Next Article in Journal
Photophysics and Photochemistry of Iron Carbene Complexes for Solar Energy Conversion and Photocatalysis
Previous Article in Journal
Oxidative Bio-Desulfurization by Nanostructured Peroxidase Mediator System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Role of Nitrate on the Sol-Gel Spread Self-Combustion Process and Its Effect on the NH3-SCR Activity of Magnetic Iron-Based Catalyst

1
School of Energy and Power Engineering, University of Shanghai for Science & Technology, Jungong Road #516, Shanghai 200093, China
2
Shanghai Key Laboratory of Multiphase Flow and Heat Transfer in Power Engineering, University of Shanghai for Science & Technology, Jungong Road #516, Shanghai 200093, China
3
SPIC Powder Plant Operation Technology (Beijing) Co., Ltd., Future Technology City, Beijing 102209, China
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(3), 314; https://doi.org/10.3390/catal10030314
Submission received: 25 February 2020 / Revised: 6 March 2020 / Accepted: 6 March 2020 / Published: 10 March 2020

Abstract

:
Sol-gel spread self-combustion is the burning of the complexing agent in dried gel and the oxidant. Meanwhile, high temperature takes place during the combustion process, which is harmful to the pore structure of the catalyst. The nitrate from metal nitrate precursors as an oxidant could participate in the spread of the self-combustion process. Therefore, the influence of nitrate from metal nitrate on the spread self-combustion of an iron–cerium–tungsten citric acid gel and its catalytic performance of NOx reduction were investigated by removing nitrate via the dissolution of washing co-precipitation with citric acid and re-introducing nitric acid into the former solution. It was found that the removal of nitrate contributes to enhancing the NH3–SCR activity of the magnetic mixed oxide catalyst. The NOx reduction efficiency was close to 100% for Fe85Ce10W5–CP–CA at 250 °C while the highest was only 80% for the others. The results of thermal analysis demonstrate that the spread self-combustion process of citric acid dried gel is enhanced by re-introducing nitric acid into the citric acid dissolved solution when compared with the removal of nitrate. In addition, the removal of nitrate helps in the formation of γ-Fe2O3 crystallite in the catalyst, refining the particle size of the catalyst and increasing its pore volume. The removal of nitrate also contributes to the formation of Lewis acid sites and Brønsted acid sites on the surface of the catalyst compared with the re-introduction of nitric acid. The in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) demonstrates that both Eley–Rideal (E–R) and Langmuir–Hinshelwood (L–H) mechanisms exist over Fe85Ce10W5–CP–CA at 250 °C with E–R as its main mechanism.

1. Introduction

Nitrogen oxide (NOx) emitted from coal-fired power plants and automobile engines has a strong negative influence on the environment and human health [1,2,3,4,5,6]. Selective catalytic reduction of NOx with NH3 (NH3–SCR) is well known as the best available control technology (BACT) to reduce nitrogen oxides due to its high efficiency [7,8]. Meanwhile, there exists some drawbacks such as high cost, high-temperature conversion of SO2 to SO3, the toxicity, and volatility of vanadium species for the commercial vanadium-based catalyst [9,10,11,12,13,14,15,16]. Due to the non-toxicity, low cost, environment, and the outstanding redox ability between FeIII and FeII, a series of iron-based mixed oxide catalysts prepared through co-precipitation, sol-gel, and impregnation methods have been developed by many researchers [17,18,19,20]. However, the sol-gel spread self-combustion method, also called low-temperature combustion synthesis (LCS), which takes advantage of organics (citric acid, glucose, urea, and so on) as reactants and nitrates from metal nitrates as oxidants, has been widely used to obtain nano-particles, ultrafine powders, and metal oxide catalysts [21,22,23,24,25,26]. In our precious studies, a novel Fe–Ce–W mixed oxide catalyst synthesized via the citric acid sol-gel spread self-combustion method exhibited high catalytic activity of NOx reduction with high-dispersive γ-Fe2O3 crystallite formed in it [27]. However, a large amount of heat is released during the spread self-combustion of the citric acid dried gel, and brings about a high temperature that acts on the obtained composite oxide catalyst, which might destroy the physical structure of catalyst, thereby influencing its catalytic performance of NOx reduction even though the duration of high temperature is short. Previous research has demonstrated that the amount of complexing agent as the fuel participated in the spread self-combustion of the dried gel, and affected the rapid oxidation of FeII to FeIII [28]. Meanwhile, the oxygen from the surrounding air as an oxidant showed an enhancement effect on the spread self-combustion process of dried gel, which affected the structure properties of the obtained powder [29]. The nitrate from metal nitrate precursors as another oxidant could also participate in the spread self-combustion of dried gel. Therefore, it is necessary to investigate the influence of nitrate on the spread self-combustion of dried gel, especially its effect on the NH3–SCR activity over the above magnetic iron–cerium–tungsten mixed oxide catalyst.
Herein, in the present work, two kinds of magnetic iron–cerium–tungsten mixed oxide catalysts were synthesized via the spread self-combustion of citric acid gel by removing nitrate through the dissolution of washing co-precipitation with citric acid, and re-introducing nitric acid into the former citric acid dissolved solution, respectively. Thermo-gravimetric analysis (TG-DTG-DSC) was used to study the influence of nitrate on the combustion of citric acid gel. In addition, x-ray diffraction (XRD), N2-adsorption–desorption, x-ray photoelectron spectroscopy (XPS), temperature-programmed reduction (H2-TPR), temperature-programmed desorption (NH3-TPD), and scanning electron microscope (SEM) were eventually used to characterize the physical structural and chemical properties of the catalyst.

2. Results and Discussion

2.1. NH3–SCR Activity

As can be observed from Figure 1, magnetic Fe85Ce10W5–CP–CA synthesized through the spread self-combustion of the citric acid dried gel without the nitrate from the metal nitrate precursors exhibited excellent catalytic performance of NOx reduction at 150~400 °C, and more than 90% of NOx reduction was achieved at 225~400 °C and over, under a gaseous hourly space velocity (GHSV) of 60,000/h. Meanwhile, the re-introduction of nitric acid into the citric acid dissolved solution decreased the NH3–SCR activity of Fe85Ce10W5–CP–CA, and the enhancement of nitrate ions from 0.5 to 2.0 further decreased its catalytic performance. This demonstrates that the nitrate brought from the precursors of metal nitrate shows an inhibition on the NH3–SCR activity of the magnetic Fe85Ce10W5 catalyst prepared through the spread self-combustion of citric acid gel. As shown in Table 1, the NOx conversion over per gram of Fe85Ce10W5–CP–CA at low temperature (125~200 °C) in one hour was still higher than those of Fe85Ce10W5–CP–CA(NA1.0), although the bulk density of Fe85Ce10W5–CP–CA was 0.7124 g/mL, which was higher than that of Fe85Ce10W5–CP–CA(NA1.0). Apparently, the re-introduction of nitric acid could participate in the spread self-combustion of citric acid dried gel, thereby decreasing the bulk density of Fe85Ce10W5–CP–CA. Therefore, the nitrate brought from the metal nitrate precursors shows an important role in the spread self-combustion of citric acid gel, thus affecting the physical structure and the redox properties of magnetic iron–cerium–tungsten mixed oxide catalyst. Herein, TG-DTG-DSC was used to investigate the combustion of the critic acid dried gels of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0).

2.2. Thermo-Gravimetric Analysis (TG-DTG-DSC)

Thermal analysis was carried out to investigate the relationship between the weight loss of the catalyst and temperature. The thermo-gravimetric (TG), differential thermo-gravimetric (DTG), and differential scanning calorimetry (DSC) traces of the Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) citric acid gels were tested and the results are shown in Figure 2. The gel of Fe85Ce10W5–CP–CA(NA1.0) illustrates about 20% weight loss before approximately 140 °C, and a larger exothermic peak can be clearly observed from its DSC, which is mainly attributed to the decomposition of nitrate. However, the gel of Fe85Ce10W5–CP–CA shows two steps weight loss compared to the three steps weight loss of Fe85Ce10W5–CP–CA(NA1.0) gel. During the ignition, the gel of Fe85Ce10W5–CP–CA(NA1.0) presents a sharply spread self-combustion at about 140 °C with a large quantity of reddish brown gas released due to the decomposition of nitrate. Meanwhile, there exists an exothermic peak at 110~160 °C for Fe85Ce10W5–CP–CA(NA1.0) gel, or not an endothermic peak. This indicates that there exists a certain burning of citric acid at 110~160 °C, and this also enhances the main burning temperature of dried gel compared with that of Fe85Ce10W5–CP–CA with a major mass loss about 37% at 160~220 °C. Therefore, the presence of NO3 helps with the burning or/and decomposing of the citric acid dried gel [30]. Finally, the exothermic peak at 240~350 °C with a smaller weight loss could be considered as the gradual decomposition of citrates (ferric citrate, etc.) for these two dried gels. Meanwhile, the decomposition temperature of the gradual citrates for Fe85Ce10W5–CP–CA is lower than those of Fe85Ce10W5–CP–CA(NA1.0). Thus, it can be concluded that the existence of nitrate could enhance the spread self-combustion of citric acid gel, which affects the physical structure and redox properties of magnetic iron–cerium–tungsten mixed oxide catalyst, thereby influencing its low-temperature NH3–SCR activity.

2.3. Structural Properties

2.3.1. X-Ray Diffraction

The X-ray diffraction patterns of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) were measured and the results are shown in Figure 3. It can be noted that there exists some obvious sharp diffraction peaks at 2θ = 30.2°, 35.6°, 43.3°, 53.7°, 57.4°, 62.7° in the samples, which are attributed to γ-Fe2O3 crystallite (#25-1402), and the diffraction peak at 2θ = 28.9° could be assigned to CeO2 crystallite (#43-1002), according to the Joint Committee on Powder Diffraction Standards (JCPDS). This indicates that γ-Fe2O3 and CeO2 are the main crystallites of magnetic Fe–Ce–W mixed oxide catalysts [19,31]. Meanwhile, the intensity of diffraction peaks attributed to the γ-Fe2O3 crystallite in Fe85Ce10W5–CP–CA is stronger than that of Fe85Ce10W5–CP–CA(NA1.0). Therefore, the removal of nitrate promotes the formation of γ-Fe2O3 crystallite, which is usually thought to be an important active specie for NH3–SCR reaction [27]. The re-introduction of nitrate also weakens the crystallization of CeO2 in Fe85Ce10W5–CP–CA. Meanwhile, the average γ-Fe2O3 crystallite sizes of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP-CA(NA1.0) as calculated according to the Scherrer equation were 13.0 and 14.4 nm, respectively. Therefore, the re-introduction of nitrate causes a large amount of combustion heat to be released and results in a high temperature, thereby enlarging the particle size of Fe85Ce10W5–CP–CA. Then, the removal of nitrate helps in the formation of dispersive γ-Fe2O3 crystallites during the spread self-combustion of citric acid gel and restrains the inter-particle agglomeration and growth of the magnetic iron–cerium–tungsten mixed oxide catalyst.

2.3.2. Scanning Electron Microscopy (SEM)

Scanning electron microscopy (SEM) is a useful technique to study the surface morphology, shape, and macroscopic particle size of the catalyst. Typical SEM pictures of Fe85Ce10W5–CP–CA, Fe85Ce10W5–CP–CA(NA1.0), and their precursors before being ignited are shown in Figure 4. As shown in Figure 4A,B, the precursor of Fe85Ce10W5–CP–CA(NA1.0) shows stronger agglomeration than that of Fe85Ce10W5–CP–CA, and the combustion of dried gel at the presence of nitrate contributes to the agglomeration of the particles, and results in poor pore connectivity of Fe85Ce10W5–CP–CA(NA1.0), thus decreasing its pore volume. The particles in Fe85Ce10W5–CP–CA have an outstanding distribution, contributing to the smaller pores and larger pore volume, which is beneficial to mass transfer and diffusion. Meanwhile, the diffusion of reactant gases and product gases among the pores of catalysts is important for NH3–SCR reaction [31,32,33]. In addition, the removal of nitrate also decreases the particles’ average diameter of Fe85Ce10W5–CP–CA as calculated from the SEM in Figure 4C,D, which is in accordance with the particle sizes calculated according to the Scherrer equation.

2.3.3. N2 Adsorption–Desorption

The porosity and the pore size distribution of the as-prepared two catalysts were determined using N2 adsorption–desorption. Figure 5 displays the N2 adsorption–desorption isotherms, the pore size distributions of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0), and their NOx conversions over per surface area in one hour (mg/(m2·h)). As can be observed, the isotherm of Fe85Ce10W5–CP–CA(NA1.0) can be recognized as a type IV N2 adsorption/desorption isotherm according to the International Union of Pure and Applied Chemistry (IUPAC) classification, and it presents mainly meso-pores (2–50 nm), however, the hysteresis loops of Fe85Ce10W5–CP–CA(NA1.0) and Fe85Ce10W5–CP–CA are the H2 and H1 type [32,34], respectively. This demonstrates that the removal of nitrate promotes the formation of meso-pores in magnetic, and the Fe85Ce10W5–CP–CA catalyst shows uniform and regular meso-pores, which was confirmed by the results of the pore diameter distribution in Figure 5B [35]. Interesting, the Brunauer-Emmett and Teller (BET) surface area of Fe85Ce10W5–CP–CA was 90.85 m2/g, a little smaller than that of Fe85Ce10W5–CP–CA(NA1.0) (93.13 m2/g), as shown in Table 2. Usually, a large BET surface area is beneficial to enhance the catalytic ability of the catalyst. Thus, the NOx conversions at low-temperature over per surface area of catalysts in one hour were calculated, and it was found that Fe85Ce10W5–CP–CA showed a higher NOx conversion over per surface area in one hour than Fe85Ce10W5–CP–CA(NA1.0) at 150~200 °C.

2.3.4. X-ray Photoelectron Spectroscopy (XPS) and H2-Temperature Program Reduction (H2-TPR)

To investigate the influence of nitrate on the elements’ concentrations and chemical states on the surface of the catalyst, the XPS spectra of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) were carried out. As can be noted from Table 3, the re-introduction of nitrate decreased the concentrations of cerium and tungsten on the surface of Fe85Ce10W5–CP–CA, and increased its surface concentration of iron. Higher surface Ce concentration contributes to the excellent reduction ability, which is widely known to be conducive for NH3–SCR reaction [36]. To further study the influence of nitrate on the redox properties of the magnetic Fe–Ce–W mixed oxides catalyst, the H2–TPR curves of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) were obtained. The results in Figure 6 show that there exists one peak at the temperature range of 300~500 °C for these two samples, attributed to the reduction from Fe2O3 to Fe3O4 at the range of 300~400 °C and the further reduction of Fe3O4 to FeO at about 500 °C [37,38,39]. Interesting, the re-introduction of nitrate enhanced the low-temperature reducibility of Fe85Ce10W5–CP–CA by increasing its surface concentration of iron species, which is confirmed by the results of XPS in Figure 5. Meanwhile, the re-introduction of nitrate decreased the NH3–SCR activity of the magnetic Fe85Ce10W5–CP–CA catalyst.

2.3.5. NH3–Temperature Programmed Desorption (NH3-TPD)

The peak position of NH3–TPD refers to the adsorption strength, and the peaks at 100~200, 200~350 and >350 °C are attributed to the weakly acidic site, the medium-strong acid site, and the strong acid site [40], respectively. From the results in Figure 7, it can be seen that there exists weak acid sites, medium acid site, and strong acid sites on the surface of the magnetic Fe–Ce–W mixed oxide catalysts. The NH3 species adsorbed on the weak acid sites and the medium acid sites were mainly assigned to the coordinated NH3 bound to Lewis acid sites and the partial ionic NH4+ bound to Brønsted acid sites [41,42,43]. The re-introduction of nitrate decreased the adsorption of NH3 on the magnetic Fe85Ce10W5–CP–CA catalyst, particularly decreasing the intensity of its weak and medium-strong acidic sites. Combined with the results of XRD and SEM, we speculate that the re-introduction of nitrate results in a severe agglomeration of amorphous iron and tungsten species of the magnetic Fe85Ce10W5–CP–CA catalyst, and its acid center could be covered or decompose in this process, thereby decreasing the intensity of its weak and medium-strong acidic sites. Therefore, the re-introduction of nitrate modifies the surface structure of the magnetic Fe85Ce10W5–CP–CA catalyst, and shows better adsorption of the NH3 reactants than Fe85Ce10W5–CP–CA(NA1.0) at 125~250 °C, which is considered to be a key factor in improving the low-temperature NH3–SCR activity of the catalyst.

2.4. Catalytic Mechanism

To identify the presence of the adsorbed NH3 species in the SCR process on the surface of Fe85Ce10W5–CP–CA, the in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) spectra of three transient experiments were recorded under a steady-state condition, and the results are presented in Figure 8. As illustrated in Figure 8A, after NH3 adsorption and N2 purge at 250 °C, Fe85Ce10W5–CP–CA showed several bands located at 1188, 1406, 1587, 3256, and 3351 cm−1. The bands located at 1188 and 1587 cm−1 can be attributed to the coordinated NH3 on the Lewis acid sites, and the band located at 1406 cm−1 can be ascribed to the ionic NH4+ bound to the Brønsted acid site, while the bands of 3256 and 3351 cm−1 corresponded to the N–H stretching modes of the coordinated NH3 connected to the Lewis acid sites [27,44,45,46,47,48]. After the introduction of NO + O2 for 10 s, the adsorption peaks of the coordinated NH3 on the Lewis acid sites (located at 1188 and 1587 cm−1) and the N–H stretching modes of coordinated NH3 connected to the Lewis acid sites (located at 3256 and 3351 cm−1) disappeared, and the intensity of ionic NH4+ bound to the Brønsted acid site at 1406 cm−1 also became weakened. At the same time, the bidentate nitrates (1002 and 1547 cm−1), M–NO2 formed by the reaction between M–OH and NOx (1350 and 3639 cm−1) and the bridging nitrate (1618 cm−1) species appeared [44,45,46,47]. As shown in Figure 8B, after nitrogen oxide species adsorption and N2 purge at 250 °C, Fe85Ce10W5–CP–CA showed several bands ascribed to M–NO2 formed by the reaction between M–OH and NOx(1353 cm−1), bidentate nitrates (1560 cm−1), and bridging nitrate (1621 cm−1), respectively [48,49,50]. After the introduction of NH3 for 10 s, the intensity of adsorption peaks of M–NO2 (1353 cm−1) and bridging nitrate (1621 cm−1) became weakened, and the band at 1560 cm−1 ascribed to the bidentate nitrates disappeared. The IR bands assigned to the coordinated NH3 on the Lewis acid sites (1189 and 1587 cm−1), the ionic NH4+ bound to the Brønsted acid site (1439 cm−1), and the N–H stretching modes of the coordinated NH3 connected to the Lewis acid sites (3255 and 3364 cm−1) appeared [27,47,48,49,50]. Figure 8C shows the experimental results of the introduction of NH3 + NO+O2 gases over Fe85Ce10W5–CP–CA at 250 °C. It can clearly be seen that the IR bands ascribed to the coordinated NH3, the N–H stretching modes of the coordinated NH3 and the ionic NH4+ appeared, and their intensity gradually became stronger with the increase in the NH3 + NO + O2 introduction. However, the adsorption peaks at about 1350 and 1562 cm−1 ascribed to M–NO2 and bidentate nitrate also appeared and then quickly vanished when NH3 + NO + O2 gases were introduced into the reaction tank. Therefore, it can be concluded that the reaction between the adsorbed NH3 species with gaseous NO + O2 or the adsorbed NOx species might occur over Fe85Ce10W5–CP–CA at 250 °C, obeyed both E–R and L–H mechanisms, and E–R was its main mechanism.

3. Experimental

3.1. Synthesis of the Catalyst

FeIII(NO3)3·9H2O, CeIII(NO3)3·6H2O, (NH4)6H2WVI12O40·nH2O, NH3·H2O, and citric acid were used as the precursors, the precipitator, and the complexing agent, respectively. For the preparation of magnetic iron–cerium–tungsten mixed oxide catalyst through the spread self-combustion of dried gel without nitrate, a certain amount of FeIII(NO3)3·9H2O, CeIII(NO3)3·6H2O, and (NH4)6H2WVI12O40·nH2O were dissolved in the de-ionized water by keeping their molar ratio at 85:10:5. After magnetic stirring evenly, this solution was titrated into ammonia water of 2 mol/L until the pH value of 9–10. The obtained precipitation was fully washed by the de-ionized water to reduce the nitrate ion in it, and then a certain amount of citric acid was added into the washed precipitation with the molar ratio of citric acid/(Fe + Ce + W) as 1.0. After being stirred for 3 h under water bath at 50 °C, the citric acid dissolved solution were treated by microwave irradiation for 10 min with 36.4% power (microwave irradiation 8 s, 14 s suspended for a cycle with full power) using a household microwave oven (EG8MEA6-NR, 2.45 GHz, 800 W) to ignite, and this catalyst is denoted as Fe85Ce10W5–CP–CA. The preparation of the catalyst by re-introducing nitric acid into the citric acid dissolved solution is similar to the preparation of Fe85Ce10W5–CP–CA, which can be expected by adding a certain amount of nitric acid into the citric acid dissolved co-precipitation precursor before the water bath, which can be denoted as Fe85Ce10W5–CP–CA (NAx, x = 0.5, 1.0, 2.0), where x is the molar ratio of nitric acid/NO3 in both Fe(NO3)3 and Ce(NO3)3.

3.2. Catalytic Measurement and Characterization

The catalytic measurements of NOx abatement with NH3 were carried out in a one-dimensional transversely fixed quartz reactor. The simulated gas consisted of 1000 ppm NH3, 1000 ppm NO, 3 vol.% O2, and the balanced N2 with the total flow of 2000 mL/min. The samples used in each experiment were 2 mL with a gas hourly space velocity (GHSV) of 60,000 h−1. The concentrations of NOx and O2 were monitored via a flue gas analyzer (Model 60i, Thermo Fisher Scientific Co. Ltd., Waltham, MA, USA). NOx conversion is defined by the following equation:
NOx   conversion   ( % )   =   C ( NOx )   inlet   C ( NOx )   outlet C ( NOx )   inlet
The thermal decomposition properties of the precursor before being ignited were determined with a thermal gravimetric analyzer (Netzsch, Selb, Germany, STA449 F3) under an air atmosphere. The surface morphology of the catalyst was measured on a SEM (Japan, Shimadzu). In addition, the physicochemical properties of the samples were also characterized via XRD, N2 adsorption–desorption, XPS, H2–TPR, and NH3–TPD as the same as our previous research [27]. The average crystallite sizes of γ-Fe2O3 in Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) were calculated according to the Scherrer equation:
D = k λ β cos ( θ )
where k is the shape factor (k = 0.9); λ is the wavelength of radiations; and β is the FWHM (full width at half maxima). Finally, the NH3–SCR mechanism over Fe85Ce10W5–CP–CA at 250 °C was acquired by in situ diffuse reflection infrared Fourier transform spectroscopy (in situ DRIFTS).

4. Conclusions

The nitrate brought from metal nitrates could participate in the spread self-combustion of citric acid dried gel. The removal of nitrate contributes to optimize the structure properties and acid sites of the magnetic Fe–Ce–W mixed oxide catalyst. The removal of nitrate helps in the formation of dispersive gamma–Fe2O3 in Fe85Ce10W5–CP–CA, accelerating the crystallite rate and refining the particle size, and it shows a wider pore size distribution than Fe85Ce10W5–CP–CA(NA1.0). At the same time, the removal of nitrate also enhances the concentrations of cerium/tungsten on the surface of the magnetic catalyst and its low-temperature adsorption of NH3. Therefore, the re-introduction of nitrate decreases the low-temperature NH3–SCR activity of Fe85Ce10W5–CP–CA, and both E–R and L–H mechanisms exist over Fe85Ce10W5–CP–CA at 250 °C.

Author Contributions

Formal analysis, S.-m.W.; Investigation, B.Y.; Resources, W.L.; Writing-original draft, X.N.; Writing-review & editing, Z.-b.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Bin Yang grant number No. 2016YFB0600601, Zhi-bo Xiong grant number No. 51406118 and Wei Lu grant number No. QD2015017. And The APC was funded by Bin Yang.

Acknowledgments

This work was supported by the National Key Research and Development Program of China (No. 2016YFB0600601), the National Science Foundation of China (No. 51406118), and the Program of Special Appointment (Eastern Scholar) at the Shanghai Institutions of Higher Learning (No. QD2015017).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Parvulescu, V.; Delmon, B.; Grange, P. Catalytic removal of NO. Catal. Today 1998, 46, 233–316. [Google Scholar] [CrossRef]
  2. Forzatti, P.; Nova, I.; Tronconi, E. Enhanced NH3 selective catalytic reduction for NOx abatement. Angew. Chem. Int. Edit. 2009, 48, 8366–8368. [Google Scholar] [CrossRef] [PubMed]
  3. Husain, D. Chemistry of atmospheres: An introduction to the chemistry of atmospheres of earth the planets and their satellites by R. P. Wayne. J. Photo. Chem. Photobiol. A Chem. 1992, 63, 253–254. [Google Scholar] [CrossRef]
  4. Zhao, S.J.; Wang, L.; Wang, Y.; Li, X. Hierarchically porous LaFeO3 perovskite prepared from the pomelo peel bio-template for catalytic oxidation of NO. J. Phys. Chem. Solids 2018, 116, 43–49. [Google Scholar] [CrossRef]
  5. Amanatidis, S.; Ntziachristos, L.; Giechaskiel, B.; Bergmann, A.; Samaras, Z. Impact of selective catalytic reduction on exhaust particle formation over excess ammonia events. Environ. Sci. Technol. 2014, 48, 11527–11534. [Google Scholar] [CrossRef]
  6. Gu, Q.; Wang, L.; Wang, Y.; Li, X. Effect of praseodymium substitution on La1-xPrxMnO3 (x=0-0.4) perovskites and catalytic activity for NO oxidation. J. Phys. Chem. Solids 2019, 133, 52–58. [Google Scholar] [CrossRef]
  7. Ning, P.; Song, Z.X.; Li, H.; Zhang, Q.L.; Liu, X.; Zhang, J.H.; Tang, X.S.; Huang, Z.Z. Selective catalytic reduction of NO with NH3 over CeO2-ZrO2-WO3 catalysts prepared by different methods. Appl. Surf. Sci. 2015, 332, 130–137. [Google Scholar] [CrossRef]
  8. Shang, D.H.; Zhong, Q.; Cai, W. High performance of NO oxidation over Ce-Co-Ti catalyst: The interaction between Ce and Co. Appl. Surf. Sci. 2015, 325, 211–216. [Google Scholar] [CrossRef]
  9. Dunn, J.P.; Koppula, P.R.; Stenger, H.G.; Wachs, I.E. Oxidation of sulfur dioxide to sulfur trioxide over supported vanadia catalysts. Appl. Catal. B 1998, 19, 103–117. [Google Scholar] [CrossRef]
  10. Boningari, T.; Ettireddy, P.R.; Somogyvari, A.; Liu, Y.; Vorontsov, A.; McDonald, C.A.; Smirniotis, P.G. Influence of elevated surface texture hydrated titania on Ce-doped Mn/TiO2 catalysts for the low-temperature SCR of NOx under oxygen-rich conditions. J. Catal. 2015, 325, 145–155. [Google Scholar] [CrossRef]
  11. Narasimharao, K.; Malik, M.A.; Mokhtar, M.M.; Basahel, S.N.; Al-Thabaiti, S.A. Iron oxide supported sulfated TiO2 nanotube catalysts for NO reduction with propane. Ceram. Int. 2014, 40, 4039–4053. [Google Scholar] [CrossRef]
  12. Jiang, Y.; Gao, X.; Zhang, Y.X.; Wu, W.H.; Song, H.; Luo, Z.Y.; Cen, K.F. Effects of PbCl2 on selective catalytic reduction of NO with NH3 over vanadia-based catalysts. J. Hazard. Mater. 2014, 274, 270–278. [Google Scholar] [CrossRef] [PubMed]
  13. Kröcher, O.; Elsener, M. Chemical deactivation of V2O5-WO3/TiO2 SCR catalysts by additives and impurities from fuels, lubrication oils, and urea solution: I. catalytic studies. Appl. Catal. B 2008, 77, 215–227. [Google Scholar] [CrossRef]
  14. Li, X.D.; Wang, C.; Huang, J.Q.; Yuan, Y.; Wang, B.; Zhang, H.B.; Xia, F.; Xiao, J.Z. The effects of Cu-content on Mg2CuxFe1O3.5+x electrodes for YSZ-based mixed-potential type NH3 sensors. Ceram. Int. 2016, 42, 9363–9370. [Google Scholar] [CrossRef]
  15. Balle, P.; Geiger, B.; Kureti, S. Selective catalytic reduction of NOx by NH3 on Fe/HBEA zeolite catalysts in oxygen-rich exhaust. Appl. Catal. B 2009, 85, 109–119. [Google Scholar] [CrossRef]
  16. Duan, L.B.; Duan, Y.; Sarbassov, Y.; Li, Y.; Anthony, E. SO3 formation under oxy-CFB combustion conditions. Int. J. Greenh. Gas Con. 2015, 43, 172–178. [Google Scholar] [CrossRef]
  17. Zhao, S.J.; Wang, Y.; Wang, L.; Jin, Y.L. Preparation, characterization and catalytic application of hierarchically porous LaFeO3 from a pomelo peel template Inorg. Chem. Front. 2017, 4, 994–1002. [Google Scholar]
  18. Chen, J.; Duan, L.; Donat, F.; Müller, C.; Anthony, E.; Fan, M. Self-activated nanostructured composite for improved CaL-CLC technology. Chem. Eng. J. 2018, 351, 1038–1046. [Google Scholar] [CrossRef] [Green Version]
  19. Kazantsev, P.A.R.V.; Pankina, G.V.; Maslakov, K.I.; Lunin, B.S.; Eliseev, O.L. Carbon-Silica composite as an effective support for iron fischer-tropsch synthesis catalysts. Energy Technol. 2019, 7, 1800961. [Google Scholar]
  20. Chen, J.; Duan, L.; Sun, Z. Accurate control of cage-Like CaO hollow microspheres for enhanced CO2 capture in calcium looping via a template-assisted synthesis approach. Environ. Sci. Technol. 2019, 53, 2249–2259. [Google Scholar] [CrossRef]
  21. Wu, D.; Shi, Z.F.; Zhang, X.P.; Wu, X.H. Progress of studies on preparation of TiO2 photocatalysts with sol-gel auto igniting synthesis. Earth Environ. Sci. 2017, 94, 1–4. [Google Scholar] [CrossRef] [Green Version]
  22. Parida, K.M.; Reddy, K.H.; Martha, S.; Das, N.; Bisawal, N. Fabrication of nanocrystalline LaFeO3: An efficient sol–gel auto-combustion assisted visible light responsive photocatalyst for water decomposition. Int. J. Hydrogen Energy 2010, 35, 12161–12168. [Google Scholar] [CrossRef]
  23. Sijo, A.K. Magnetic and structural properties of CoCrxFe2−xO4 spinels prepared by solution self-combustion method. Ceram. Int. 2017, 43, 2288–2290. [Google Scholar] [CrossRef]
  24. Cannas, C.; Musinu, A.; Peddis, D.; Piccaluga, G. New synthesis of ferrite-silica nanocomposites by a sol-gel auto-combustion. J. Nanopart. Res. 2004, 6, 223–232. [Google Scholar] [CrossRef]
  25. Kaitlyn, A.; George, K.S. Synthesis of GAGG: Ce3+ powder for ceramics using mechanochemical and solution combustion methods. J. Am. Ceram. Soc. 2018, 101, 3837–3849. [Google Scholar]
  26. Yeh, C.L.; Ke, C.Y. Synthesis of TiB2-Al2O3-FeAl composites via self-sustaining combustion with Fe2O3/TiO2-based thermite mixtures. Ceram. Int. 2018, 44, 16030–16034. [Google Scholar] [CrossRef]
  27. Xiong, Z.B.; Ning, X.; Zhou, F.; Yang, B.; Tu, Y.W.; Jin, J.; Lu, W.; Liu, Z.H. Environment-friendly magnetic Fe-Ce-W catalyst for the selective catalytic reduction of NOx with NH3: Influence of citric acid content on its activity-structure relationship. RSC Adv. 2018, 8, 21915–21925. [Google Scholar] [CrossRef] [Green Version]
  28. Ianoş, R.; Tăculescu, A.; Păcurariu, C.; Lazău, I. Solution combustion synthesis and characterization of magnetite, Fe3O4, nanopowders. J. Am. Ceram. Soc. 2012, 95, 2236–2240. [Google Scholar] [CrossRef]
  29. Ianoş, R.; Moacă, E.A.; Căpraru, A.; Lazău, R.; Păcurariu, C. Maghemite, γ-Fe2O3, nanoparticles preparation via carbon-templated solution combustion synthesis. Ceram. Int. 2018, 44, 14090–14094. [Google Scholar] [CrossRef]
  30. Zhao, J.; Liang, Y.; Yan, H.; Gu, A.; He, S.; Liu, J. Preparation of nano-MnZn ferrite by gel-self-combustion technique. J. Mater. Sci. Eng. 2003, 21, 68–71. [Google Scholar]
  31. Xu, L.T.; Niu, S.L.; Lu, C.M.; Zhang, Q. Influence of calcination temperature on Fe0.8Mg0.2Oz catalyst for selective catalytic reduction of NOx with NH3. Fuel 2018, 219, 248–258. [Google Scholar] [CrossRef]
  32. Zhao, K.; Meng, J.P.; Lu, J.Y.; He, Y.; Huang, H.Z.; Tang, Z.C.; Zhen, X.P. Sol-gel one-pot synthesis of efficient and environmentally friendly iron-based catalysts for NH3-SCR. Appl. Surf. Sci. 2018, 445, 454–461. [Google Scholar] [CrossRef]
  33. Zhang, Q.L.; Liu, X.; Ning, P.; Song, Z.X. Enhanced performance in NOx reduction by NH3 over a mesoporous Ce-Ti-MoOx catalyst stabilized by a carbon template. Catal. Sci. Technol. 2015, 5, 2260–2269. [Google Scholar] [CrossRef]
  34. Liu, F.D.; He, H. Structure-activity relationship of iron titanate catalysts in the selective catalytic reduction of NOx with NH3. J. Phys. Chem. C 2010, 114, 16929–16936. [Google Scholar] [CrossRef]
  35. Wang, R.N.; Wu, X.; Zou, C.L.; Li, X.J.; Du, Y.L. NOx removal by selective catalytic reduction with ammonia over a hydrotalcite-derived NiFe mixed oxide. Catalysts 2018, 8, 384. [Google Scholar] [CrossRef] [Green Version]
  36. Tan, J.B.; Wei, Y.C.; Sun, Y.Q.; Liu, J.; Zhao, Z.; Song, W.Y.; Li, J.M.; Zhang, X. Simultaneous removal of NOx and soot particulates from diesel engine exhaust by 3DOM Fe-Mn oxide catalysts. J. Ind. Eng. Chem. 2018, 63, 84–94. [Google Scholar] [CrossRef]
  37. Wang, H.; Ning, P.; Zhang, Q.; Liu, X.; Zhang, T.; Fan, J.; Wang, J.; Long, K. Promotional mechanism of WO3 over RuO2-Fe2O3 catalyst for NH3-SCO reaction. Appl. Catal. A 2018, 561, 158–167. [Google Scholar] [CrossRef]
  38. Wang, H.; Qu, Z.P.; Xie, H.B.; Maeda, N.; Lei, M.; Wang, Z. Insight into the mesoporous FexCe1−xO2-δ catalysts for selective catalytic reduction of NO with NH3: Regulable structure and activity. J. Catal. 2016, 338, 56–67. [Google Scholar] [CrossRef]
  39. Abolfazl, B.R.; Rezaei, M. Low temperature CO oxidation over Fe-Co mixed oxide nanocatalysts. Chem. Eng. J. 2012, 184, 141–146. [Google Scholar]
  40. Li, X.S.; Li, K.; Yue, Z.P.; Li, X.Y.N.; Zhang, D.; Wang, J.J.; Chen, J.; Li, H. Interaction of phosphorus with a FeTiOx catalyst for selective catalytic reduction of NOx with NH3: Influence on surface acidity and SCR mechanism. Chem. Eng. J. 2018, 347, 173–183. [Google Scholar] [CrossRef]
  41. Xiong, Z.B.; Liu, J.; Zhou, F.; Liu, D.Y.; Lu, W.; Jin, J.; Ding, S.F. Selective catalytic reduction of NOx with NH3 over iron-cerium-tungsten mixed oxide catalyst prepared by different methods. Appl. Surf. Sci. 2017, 406, 218–225. [Google Scholar] [CrossRef] [Green Version]
  42. Xiong, Z.B.; Bai, P.; Zhou, F.; Wu, C.; Lu, W.; Jin, J.; Ding, S.F. Magnetic iron-cerium-tungsten mixed oxide pellets prepared through critic acid sol-gel process assisted by microwave irradiation for selective catalytic reduction of NOx with NH3. Powder Technol. 2017, 319, 19–25. [Google Scholar] [CrossRef]
  43. Li, X.; Li, J.H.; Yue, P.; Xu, B. Selective catalytic reduction of NO with NH3 over novel iron-tungsten mixed oxide catalyst in a broad temperature range. Catal. Sci. Technol. 2015, 5, 4556–4564. [Google Scholar] [CrossRef]
  44. Zhang, G.; Ren, J.J.; Liu, B.L.; Tian, M.; Zhou, H.W.; Zhao, J.S. In situ hydrothermal preparation and photocatalytic desulfurization performance of metallophthalocyanine sensitized SnO2. J. Inorg. Chim. Acta 2018, 471, 782–787. [Google Scholar] [CrossRef]
  45. Liu, Y.; Gu, T.T.; Weng, X.L.; Wang, Y. DRIFT studies on the selectivity promotion mechanism of Ca-modified Ce-Mn/TiO2 catalysts for low-temperature NO reduction with NH3. J. Phys. Chem. A 2012, 116, 16582–16592. [Google Scholar] [CrossRef]
  46. Chen, P.; Tao, L.; Zhu, J.; Zhao, G.; Liu, Y.; Lu, Y. Morphology-controllable hexagonal-phase indium oxide in situ structured onto a thin-felt Al2O3/Al-fiber for the hydrogenation of CO2 to methanol. Energy Technol. 2019, 7, 1800747. [Google Scholar] [CrossRef]
  47. Liu, J.; Xiong, Z.B.; Zhou, F.; Lu, W.; Jin, J.; Ding, S.F. Promotional effect of H2O2 modification on the cerium-tungsten-titanium mixed oxide catalyst for selective catalytic reduction of NO with NH3. J. Phys. Chem. Solids 2018, 121, 360–366. [Google Scholar] [CrossRef]
  48. Castellanos, I.; Marie, O. An operando FT-IR study of the NOx SCR over Co-HFER and Fe-HFER using acetylene as a reducing agent. Catal. Today 2017, 283, 54–65. [Google Scholar] [CrossRef]
  49. Shi, J.; Zhang, Y.; Zhang, Z.H.; Fan, Z.Y.; Chen, M.X.; Zhang, Z.X.; Shangguan, W.F. Water promotion mechanism on the NH3-SCR over Fe-BEA catalyst. Catal. Commun. 2018, 115, 59–63. [Google Scholar] [CrossRef]
  50. Devadas, M.; KroCher, O.; Elsener, M.; Wokaun, A.; Mitrikas, G.; Sӧger, N.; Pfeifer, M.; Demel, Y.; Mussmann, L. Characterization and catalytic investigation of Fe-ZSM5 for urea-SCR. Catal. Today 2007, 119, 137–144. [Google Scholar] [CrossRef]
Figure 1. The NOx conversion over the catalyst. Reaction conditions: [NO] = [NH3] = 1000 ppm, [O2] = 3.0 vol.% and 2000 mL/min of total flow rate. A total of 2 mL of catalyst with a gas hourly space velocity (GHSV) of 60,000/h.
Figure 1. The NOx conversion over the catalyst. Reaction conditions: [NO] = [NH3] = 1000 ppm, [O2] = 3.0 vol.% and 2000 mL/min of total flow rate. A total of 2 mL of catalyst with a gas hourly space velocity (GHSV) of 60,000/h.
Catalysts 10 00314 g001
Figure 2. Thermal analysis curves of the precursor mixtures of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) recorded under an air atmosphere. (A) thermo-gravimetric (TG), (B) differential thermo-gravimetric (DTG), and (C) differential scanning calorimetry (DSC).
Figure 2. Thermal analysis curves of the precursor mixtures of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) recorded under an air atmosphere. (A) thermo-gravimetric (TG), (B) differential thermo-gravimetric (DTG), and (C) differential scanning calorimetry (DSC).
Catalysts 10 00314 g002
Figure 3. The x-ray diffraction (XRD) spectra of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0). (★ gama-Fe2O3 25-1402, ▼ CeO2 43-1002).
Figure 3. The x-ray diffraction (XRD) spectra of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0). (★ gama-Fe2O3 25-1402, ▼ CeO2 43-1002).
Catalysts 10 00314 g003
Figure 4. The scanning electron microscope (SEM) images of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) before and after calcination, respectively.(A) Fe85Ce10W5–CP–CA before calcination, (B) Fe85Ce10W5–CP–CA(NA1.0) before calcination, (C) Fe85Ce10W5–CP–CA after calcination, (D) Fe85Ce10W5–CP–CA(NA1.0) after calcination.
Figure 4. The scanning electron microscope (SEM) images of Fe85Ce10W5–CP–CA and Fe85Ce10W5–CP–CA(NA1.0) before and after calcination, respectively.(A) Fe85Ce10W5–CP–CA before calcination, (B) Fe85Ce10W5–CP–CA(NA1.0) before calcination, (C) Fe85Ce10W5–CP–CA after calcination, (D) Fe85Ce10W5–CP–CA(NA1.0) after calcination.
Catalysts 10 00314 g004
Figure 5. N2 adsorption and desorption isotherms (A), pore diameter distributions (B), and desorption cumulative pore volume (C) of catalysts.
Figure 5. N2 adsorption and desorption isotherms (A), pore diameter distributions (B), and desorption cumulative pore volume (C) of catalysts.
Catalysts 10 00314 g005
Figure 6. H2-temperature program reduction (H2-TPR) profiles of the catalysts.
Figure 6. H2-temperature program reduction (H2-TPR) profiles of the catalysts.
Catalysts 10 00314 g006
Figure 7. NH3–temperature programmed desorption (NH3-TPD) profiles of the catalysts.
Figure 7. NH3–temperature programmed desorption (NH3-TPD) profiles of the catalysts.
Catalysts 10 00314 g007
Figure 8. In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) of Fe85Ce10W5–CP–CA catalyst under transient reactions at 250 °C. (A) Between nitrogen oxides and pre-adsorbed NH3 species. (B) Between NH3 and pre-adsorbed nitrogen oxides species. (C) The NH3, NO, and O2 species.
Figure 8. In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) of Fe85Ce10W5–CP–CA catalyst under transient reactions at 250 °C. (A) Between nitrogen oxides and pre-adsorbed NH3 species. (B) Between NH3 and pre-adsorbed nitrogen oxides species. (C) The NH3, NO, and O2 species.
Catalysts 10 00314 g008
Table 1. The NOx conversion over per gram of catalysts under a gas hourly space velocity (GHSV) of 60,000/h.
Table 1. The NOx conversion over per gram of catalysts under a gas hourly space velocity (GHSV) of 60,000/h.
SamplesNOx Conversion (mg/(g·h))
125 °C150 °C175 °C200 °C
Fe85Ce10W5–CP–CA8.521.051.687.3
Fe85Ce10W5–CP–CA(NA1.0)7.415.427.250.4
Table 2. The surface area, pore volume, diameter, and average crystallite size of the catalysts.
Table 2. The surface area, pore volume, diameter, and average crystallite size of the catalysts.
SamplesSBET a (m2/g)Pore Volume b (cm3/g)Pore Diameter c (nm)D d (nm)
Fe85Ce10W5–CP–CA90.8500.1786.17013.000
Fe85Ce10W5–CP–CA(NA1.0)93.1300.1464.92014.400
a Brunauer-Emmett and Teller (BET) surface area; b Barrett-Joyner and Halenda (BJH) desorption pore volume; c Barrett-Joyner and Halenda (BJH) desorption pore diameter; d Calculated according to the Scherrer equation.
Table 3. The X-ray photoelectron spectroscopy (XPS) results of catalysts.
Table 3. The X-ray photoelectron spectroscopy (XPS) results of catalysts.
SamplesSurface Atomic Concentrations (%)
Fe2+Fe3+FetotalCeWOαOβOtotal
Fe85Ce10W5-CP-CA5.6511.7117.367.242.8915.8256.6972.51
Fe85Ce10W5-CP-CA(NA1.0)5.7311.8317.566.332.3616.0957.6673.75

Share and Cite

MDPI and ACS Style

Ning, X.; Xiong, Z.-b.; Yang, B.; Lu, W.; Wu, S.-m. The Role of Nitrate on the Sol-Gel Spread Self-Combustion Process and Its Effect on the NH3-SCR Activity of Magnetic Iron-Based Catalyst. Catalysts 2020, 10, 314. https://doi.org/10.3390/catal10030314

AMA Style

Ning X, Xiong Z-b, Yang B, Lu W, Wu S-m. The Role of Nitrate on the Sol-Gel Spread Self-Combustion Process and Its Effect on the NH3-SCR Activity of Magnetic Iron-Based Catalyst. Catalysts. 2020; 10(3):314. https://doi.org/10.3390/catal10030314

Chicago/Turabian Style

Ning, Xing, Zhi-bo Xiong, Bin Yang, Wei Lu, and Shui-mu Wu. 2020. "The Role of Nitrate on the Sol-Gel Spread Self-Combustion Process and Its Effect on the NH3-SCR Activity of Magnetic Iron-Based Catalyst" Catalysts 10, no. 3: 314. https://doi.org/10.3390/catal10030314

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop