS.
  24. B. Y. Bai, J. H. Li and J. M. Hao, Appl. Catal., B, 2015, 164, 241–250 CrossRef CAS.
  25. Z. C. Bai, B. Sun, N. Fan, Z. C. Ju, M. H. Li, L. Q Xu and Y. T. Qian, Chem.–Eur. J., 2012, 18, 5319–5324 CrossRef CAS PubMed.
  26. J. F. Li, B. J. Xi, Y. C. Zhu, Q. W. Li, Y. Yan and Y. T. Qian, J. Alloys Compd., 2011, 509, 9542–9548 CrossRef CAS.
  27. X. B. Zhu, X. N. Li, Y. C. Zhu, S. S. Jin, Y. Wang and Y. T. Qian, Electrochim. Acta, 2014, 121, 253–257 CrossRef CAS.
  28. Y. Ren, R. Armstrong, F. Jiao and P. G. Bruce, J. Am. Chem. Soc., 2010, 132, 996–1004 CrossRef CAS PubMed.
  29. X. Gu, L. Chen, Z. C. Ju, H. Y. Xu, J. Yang and Y. T. Qian, Adv. Funct. Mater., 2013, 23, 4049–4056 CrossRef CAS.
  30. J. Yang, J. H. Zeng, S. H. Yu, L. Yang, G. E. Zhou and Y. T. Qian, Chem. Mater., 2000, 12, 3259–3263 CrossRef CAS.
  31. C. Fang, D. S. Zhang, S. X. Cai, L. Zhang, L. Huang, H. R. Li, P. Maitarad, L. Y. Shi, R. H. Gao and J. P. Zhang, Nanoscale, 2013, 5, 9199–9207 RSC.
  32. X. Y. Wang, W. Wen, J. X. Mi, X. X. Li and R. H. Wang, Appl. Catal., B, 2015, 176, 454–463 CrossRef.
  33. J. Yu, Z. C. Si, L. Chen, X. D. Wu and D. Weng, Appl. Catal., B, 2015, 163, 223–232 CrossRef CAS.
  34. L. Zhang, D. S. Zhang, J. P. Zhang, S. X. Cai, C. Fang, L. Huang, H. R. Li, R. H. Gao and L. Y. Shi, Nanoscale, 2013, 5, 9821–9828 RSC.
  35. L. Zhang, L. Y. Shi, L. Huang, J. P. Zhang, R. H. Gao and D. S. Zhang, ACS Catal., 2014, 4, 1753–1763 CrossRef CAS.
  36. X. Xiao, Z. Y. Sheng, L. Yang and F. Dong, Catal. Sci. Technol., 2016, 6, 1507–1514 CAS.
  37. B. H. Zhao, R. Ran, X. D. Wu and D. Weng, Appl. Catal., A, 2016, 514, 24–34 CrossRef CAS.
  38. X. Wang, Y. Y. Zheng, Z. Xu, X. L. Wang and X. P. Chen, RSC Adv., 2013, 3, 11539–11542 RSC.
  39. X. F. Tang, J. H. Li and J. M. Hao, Catal. Commun., 2010, 11, 871–875 CrossRef CAS.
  40. F. D. Liu, W. P. Shan, Z. H. Lian, L. J. Xie, W. W. Yang and H. He, Environ. Sci. Technol., 2013, 3, 2699–2707 CAS.
  41. X. F. Tang, J. H. Li, L. Sun and J. M. Hao, Appl. Catal., B, 2010, 99, 156–162 CrossRef CAS.
  42. J. H. Chen, M. Q. Shen, X. Q. Wang, G. S. Qi, J. Wang and W. Li, Appl. Catal., B, 2013, 134–135, 251–257 CrossRef CAS.
  43. J. L. Zuo, Z. H. Chen, F. R. Wang, Y. H. Yu, L. F. Wang and X. H. Li, Ind. Eng. Chem. Res., 2014, 53, 2647–2655 CrossRef CAS.
  44. D. Wang, Y. Jangjou, Y. Liu, M. K. Sharma, J. Y. Luo, J. H. Li, K. Kamasamudram and W. S. Epling, Appl. Catal., B, 2015, 165, 438–445 CrossRef CAS.
  45. D. S. Zhang, L. Zhang, L. Y. Shi, C. Fang, H. R. Li, R. H. Gao, L. Huang and J. P. Zhang, Nanoscale, 2013, 5, 1127–1136 RSC.
  46. I. Ascoop, V. V. Galvita, K. Alexopoulos, M. F. Reyniers, V. D. Voort, V. Bliznuk and G. B. Marin, J. Catal., 2016, 335, 1–10 CrossRef CAS.
  47. R. Y. Qu, X. Gao, K. F. Cen and J. H. Li, Appl. Catal., B, 2013, 142, 290–297 CrossRef.
  48. D. S. Zhang, L. Zhang, L. Y. Shi, C. Fang, H. R. Li, R. H. Gao, L. Huang and J. P. Zhang, Nanoscale, 2013, 5, 1127 RSC.
  49. Z. H. Chen, Q. Yang, H. Li, X. H. Li, L. F. Wang and S. C. Tsang, J. Catal., 2010, 276, 56–65 CrossRef CAS.
  50. B. Guan, H. Lin, L. Zhu, B. Tian and Z. Huang, Chem. Eng. J., 2012, 181–182, 307–322 CrossRef CAS.
  51. H. Hu, S. X. Cai, H. R. Li, L. Huang, L. Y. Shi and D. S. Zhang, ACS Catal., 2015, 5, 6069–6077 CrossRef CAS.
  52. Z. B. Wu, B. Q. Jiang, Y. Liu, H. Q. Wang and R. B. Jin, Environ. Sci. Technol., 2007, 41, 5812–5817 CrossRef CAS PubMed.
  53. Z. M. Liu, J. Z. Zhu, J. H. Li, L. L. Ma and S. I. Woo, ACS Appl. Mater. Interfaces, 2014, 6, 14500–14508 CAS.
  54. Z. M. Liu, Y. Li, T. L. Zhu, H. Su and J. Z. Zhu, Ind. Eng. Chem. Res., 2014, 53, 12964–12970 CrossRef CAS.
  55. Y. Liu, T. T. Gu, X. L. Weng, Y. Wang, Z. B. Wu and H. Q. Wang, J. Phys. Chem. C, 2012, 116, 16582–16592 CAS.
  56. G. S. Qi and T. Yang, J. Phys. Chem. B, 2004, 108, 15738–15747 CrossRef CAS.
  57. L. Chen, J. H. Li and M. F. Ge, J. Phys. Chem. C, 2009, 113, 21177–21184 CAS.
  58. Y. L. Wang, C. Z. Ge, L. Zhan, C. Li, W. M. Qiao and L. C. Ling, Ind. Eng. Chem. Res., 2012, 51, 11667–11673 CrossRef CAS.
  59. L. Chen, J. J. Li and M. F. Ge, Environ. Sci. Technol., 2010, 44, 9590–9596 CrossRef CAS PubMed.
  60. M. Y. Qiu, S. H. Zhan, H. B. Yu, D. D. Zhu and S. Q. Wang, Nanoscale, 2015, 7, 2568–2577 RSC.
  61. S. J. Yang, J. H. Li, C. Z. Wang, J. H. Chen, L. Ma, H. Z. Chang, L. Chen, Y. Peng and N. Q. Yan, Appl. Catal., B, 2012, 44, 73–80 CrossRef.
  62. B. Guan, H. Lin, L. Zhu and Z. Huang, J. Phys. Chem. C, 2011, 115, 12850–12863 CAS.
  63. Z. C. Si, D. Weng, X. D. Wu, R. Ran and Z. R. Ma, Catal. Commun., 2012, 17, 146–149 CrossRef CAS.
  64. Z. C. Si, D. Weng, X. D. Wu, Z. R. Ma, J. Ma and R. Ran, Catal. Today, 2013, 201, 122–130 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra03108k

This journal is © The Royal Society of Chemistry 2016