Next Article in Journal
Recent Advances in Industrial Sulfur Tolerant Water Gas Shift Catalysts for Syngas Hydrogen Enrichment: Application of Lean (Low) Steam/Gas Ratio
Next Article in Special Issue
In Situ IR Study on Effect of Alkyl Chain Length between Amines on Its Stability against Acidic Gases
Previous Article in Journal
Efficient Production of Medium-Chain Structured Phospholipids over Mesoporous Organosulfonic Acid-Functionalized SBA-15 Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Critical Review of Recent Progress and Perspective in Practical Denitration Application

1
Key Laboratory for Green Processing of Chemical Engineering of Xinjiang Bingtuan, School of Chemistry and Chemical Engineering, Shihezi University, Shihezi 832003, China
2
Bingtuan Industrial Technology Research Institute, Shihezi University, Shihezi 832003, China
3
Tianjin Key Laboratory for Heavy-Metal Containing Wastewater Treatment, Tianjin 300350, China
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(9), 771; https://doi.org/10.3390/catal9090771
Submission received: 19 August 2019 / Revised: 4 September 2019 / Accepted: 9 September 2019 / Published: 13 September 2019

Abstract

:
Nitrogen oxides (NOx) represent one of the main sources of haze and pollution of the atmosphere as well as the causes of photochemical smog and acid rain. Furthermore, it poses a serious threat to human health. With the increasing emission of NOx, it is urgent to control NOx. According to the different mechanisms of NOx removal methods, this paper elaborated on the adsorption method represented by activated carbon adsorption, analyzed the oxidation method represented by Fenton oxidation, discussed the reduction method represented by selective catalytic reduction, and summarized the plasma method represented by plasma-modified catalyst to remove NOx. At the same time, the current research status and existing problems of different NOx removal technologies were revealed and the future development prospects were forecasted.

Graphical Abstract

1. Introduction

NOx is one of the main sources of haze generation and air pollution in all worldwide. On one hand, NOx comes from the natural factors of nitrate caused by decay of organism after biological death and plant straw-burning as well as in the lightning process; on the other hand, it stems from the artificial factors, for instance, iron and steel smelting; petroleum cracking and coal-based fossil fuel power generation; fuel vehicle exhaust emissions [1,2]; and NOx decomposition under the action of soil microorganisms, which is caused by the large amount of nitrogen fertilizer applied in farmland [3,4]. The emission of toxic waste gases has resulted in a significant increase in the content of NOx in the atmosphere, which is damaging to human organs, such as the heart and lungs, and reducing the body’s immunity [5], combining with volatile organic compounds to produce ozone [6], making the human body susceptible to respiratory diseases such as coughs, sore throat, and bronchitis [7,8]; it can also lead to premature birth [9], severely inhibiting photosynthesis, and thus affects plant growth and oxygen conversion [10,11]; but, at the same time, produces photochemical smog [12,13] and acid rain [14]. Furthermore, it forms aerosol with SO2, dust, and water vapor, then produces PM2.5 and other small particle size air pollution suspensions, namely, haze, which cause serious air pollution worldwide [15,16].
Therefore, the treatment of NOx is imminent. The supervision of flue gas emission from fixed source factories and the treatment of motor vehicle exhaust emission from mobile sources have been widely controlled [17] recently, appearing to be the following mature technologies; removal of NOx by adsorption, oxidation, reduction, and plasma methods. In view of these attentive findings, adsorption method uses recyclable solid adsorbent material to adsorb NOx from flue gas through microporous structure, remarkably, environmentally friendly, no secondary pollution, and energy-saving and -efficient features make it stand out [18,19,20]. The oxidation process is a method that oxidizes NO into NO2 or N2O3 with high solubility under the action of an oxidant that is then absorbed by water or alkali solution. The process is simple and cost-saving, and is widely used in the wet flue gas denitrification process with relatively low cost and high removal efficiency of NOx; except that the oxidized NO and SO2 can be removed simultaneously to produce high value-added products [21,22,23]. The reduction method has the characteristics of high efficiency, cleanliness, and mildness; NOx removal can be realized at low temperature at present [24,25,26]. The plasma method is used to modify the surface of the catalyst, the dispersion of active species can be improved by plasma treatment, and the stability and low temperature activity of catalyst can be improved [27,28].
Seldom, if ever, does a comprehensive article to introduce the current situation and treatment methods of removing NOx. In this paper, we tried to renovate, classify, and summarize the significant perspectives and most relevant findings reported in recent research articles and earlier reviews generalizing the mechanism analysis and research progress of NOx removal by adsorption, oxidation, reduction, and plasma methods, which can provide means for promoting the technological progress of pollution prevention, maintaining healthy development of factories continuously, studying the methods of removing NOx deeply, and mastering different technologies of removing NOx, as well as providing guidance for controlling the emission of NOx from motor vehicles such as diesel, gasoline, and so on. In order to improve the environmental quality and save the ecological environment, this paper further provides a convenient, low-cost, efficient, and economic guidance line.

2. Removal of NOx by Adsorption

The adsorption method for NOx treatment is simple to operate, controllable in process, and can capture toxic waste gases with very low concentration efficiently. NOx in flue gas is adsorbed by recoverable solid adsorbents through microporous structure [29,30,31]. The adsorption materials for NOx in waste gas mainly include molecular sieve, activated carbon, zeolite, heterophony acid, silica gel, and peat [32]. The adsorption method is efficient and thorough, and can produce nitric acid with high added value. Meanwhile, in recent years, adsorption denitrification technology mainly includes the organic framework of copper (Cu-BTC) adsorption and carbon material adsorption methods [33].

2.1. NOx Removed by Cu-BTC Adsorption

Metal–organic frameworks (MOFs) have many advantages, e.g., porous, high specific surface area, and abundant Lewis acidic active sites, and make excellent gas adsorption materials [34], among them, Cu-based MOFs have excellent catalytic performance [35]. Cu-BTC is a special organic metal framework in MOFs. Qin et al. [36] prepared Agx-Cu-BTC dispersed in bimetallic organic framework by preassembly method to absorb NOx, this framework had regular octahedral shapes with particle sizes ranging from 10 to 30 microns. Ag-doping into the metal–organic framework of Cu-BTC resulted in a heterojunction, which made the organic functional group, –COO–, blue shift and accompanied with good dispersion of Ag and Cu atoms in the organic framework.
In order to further study the adsorption of NOx by Cu-BTC material, Khan et al. [37] studied the adsorption of NO in MOFs-Cu3(BTC)2 and Cu3(NH2BTC)2 by 1H solid-state NMR, and found that 1H in MOFs-Cu3(BTC)2 had a linear relationship with NO adsorption. Kaur et al. [38] explained the synthesis and adsorption properties of Cu-BTC porous octahedral nanocrystals. Meng et al. [39] found that in the framework of Cu-BTC, CO2 varies with the Langmuir distribution, while the N atom in the NO molecule is mainly responsible for the coordination with the Cu atom. Qin et al. [40] studied the adsorption mechanism of NO on A-Cu-BTC surface (Figure 1a). It was found that the doping of Sr, Ce, Al, and other elements in Cu-BTC could increase the NO conversion (Figure 1b). The modification of Ce by X-ray diffraction analysis enhanced the dispersion of Cu in metal–organic framework BTC (Figure 1c). Scanning electron microscopy further confirmed that the formation of Ce–Cu heterojunction enhanced the low degree (Figure 1d).

2.2. NOx Removed by Carbonaceous Material Adsorption

Removal of NOx by adsorption, especially activated carbon, has undergone a long historical evolution [41]; activated carbon originated from 1900, and its inventor was Raphael von Ostrejko, by whom, British and German patents had been obtained. In 1920, the demand for activated carbon continued to expand through the extensive application of gas masks; at the same time, the application of activated carbon in sewage treatment and catalysis was gradually developed [42,43], activated carbon plants in major industrial countries, including the United States, had been gradually opened and expanded. In the mid-20th century, the field of activated carbon continued to aggrandize and was regarded as “universal adsorbent”. Today, activated carbon is used in pollution control [44,45] with a high-precision instrument to produce pore structure matching the molecular size of the toxic waste gas, which is used to adsorb harmful gases such as benzene, formaldehyde, NOx, NH3, and so on [46,47].
Activated carbon, which is not a natural material, but an artificial material made by processing carbon-containing organic matter, has been widely used as catalyst [48]. Activated carbon comes from a wide range of raw materials, which can be almost any organic compound material, including sawdust, bagasse, branches, shells and other wood materials [49], coking coal, bituminous coal, and other bulk commodities [50]. In the process of activation smelting, a huge specific surface area and complex pore structure are gradually formed, which can effectively remove chroma and odor, and remove most organic pollutants and some inorganic substances in secondary effluent [51], containing some toxic heavy metals [52,53]. Furthermore, activated carbon has excellent thermal and chemical stability, high wear resistance and a variety of oxygen-containing functional groups on the surface, so activated carbon adsorption is a very promising means in the process of flue gas treatment.
Different kinds of oxygen-containing groups are the main active sites on activated carbon [54]; however, the untreated activated carbon has smaller pore volume and specific surface area, higher ash content, and poor adsorption performance. The chemical properties of activated carbon surface can be changed by chemical oxidation, reduction, and loading [55,56]. Jie et al. [57] found that proper oxidation could improve the degradation ability of activated carbon to nitrogen compounds. Wang et al. [21] found that pretreatment of carrier activated carbon with HNO3 could improve the NO conversion. Li et al. [58] discovered that the pore size of activated carbon fibers decreased and the surface functional groups increased after modification with Cu2O/TiO2, which improved the adsorption capacity of activated carbon fibers for NO. Javier et al. [59] prepared goethite and hematite twin grains on activated carbon, which improved the total denitrification capacity. You et al. [60] caused defects on the surface of activated carbon by changing pore size and nitrogen doping, which greatly improved its NO catalytic oxidation activity at room temperature. Sun et al. [61] used Cu-Fe modified microwave coconut shell activated carbon (CuFe/MCSAC) to remove NO from industrial tail gas, and studied the adsorption process of NO on the surface of activated carbon (Figure 2a). It was found that 3% O2 was beneficial to increase the conversion of NO, which was attributed to NO oxidation as the first step of selective catalytic reduction reaction (Figure 2b). In addition, the sulfur resistance was studied (Figure 2c,d). The preparation methods and other conditions of different adsorption materials are shown in Table 1.

2.3. Other Crucial NOx Adsorbents

There are numerous kinds of NOx adsorbents; low-priced and efficient adsorbents are required for denitrification. Li et al. [63] prepared fly ash derivative Cu/SAPO-34 by acid–alkali combined hydrothermal method to absorb NOx. Mire et al. [64] studied the removal of NOx using H-ZSM-5 as adsorbents, and found that Fe-modified Fe/H-ZSM-5 had the strongest adsorption capacity. Chen et al. [65] synthesized Fe-ZSM-5@CeO2 adsorbent for adsorbing NOx. Baran et al. [66] prepared Cu-BEA molecular sieve to study the adsorption characteristics of NOx. Ma et al. [67] developed a new Ag nano-g-C3N4/WS2 material to adsorb NOx. Imai et al. [68] developed a niobium phosphate adsorbent to remove NOx. Xiao et al. [69] studied the adsorption of NOx on mesoporous molecular sieve doped with platinum (Pt/SBA-15). Wang et al. [70] found that when iron and cobalt were loaded on the active semi-coke material; it was beneficial to NOx adsorption. Chen et al. [71] studied NO removal performance based on sinter adsorption materials. The synthetic methods and reaction conditions of different adsorption materials are shown in Table 2.
Chemical adsorption usually requires a certain activation energy, and the adsorption rate is slow at low temperature [72]. The adsorption method is closely related to the surface chemical properties of adsorbents, and it can not only eliminate the pollution of NOx to the atmosphere, but recycle them to produce high value-added products. However, due to its limited adsorption capacity, the size of the absorber is huge, the adsorbent needs to be regenerated frequently, and the one-time investment is large. In particular, activated carbon has the possibility of spontaneous combustion above 300 °C, which causes considerable difficulties in the regeneration of adsorbents and limits its application.

3. Removal of NOx by Oxidation

Since MeKee oxidized low-concentration NO to NO2 in 1921; it laid the foundation for NO oxidation [73]. In the early 1980s, Japanese scholars made a more in-depth study on the oxidative denitrification. At this time, NO oxidative denitrification technology rose gradually. In the late 1990s, Takashi Ibusuki et al. [74] found that TiO2 can oxidize NO to HNO3 under ultraviolet irradiation rapidly. Nowadays, the technology of removing NOx by oxidation is relatively mature. It can not only control the morphology of catalysts by some means, but study the law that affects the activity of catalysts. At the same time, investigators often only study the activity of catalysts in a low-temperature window gradually [75,76,77]. This method uses oxidizer to oxidize NO, which is not soluble in water, to NO2 or N2O3 in flue gas [78,79,80], which is then sprayed and washed by water, alkali solution, acid solution, or metal complex solution so that the NOx in the gas phase are transferred to the liquid phase to realize the denitrification treatment of the flue gas. O3, ClO2, NaClO2, NaClO, KMnO4, H2O2, Cl2, and HNO3 can be used as oxidants [81]. According to the types of oxidation technology, it can be divided into gas-phase oxidation, liquid-phase oxidation, and catalytic oxidation.

3.1. NOx Removed by Gas-Phase oxidation

Since the development of a low temperature oxidation technology called LoTOx by Linde Industrial Gases Company in the United States, more and more researchers have been studying the denitrification performance of O3. The gas-phase oxidation method has been used to purify flue gas from boilers fueled with natural gas, and the denitrification rate is over 90%. The purification process is simple and has no effect on the normal operation of the boiler. It can recover high-grade HNO3 [82,83]. O3 is highly oxidizable, and the reduction potential in the process of O3 ↔ O2 can reach 2.07 eV, which makes the oxidation of NO possible.
At present, ClO2 and O3 are widely used in oxidative denitrification as gas phase oxidants. Hultén et al. [84] found that when the molar ratio of ClO2 to NO was 0.6 at 160 °C, the total amount of NOx removed was 94%. Sun et al. [85] discovered that the removal efficiency of NOx increased with increasing O3 concentration. The removal efficiency of NOx reached 82% when the flue gas temperature exceeded 40 °C. Mok et al. [86] proposed a two-step process consisting of an O3 chamber and an absorber containing a reductant solution for simultaneous removal of NOx. When the flue gas passes through the O3 chamber, and in turn the absorber, the removal rate of NOx is ~95%. Lin et al. [87] studied the deep oxidation of NO by O3 on the MnOx-based catalyst. The experimental device of deep oxidation showed that the catalyst had good stability and sulfur resistance. Han et al. [88] prepared the black TiO2 catalyst by sol–gel method. The deep oxidation of NO to HNO3 was further strengthened (Figure 3a): black TiO2 had higher NO conversion than TiO2 at the low temperature of 60 °C (Figure 3b). New peaks of black TiO2 at 532.6 eV were found by X-ray photoelectron spectroscopy (XPS) chromatography, which further confirmed the excellent catalytic activity of black TiO2 (Figure 3c). In addition, the sulfur resistance of black TiO2 was investigated (Figure 3d).

3.2. NOx Removed by Liquid-Phase Oxidation

The liquid-phase oxidation absorption method is that NO is oxidized by liquid phase oxidant and then absorbed by alkali absorption method. The de-NOx rate can reach 90–95% when using KMnO4 or NaClO2 as liquid phase oxidizers [89]. NaClO2 has strong oxidizing property; its solution can oxidize NO to NO2, and then NaNO3 was obtained. It is easy to oxidize NO with NaClO2 and the removal rate of NO is high [90]. In order to cope with the increasingly severe environmental problems and find effective ways to control air pollution, the NO gas discharged from flue gas can be removed simultaneously through a wet NaClO2 scrubber combined with a plasma electrostatic precipitator.
Deshwal et al. [91] used NaClO2 solution to remove NOx from simulated flue gas. NaClO2 was a better oxidant when the pH was less than 4. NaClO2 was decomposed into ClO2 gas in acidic medium. It was considered that NaClO2 participates not only in NO oxidation but in NO2 absorption. Hao et al. [92] prepared a new composite oxidant NaClO2/Na2S2O8, removing SO2 and NO simultaneously by pre-oxidation. The removal rates of SO2 and NO were 100% and 82.7%, respectively, under the optimum conditions. Yang et al. [93] used ultraviolet radiation sodium chlorite (UV/NaClO2) solution to remove NOx from simulated flue gas in a small scrubbing reactor. The results showed that the addition of Cu and Mn resulted in the lattice expansion and micro-strain decrease of CeO2-ZrO2, which led to the formation of oxygen vacancies. Besides, there was a strong interaction between surface Cu, Mn, and Ce through charge transfer.

3.3. NOx Removed by Catalytic Oxidation

O3 has a high preparation cost in gas-phase oxidation, which leads to its narrow promotion area; the liquid-phase oxidation method requires higher corrosion resistance of equipment, and the price of oxidant is relatively expensive, which restricts its industrial application. Therefore, the preponderance of catalytic oxidation method are beginning to highlight [94]. The oxidant is the excess oxygen in flue gas, the catalysts are vanadium, tungsten, and titanium and rare earth metal oxides supported on activated carbon, alumina, and silica. This method has been widely used in flue gas purification with the denitrification rate is over 90% [95]. The preponderance of catalytic oxidation are closed-circuit circulation; simultaneous removal of SO2, dust, and other pollutants without additional oxidizer; injection of 5% steam into flue gas can improve the efficiency and life of catalyst; and low investment and operation cost of equipment. Catalytic oxidation technology for NOx removal includes molecular sieve catalytic oxidation, metal oxide catalytic oxidation and Fenton system catalytic oxidation, which have their respective applicable conditions.

3.3.1. Catalytic Oxidation of Molecular Sieve

A kind of natural aluminosilicate, which has the functions of ion exchange, molecular sieving, catalysis and adsorption, called zeolite, and synthetic zeolite is also called molecular sieve. The basic structure of molecular sieve skeleton is Al2O3 and SiO4 tetrahedrons, which form a three-dimensional network structure by combining the common O atoms [96]. It has strong selectivity, high-temperature resistance, and high adsorption capacity, which is widely used in organic chemical industry and petrochemical industry. It is also an excellent adsorbent for gas dehydration. At the same time, more and more attention has been paid to the purification of waste gas. Molecular sieves have regular and uniform intracrystalline channels, and the pore size is close to the molecular size. The catalytic performance of molecular sieves varies significantly with the geometrical size of reactant molecules, product molecules or reaction intermediates [97].
Molecular sieve catalysts for NO treatment are mostly used in oxidation and decomposition processes, such as octahedral molecular sieve (OMS-2) catalysts. Molecular sieve catalysts for oxidation reaction are mainly transition metal ion exchange zeolites and specific types of zeolite zeolites, and these zeolite catalysts only show high temperature activity. Liu et al. [98] first used solvent-free synthesis of cryptomelane OMS-2 to oxidize waste material in coal-fired flue gas, which provided a new idea for oxidative denitrification by molecular sieve. Geng et al. [99] prepared zinc oxide nanoparticle molecular sieves by melt infiltration method to remove exhaust gas at room temperature. Saminda et al. [100] synthesized Mn octahedral molecular sieve (OMS-2) to remove flue gas by catalytic oxidation. Park et al. [101] studied the equilibrium and kinetics of adsorption of N2O on carbon molecular sieves, concluding that the adsorption rate of N2O was influenced by kinetics and Lewis structure. Sarmah et al. [102] analyzed the microstructure and morphology of OMS by scanning electron microscopy, which showed the highly cross-linked nanowire morphology. Hamaguchi et al. [103] studied the mechanism of NO removal on the surface of molecular sieve K1.14Mn8O16 (Figure 4a), finding that the thermal unstable oxygen potential acted as the oxidation site of NO at low temperature. The doping of K increased the adsorption of NO (Figure 4b). The particle size of K was 0.46 nanometers (Figure 4c). NO generated in K-OMS-2 molecular sieves NO3, the formation of NO3 is proved by Kubelka–Mung theory (Figure 4d).

3.3.2. Catalytic Oxidation of Metal Oxides

Metal oxide, the compound of O and another metal element, is a kind of excellent catalyst. As the main catalyst, promoter and carrier, it has been widely used in the field of catalysis. After nanotreatment of metal oxides, nano-materials with uniform dispersion, small particle size, and high purity can be obtained, and their catalytic performance is better [104]. Metal oxides provide oxygen free radical formation products for reactants and provide active sites for adsorption (or coordination) of reactants to make them active as well.
Metal oxide is a kind of catalyst that has been studied earlier in the field of NO catalytic oxidation and developed vigorously. Jogi et al. [105] found that the presence of TiO2 greatly enhanced the efficiency of O3 oxidation of NO2 to N2O5 at 100 °C. Wang et al. [106] synthesized manganese oxides with different valences (MnO2, Mn2O3, and Mn3O4) by hydrothermal method. At the space velocity of 48,000 mLg−1h−1, with oxygen as oxidant, the maximum NO conversion rate was 91.4%. Ma et al. [107] synthesized nanoparticle-like monocrystalline SmMn2O5 by adjustable hydrothermal method; among them, nanoparticle-like SmMn2O5 achieved more than 90% NO conversion at 300 °C. Jia et al. [79] synthesized a series of Mn–Fe catalysts for oxidative denitrification by hydrothermal method. It was found that the addition of iron oxides could effectively improve the catalytic oxidation performance of manganese dioxide at low temperature. Cai et al. [108] synthesized Cr–Ce oxidative denitrification catalyst with three-dimensional structure by hydrothermal method; more active sites could be exposed with the increase of surface particle size and roughness (Figure 5a). The chromium-Ce/Et catalyst synthesized with ethanol as solvent had the highest catalytic oxidative denitrification activity (Figure 5b). Through a series of characterization catalysts, the Cr–Ce/Et coexistence effect formed the phase interface between atoms (Figure 5c). In addition, the effects of steam and SO2 on NO conversion over Cr–Ce/Et catalysts at 250 and 300 °C were studied (Figure 5d).

3.3.3. Photocatalytic Oxidation

Photocatalytic oxidation is a new process of NOx treatment, which has been developed gradually over the last ten years. The principle of photocatalytic oxidation is to irradiate semiconductor catalysts with light of specific wavelength, to stimulate valence band electrons on semiconductor materials to transit into conduction band, and to generate holes in valence band. Conduction band electrons and valence band holes have strong reductive and oxidative properties, respectively. When they are in contact with flue gas, H2O, O2, and NOx adsorbed on the catalyst surface will generate active free radicals under the action of light, and then catalytic oxidation reaction will take place to oxidize NOx to NO3. Because it does not need to inject additional reductant and is simple to operate with low-cost and no secondary pollution, it is the focus of photocatalytic technology research at present [109,110].
The catalysts for photocatalytic oxidation are mainly metal oxide semiconductor materials. Among them, TiO2 has the merits of high catalytic activity, photochemical stability and low price. It is the most commonly used catalytic material in photocatalytic reaction. Duan et al. [111] found that a small amount of Ag nanoclusters supported on TiO2 was beneficial to improve photocatalytic activity. He et al. [112] synthesized carbon-encapsulated nanocrystalline TiO2 catalyst, which achieved 71% NO conversion under visible light. Yuan et al. [113] prepared TiO2 aluminosilicate fiber nanocatalyst for photocatalytic removal of NOx. Jin et al. [114] prepared SrTiO3 catalyst with SrCO3 successfully as auxiliary by one-step in situ pyrolysis for removal of NOx. Huy et al. [115] reported for the first time the removal of NOx by visible light on SnO2/TiO2 nanotube heterojunction catalysts. The reaction mechanism was shown in Figure 6a. The removal efficiency of NOx under visible light was represented in Figure 6b. Figure 6c shows that SnO2/TiO2 nanotube heterojunctions had been successfully synthesized. Figure 6d proves that the superoxide anion radical played an important role in visible light photocatalytic oxidation by using active oxygen detection materials.

3.3.4. Catalytic Oxidation of Fenton System

The essence of Fenton system is that the chain reaction between Fe2+ and H2O2 catalyzes the production of OH· with strong catalytic oxidation ability. Its oxidation potential is only 2.80 V after that of fluorine, which is a strong oxidation system. In addition, OH· has high electrophilicity and electronegativity and its electron affinity can reach 569.3 kJ, which is associated with strong additive reaction properties. Therefore, Fenton reagent can selectively degrade most organic compounds in water, especially suitable for the treatment of organic wastewater which is difficult to be degraded by biology or chemical oxidation. The prospect of catalytic oxidation of NOx in flue gas is also extremely promising [116,117].
NO can be oxidized efficiently at low temperature with H2O2. At present, Fenton reagents and Fenton-like reagents are widely studied [118,119]. H2O2, no secondary pollution and low price, is one of the main components of Fenton reagent and can achieve high denitrification rate [120]. Fenton reagent refers to the H2O2/Fe2+ system, through a series of catalytic actions of Fe2+ to produce iso-strong oxides OH·, HO2·, and O2. Wang et al. [121] adopted a new Fenton process catalyzed by Cu2+ and Fe2+ in coordination, which inducing more OH· to deoxidize NO, and enhancing the denitrification activity of Fenton catalyst. Zhao et al. [122] carried out simultaneous denitrification of flue gas in bubbling reactor with Fenton reagent. The removal rate of NO was over 90%. Cheng et al. [123] found that too high Fe2+ concentration and pH were not conducive to NO oxidation and when the pH was only 2–5, the formation rate of OH reached the maximum. Guo et al. [124] confirmed that the oxidation of NO by Fenton reagent occurred on the liquid membrane side, low concentration of NO was not conducive to mass transfer to the liquid side. Cui et al. [125] took fly ash as raw material and removed NOx with hydrogen peroxide. It is found that Fenton heterogeneous catalyst modified by alkali, acid, and alkali has high catalytic activity. Zhao et al. [126] used heterogeneous ultraviolet spectrophotometry to remove SO2 and NO from flue gas for the first time. Yuan et al. [127] oxidized NO with H2O2 as oxidant in the two-ion Fenton reaction system (Figure 7a,b). The modification of Ce could significantly improve the conversion of NO (Figure 7c). The effect of Ce on the dispersion of Fe2+ was further investigated by XRD (Figure 7d).
In recent years, oxidation method has been widely used in the United States and European countries to remove NOx. Selective catalytic oxidation method has the advantages of relatively low cost and high removal efficiency of NOx. The oxidized NO can be removed simultaneously with SO2 to produce high value-added products. The process is simple and cost-saving [128,129]. The reagents materials and other conditions of different oxidation methods are shown in Table 3.

4. Removal of NOx by Reduction

As early as the late 1970s, reduction denitrification technology was put into use in Japan. Over time, reduction technology had been widely used in Europe and the United States, and is considered to be the most advantageous technology for removing NOx. Because of the limitations of NOx control before and during combustion, the control of NOx in flue gas after combustion has attracted wide attention [131]. There are many technologies for removing NOx from flue gas, according to different principles, they can be divided into catalytic reduction and adsorption, besides, in the view of different working media, and they can be divided into dry and wet methods. Different treatment methods are selected according to the concentration of NOx tail gas and other working conditions [132,133,134]. Reduction denitrification technology includes selective noncatalytic reduction and selective catalytic reduction.

4.1. Selective Noncatalytic Reduction (SNCR) for NOx Removal

Selective noncatalytic reduction (SNCR) has been developed earlier and its technology is relatively mature, which can generally reduce NOx by 50 to 60%. The main principle is that under the condition of catalyst, ammonia, or urea with reductive groups are injected into the flue gas of high temperature furnace was between 850 and 1100 °C. NOx in the flue gas is reduced to harmless nitrogen after atomizing and spraying into the reactor. The advantage of SNCR technology is that there is less investment and occupied area a relatively small, SNCR technology has certain limitations and is generally only suitable for small coal-fired boilers with low NOx content [135]. Since there is no catalyst to accelerate the reaction, the operating temperature is higher than that of SCR method. In order to avoid NH3 being oxidized, the temperature should not be too high. The current trend is to use urea instead of NH3 as reducing agent.
In order to further study the development status of selective noncatalytic reduction (SNCR), Hao et al. [136] found that in the process of SNCR containing NH3, the use of water vapor could increase NO reduction rate, and the optimum water vapor content was 4–8%. Sodium and potassium additives could promote NO reduction in the order of Na2CO3 > KCl > NaCl. Fu et al. [137] found that CaO had an effect on the pyrolysis of urea during melt coalescence, thus reduced the pyrolysis rate of urea and the performance of urea SNCR denitrification process. Chen et al. [138] found that when urea was added into hydrazine hydrate solution, better denitrification efficiency could be obtained in the temperature range of 550 to 650 °C. Kang et al. [139] validated the optimized SNCR reaction mechanism, and combined the optimized mechanism with CFD software to simulate the SNCR denitrification process. Fu et al. [140,141] studied the reaction mechanism of SNCR denitrification on CaO and CaCO3 surfaces (Figure 8a). NH3 conversion and NO selectivity in different mixed gas without oxygen were analyzed (Figure 8b), HCHO was the main product of NH3 transformation on CaCO3 surface (Figure 8c), and in situ infrared spectroscopy further confirmed the species of reaction intermediates (Figure 8d).

4.2. Selective Catalytic Reduction (SCR) for NOx Removal

SCR is one of the most widely used denitrification methods at home and abroad at present. It was first developed by the United States, and then widely used in industrial production in Japan. SCR can make the removal efficiency of NOx reach more than 90% [142,143]. The basic principle of SCR technology is that under the existence of catalyst, NH3, CO, and other reducing agents are used to reduce the NOx to N2 in flue gas. At the same time, this method also has some shortcomings. The core of SCR technology lies in the selection of excellent catalysts. The complex composition of flue gas can easily deactivate and poison the catalysts. Especially for water and sulfur, SCR catalysts are more sensitive, which affects the life of the catalysts, and makes the cost of treatment and operation significantly higher [144,145,146]. SCR technology is a method of selectively reducing NOx to nitrogen and water by using ammonia or hydrocarbon or carbon monoxide as reductant under certain temperature and catalyst [147,148]. SCR is the most mature and widely used denitrification technology with high denitrification efficiency. It is the most efficient means to control NOx pollution. Generally, the denitrification efficiency can reach 80–90% based on the rational selection of reactor and catalyst [149].

4.2.1. NH3-SCR

NH3-SCR is one of most efficient processes for denitrification. Up to date, NH3-SCR systems have been widely applied in stationary industrial installations (SIIs), such as power plants and industrial boilers. Commercial vanadium-based catalysts usually work in the temperature range of 300 to 400 °C, and often suffer from high dust and SO2 contents in flue gases, while a manganese-based monolithic honeycomb catalyst (MHC) showed high activity in NH3-SCR at room temperature (as low as 20 °C). We believe that this room temperature NH3-SCR technology is effective for high concentration NOx removal with great potential for applications in SIIs.
Current SCR denitration (De-NOx) technologies come in three categories, high-temperature De-NOx at high dust content, high-temperature De-NOx at high SO2 content, and low-temperature De-NOx at high H2O content, as shown in Figure 9. These applications are particularly challenging because of the varying operation conditions of SIIs with regard to flue gas temperature. The general niches and limitations of monolithic honeycomb catalysts (MHCs) for SCR processes and practical NOx removal can be highlighted as follows.
(i)
High-temperature De-NOx unit at high dust content. In this unit, the temperature of the flue gas entering the NH3-SCR reactor is as high as 300–400 °C. This high temperature significantly favors the De-NOx performance of most of the catalysts (e.g., vanadium-type-based catalysts) [150,151]. At present, this technology is the most widely used in SIIs. However, further progress in promising applications is still somewhat bottlenecked by (a) the high dust content in the flue gas, which easily leads to erosion and plugging holes in MHCs and (b) the high SO2 content of the flue gas, leading to catalyst poisoning and deactivation. These limitations are yet to be overcome [152].
(ii)
High-temperature De-NOx unit at high SO2 content. In this unit, the temperature of the flue gas entering the NH3-SCR reactor ranges from 180 to 280 °C. A dust-removing apparatus is used to mitigate the erosion and plugging issues for longer service life of MHCs for NH3-SCR. Nevertheless, the high SO2 content continues to raise greater worries for activity declining and shorter service life [153,154]. Most of manganese- and copper-based catalysts with high SO2 resistance could be used in this process [155,156].
(iii)
Low-temperature De-NOx unit at high H2O content. In this unit, the temperature of the flue gas entering the NH3-SCR reactor is below 160 °C. Although dust and SO2 are previously removed, the H2O content in the flue gas still remains as high as 10 vol.%. Almost all current catalysts show low activity and poor water-resistance at this temperature [157,158]. Therefore, this technique is not practical for SIIs.
Additionally, in the process of low-temperature De-NOx unit after H2O removal, the water vapor in the flue gas is condensed and recovered. Furthermore, the flue gas obtained after the dust removal, desulfurization, and water recovery steps can be heated to a temperature ranging from 50 to 150 °C, depending on the flue gas flow. Most Mn-based catalysts exhibit great NH3-SCR performance at low temperature [160]. This technique has potential in application for SIIs; however, the NO removal at low temperature still exists as a big challenge.
Zhang et al. [161] found inspiring room temperature De-NOx performance in NH3-SCR with Mn-based MHCs. The MnOx-Fe2O3/vermiculite (VMT) catalyst showed high N2 selectivity of 97.1% at 20 °C. The corresponding MnOx-Fe2O3/VMT MHC exhibited a NO conversion of 62.2% at NO content of 500 ppm. Additionally, MnOx-Fe2O3/VMT MHC showed excellent SO2 resistance even at SO2 content of 300 ppm, at which sulfate might be easily formed to foil the catalyst surface. As shown in Figure 8b, this room temperature NH3-SCR technology is effective for high concentration NOx removal, showing great potential for applications in SIIs.
For low temperature NH3-SCR denitrification catalyst, it is very crucial to study denitrification mechanism for further improving its activity [162,163]. Ryu et al. [164] studied the location of active sites in NH3-SCR reaction, mainly at Lewis acid sites on the catalyst surface. Campisi et al. [165] functionalized HAP by ion exchange. The high dispersion of Fe3+ species ensured the acid and redox centers. Yang et al. [166] studied the mechanism of NH3-SCR reaction of NO on CuMn2O4 catalyst. Density functional theory calculation showed that the dehydrogenation of NH3 was the decisive step of N2 formation. Wu et al. [167] used hydrotalcite-like precursor to induce the modification of NiO by titanium dioxide (Figure 10a), which produced a good interface effect in the calcination process (Figure 10b). The active components were dispersed at atomic level, and the conversion rate was over 90% in the temperature range of 240 to 360 °C (Figure 10c,d).
MnOx-based catalysts are currently the focus of low-temperature NH3-SCR flue gas denitrification technology research [168,169,170]. As an active component, MnOx can provide free electrons, which play an important role in the SCR process [171]. Wang et al. [172] prepared MnOx catalyst with Na2CO3 as precipitator has a high specific surface area and amorphous framework structure. Zhang et al. [173] synthesized Fe modified MnOx/TiO2 catalyst by sol–gel method; when the flue gas was simulated at 250 °C, the NO conversion reached 76.8%. Wang et al. [174] developed a MnOx-CeO2-Al2O3 catalyst with enhanced activity and good water resistance by using instantaneous nano-precipitation technology. Xiang et al. [175] discussed the adsorption mechanism of NH3 and NO on MnOx-based catalysts supported on γ-Al2O3. Wang et al. [176] studied the performance of a series of W-modified MnOx/TiO2 catalysts for NH3-SCR. In situ infrared spectroscopy showed that more Brønsted and Lewis acid sites were formed on the catalyst surface after tungsten doping. Zhang et al. [177] studied the effect of Ce-Mo modified TiO2 adsorbent on NO adsorption. Zhao et al. [178] prepared two-dimensional layered MnAl double oxides by instantaneous nanoprecipitation. Wang et al. [179] prepared MnOx-CeO2-Al2O3 catalysts with different MnOx content by self-propagating high-temperature synthesis firstly, and the effects of different synthesis methods on NO conversion were compared in Table 4.
The powder catalyst in the laboratory stage is quite mature, meaning it can efficiently remove nitrogen oxides from simulated flue gas. However, how to apply these excellent catalysts to practice is a question worthy of consideration. Obviously, powder catalysts must be prepared as monolith catalysts to meet the needs of stationary industrial installations. Monolith catalysts have many dominant positions, such as strong mechanical stability, thermal conductivity, mass transfer capacity, small pressure drop and recycling, which are conducive to the catalytic process and practical application.
Yu Feng research group of Shihezi University has studied the monolith catalyst deeply. Tian et al. [185] prepared nanoporous microspheres Mn–Ce–Fe–Ti mixed oxide catalysts for NH3-SCR by spray drying (Figure 11a). The samples were processed by focused ion beam (FIB) and observed by scanning electron microscopy (Figure 11b,c). After application in flue gas treatment, it was found that the monolith catalyst still maintained good catalytic activity. Wang et al. [159] first designed and prepared spherical MnOx-CeO2-Al2O3 powder catalysts by spray drying, then applied it to monolith honeycomb catalysts(MHC). Compared with the same metal oxides catalysts prepared by coprecipitation (CP-MHC), the spray-drying (SD-MHC) method exhibited excellent SCR denitrification performance at 50–150 °C (Figure 11d,e).

4.2.2. CO-SCR

CO-SCR denitrification technology can effectively remove NO and CO from flue gas, thus performing two steps simultaneously [186]. In addition, CO-SCR as reducing agent can simultaneously remove NO and CO from tail gas and reduce the cost of purchase, transportation and storage of additional reducing agent. It is an attractive and promising denitrification technology. Therefore, it is urgent to study CO-SCR technology [187,188,189]. Besides, precious metals are expensive, have limited reserves and have poor high temperature performance. They sinter at about 800 °C. At the same time, they have poor sulfur and phosphorus resistance, so it is particularly important to develop non-precious metal catalysts. Current research focuses on rare earth elements and cheap transition metals, rare earth elements, especially Ce, are widely studied in cheap transition metals, such as Cu, Fe, Mn, etc. [190,191,192].
In order to further explore the performance of CO-SCR denitrification, Shen et al. [187] found that the operating temperature window moved towards low temperature after Fe co-doping with Mn–Ce/TiO2. Yang et al. [193] studied the application of gas sensor in denitrification.Chen et al. [194] studied the CO+NO reaction on Pd/Al2O3 and Pd/CeZrO2 catalysts. Cheng et al. [195] prepared Pd/LaFeO3 nanoparticle catalysts supported on LaFeO3 by two-step precipitation-deposition method. The visible light-induced valence exchange of Fe2+ and Fe3+ can promote the formation of oxygen vacancies. Lopes et al. [196] synthesized Cu-Ni catalyst by coprecipitation method. Bánsági et al. [197] found that isocyanate was the intermediate in the process of CO-SCR reaction. Wang et al. [198] studied the reaction of CO-SCR and the decomposition of supported metals by TG-FTIR. Liu et al. [190] prepared Cu0.1La0.1Ce0.8O catalyst, which had good water resistance and catalytic stability in CO-SCR process. Yamashita et al. [199] studied the CO-SCR reaction of nickel species supported on silica. Yao et al. [200] discussed the mechanism of CeO2 modified MnOx catalyst at low temperature (Figure 12a). Phase interfaces or synergies were formed on the surface (Figure 12b). Jin et al. [201] further studied the sulfur resistance and water resistance of Mn-Mo-WOx/TiO2-SiO2 catalysts. It was found that the catalytic activity of Mn-Mo-WOx/TiO2-SiO2 catalysts decreased significantly when H2O and SO2 existed simultaneously (Figure 12c,d), and the effects of different synthesis methods on NO conversion were compared in Table 5.

4.2.3. HC-SCR

In recent years, selective catalytic reduction of hydrocarbons (HC-SCR) has become a hot research topic worldwide. It has potential advantages in large-scale flue gas emission control, diesel engine and lean combustion engine exhaust gas treatment. At present, many studies have focused on Ag/Al2O3 catalytic system. Compared with the transition metal catalysts represented by Cu, Ag/Al2O3 catalysts have higher NO conversion and N2 selectivity, outstanding thermal stability, water sulfur resistance, which makes it to be one of the most promising catalysts for HC-SCR.
H2 can significantly promote the HC-SCR by hydrocarbons. Xu et al. [202] studied the sulfur resistance of Ag/Al2O3 catalyst assisted by H2. It was found that Ag/Al2O3 catalyst with higher Ag loading had better denitrification and sulfur resistance, especially 4% Ag/Al2O3 catalyst, which was attributed to the rapid migration of sulfate on its surface to form more active sites. Thomas et al. [203] found that in the presence of H2, the stability of nitrate was greatly reduced, and H2 greatly reduced the temperature of formation and decomposition of NOx. Xu et al. [204] found that Ag/Al2O3 catalyst had excellent water resistance in the whole temperature range. After introducing water vapor, especially at low temperature, the conversion of NOx was enhanced. Chaieb et al. [205] studied the reaction mechanism of H2 assisting C3H6 to remove NOx on Ag-loaded Al2O3 surface (Figure 13a). When the Ag-loaded density reached 0.7Ag/nm2Al2O3, the contact process between NO and C3H6 was the decisive step of C3H6-SCR process (Figure 13b). Xi et al. [206] prepared Fe-Ag/Al2O3 catalyst loaded on cordierite by sol–gel method and impregnation method. After Fe modification, the catalyst surface became loose and porous, forming Fe3O4-based needle-like and flaky crystals. It was found that Fe could effectively improve the performance of the catalyst against SO2 and H2O (Figure 13c,d), and increased the catalyst Lewis acid sites.

4.2.4. H2-SCR

H2 is the most clean and resource-rich reductant for selective catalytic removal of NOx. Compared with traditional NH3 and urea as reductant, hydrogen as reductant has the advantages of low reaction temperature, clean and no secondary pollution, and low temperature activity of H2-SCR catalyst.
Platinum and palladium, as unique active components, have excellent denitrification performance in the H2-SCR system. Wang et al. [207] prepared Pt/HZSM-5 catalyst by impregnation method. FTIR analysis showed that Pt on the outer surface of molecular sieve acted as an activation agent for H2 in H2-SCR. Xue et al. [208] directly deposited Pt with different content on MIL-96 (Al) surface by hydrothermal method. Pt was slightly loaded on MIL-96 (Al) surface and H2-SCR catalytic activity was excellent at 80 °C. Zhang et al. [209] modified H2-SCR on Pt/HZSM-5 catalyst by tungsten doping. The results showed that tungsten was beneficial to the metal state of platinum in the catalyst and inhibited the formation of nitrate. At the same time, tungsten accelerated the dissociation of nitric oxide on the platinum surface. Väliheikki et al. [210] found that tungsten doping could improve the catalytic activity through the acidity of CezZr1−zO2 surface on H2-SCR catalyst. Yang et al. [211] studied the catalytic performance of platinum supported on H-ferrous acid. In situ infrared spectroscopy showed that NO2 and nitrate were not important intermediates of H2-SCR on the catalyst, while NOδ+ on Pt was an effective material for NO removal. Huai et al. [212] studied the decomposition process of NO on Pd (111) surface and the formation pathways of N2, NH3, N2O, and H2O (Figure 14a,b). The specific temperatures of them could be obtained by relative selectivity analysis at different temperatures (Figure 14c,d).
Removal of NOx by reduction is an extraordinary efficient method for removing NOx. However, SCR is limited by a series of problems, such as high cost, high temperature at 300 °C, NH3 escaping from secondary pollution environment [213], poor sulfur and water resistance of catalysts and susceptibility to poisoning and inactivation [214,215], which affect the life of the catalysts, makes the treatment and operation cost significantly higher [216,217]. Therefore, how to promote the resource-saving and environmentally friendly denitrification process is a difficult problem to be solved urgently in the world [218]. Studying the reaction mechanism of NO on catalyst surface is of great significance for solving the problem of catalyst poisoning and deactivation; improving catalytic denitrification activity; promoting dust, H2O, and SO2 resistance; enlarging operation temperature window; and promoting the low temperature direction of reaction [219,220,221].

5. Removal of NOx by Plasma

Plasma was discovered by German physicist and chemist Krux in 1879 and was first introduced into physics by American scientists Langmuir and Tonks. The plasma is the fourth form of matter, which contains a large number of electrons, ions, molecules, neutral atoms, excited state atoms, photons, and free radicals, among which positive and negative ions have the same charge [222]. Strictly speaking, plasma is a gas mass with high potential kinetic energy. The total charged quantity of plasma is still neutral. The outer electrons are shot out by the high kinetic energy of electric field or magnetic field. As a result, electrons are no longer bound to the nucleus, but become free electrons with high potential and high kinetic energy [223,224]. The plasma is a good conductive fluid, and there is no net Coulomb force or net magnetic force between charged particles. The charged particles are positive and negative in microscopy and electrically neutral in macro, and have certain thermal effects. According to the different thermodynamic equilibrium states, plasma can be divided into thermodynamic equilibrium plasma and non-thermodynamic equilibrium plasma [225,226]. Thermodynamically balanced plasma, also known as high temperature plasma, has the same temperature of molecule, electron and ion [227]. Non-thermodynamic equilibrium plasma is also called cryogenic plasma, in which the electron temperature is 10,000–100,000 K, the temperature of ions and neutral particles is much lower than that of electrons, the temperature is only 300–500 K, and the temperature of the whole system is lower [228,229,230]. Low temperature plasma has been widely used in many fields due to its unique characteristics [231]. According to the treatment of NOx by plasma, it can be divided into three parts: direct decomposition of NOx by plasma, plasma-modified catalyst and plasma assisted catalyst.

5.1. Direct Decomposition of NOx by Plasma

The plasma can not only directly decompose organic compounds in sewage, but treat nitrogen oxides directly [232]. The study of plasma direct decomposition of NOx dated back to 1997. Oda et al. [233] used dielectric barrier discharge to decompose NOx in automobile exhaust directly. They found that NOx was more easily oxidized in the presence of oxygen. Baeva et al. [234,235] also found the same phenomenon when he decomposed NOx by pulsed microwave discharge plasma at atmospheric pressure. Hueso et al. [236,237] used microwave-induced plasma to study the decomposition of NO in Ar and N2. In the presence of oxygen, if enough reductive gases were added, the decomposition of NO was still dominant. Schütz et al. [238] introduced that dielectric barrier discharge can be realized in air atmosphere, discharge uniformly, produce high concentration of isomers, and can be effectively utilized. There are many factors affecting dielectric barrier discharge, such as material of dielectric barrier, discharge voltage, discharge frequency, discharge distance, etc. [239]. Wu et al. [240] proposed a new method of corona discharge coupled with wet absorption, which can achieve simultaneous desulfurization and denitrification. Chen et al. [241] studied the reaction pathway and mechanism of NO removal by nonthermal plasma NTP for the first time by FT-IR (Figure 15a). It was found by FT-IR analysis and recording that O· played the role of NO· + O· → NO2 oxidation (Figure 15b). The effect of plasma on NO removal in Ar, O2, N2, and other atmospheres was compared. The sulfur and water resistance were also studied, NO3 and SO42− were formed as the byproducts, which greatly attenuate the catalytic activity (Figure 15c,d).

5.2. Plasma-Modified Catalyst

NOx removal by plasma-modified catalysts is a new green and environmental technology. The modification of catalysts by plasma is mainly focused on the surface modification of catalysts. The dispersion of active species can be improved by plasma modification, and the stability and low temperature activity of catalysts can be improved [242]. Hong et al. [243] studied the denitrification performance of pulsed corona discharge plasma. The results showed that there was an obvious interaction between the process of desulfurization and denitrification. Zhang et al. [244] treated NiO-TiO2-Al2O3 catalyst with plasma. It was found that the specific surface area of the catalyst was increased, the nickel particles dispersed more evenly and the ability of the catalyst to adsorb NOx was enhanced. Huang et al. [245] prepared V2O5/ACF catalyst. After oxygen plasma modification, the surface oxygen-containing functional groups increased, the pore size distribution widened and the catalytic activity increased. Liu et al. [246] prepared a Mn-O-Ce catalyst by nonthermal plasma, it decomposes nitrate and organic matter slowly and forms defective Mn-O-Ce phase in the catalyst during plasma treatment. More details are shown in Table 6.
The plasma contains high energy, which can provide enough activation energy for many difficult reactions. An et al. [247] studied the oxidation characteristics of NO in oxygen atmosphere by surface dielectric barrier discharge plasma treatment of mixed flue gas with active material injection method. Cui et al. [248] prepared a MnCe/Ti catalyst in a dielectric barrier discharge reactor by combining nonthermal plasma discharge with catalyst. The highest NO conversion rate was 86.9%. Zhao et al. [249] developed a new process of denitrification by water-cooled dielectric barrier discharge. Peng et al. [250] used nonthermal plasma and bamboo charcoal combined with adsorption technology to catalyze the removal of NO. SEM showed that the surface of the catalyst after plasma treatment would produce defects to promote the growth of pore structure and increase the active sites. Park et al. [251] removed both NO and SO2 from flue gas through a wet NaClO2 scrubber combined with a plasma electrostatic precipitator. Wang et al. [252] prepared Mn–Ce/ZSM5 multi-walled carbon nanotubes catalyst by impregnation (Figure 16a). It was found that more oxygen holes could be generated after plasma treatment, which could significantly improve NO conversion (Figure 14b). Among them, Mn1Ce1/ZSM5 has the highest catalytic activity and its reaction accorded with Eley–Rideal mechanism (Figure 16c). The sulfur and water resistance of Mn1Ce1/ZSM5 were also slightly improved (Figure 16d).

5.3. Plasma Concerted Catalysis

Plasma concerted catalysis is the combination of plasma and catalyst, which makes them act synergistically on the reaction system and optimize the reaction process. According to the combination of catalysts and plasma, plasma-assisted catalysis can be divided into plasma-enhanced catalysis and plasma-driven catalysis. (1) Plasma-enhanced catalysis: catalysts work at the back of the plasma and the plasma and catalyst work in the optimum temperature range respectively. Under this combination mode, the role of plasma is to change the gas composition reaching the catalyst bed. Reactive gases are treated by plasma to produce more highly active intermediates, which react on catalysts to produce final products [253]. (2) Plasma-driven catalysis: plasma and catalyst are placed in the same reactor, and a large number of short-lived active species produced by plasma change the shape of gaseous reactants on the catalyst. The discharge of plasma on the catalyst surface may also change the properties of the catalyst, such as the generation of electron holes and electrons, and the surface of the catalyst. The change of work function, the activation of lattice oxygen of catalyst, the formation of new active sites on catalyst surface, etc. [254]. Tang et al. [255] studied the effect of nonthermal plasma co-catalytic treatment on the low-temperature oxidation of NO by Mn-CoOx catalyst (Figure 17a,b). It was found that the activity of the catalyst treated by nonthermal plasma increased significantly (Figure 17c), and the specific surface area and pore volume of the catalyst increased significantly after nonthermal plasma treatment, as well as the advantages of changing the relative surface concentration and oxidation state of the species on the catalyst surface. Two main weightlessness peaks were studied (Figure 17d).
To further demonstrate the concerted effect of plasma, the following datas are listed to enhance persuasion. Niu et al. [256] found that the denitrification rate was greatly increased when the plasma acted with Co-ZSM-5 zeolite catalyst. Bröer et al. [257] found that the activity of SCR at low temperature increased significantly under the action of plasma with V2O5-WO3/TiO2 catalyst. Fan [258] studied the concerted effect of plasma and H2 mordenite in C2H2-SCR. Li et al. [259] explored the enhanced effect of Mn-Co-CeOx catalyst with plasma treatment in NH3-SCR. Cho et al. [260] discussed the concerted effect in plasma-enhanced catalysis. Using Ba & Cu (C2H2O3)2 as catalyst, NO was oxidized to NO2 in plasma, while hydrocarbon was oxidized to active intermediate species CO, which eventually reduced NOx to N2. Pan et al. [261] found that the conversion of NOx as well as the water and sulfur resistance of C3H8-SCR plasma increased significantly when dielectric barrier plasma acted simultaneously. Wang et al. [262,263] found that the denitrification rate was very low when the plasma acted alone. When the plasma cooperated with the Mn-Cu/ZSM5 zeolite catalyst, the conversion of NO was greatly increased. The above data are summarized in Table 7.
In recent years, the form of plasma has been innovated rapidly [264,265]. Dielectric barrier discharge, radio frequency discharge, and microwave discharge are all new nonthermal equilibrium plasma discharge forms [266,267]. The effect of plasma denitrification is remarkable, which can improve the NO conversion rapidly on short notice. However, compared with the traditional denitrification technology, the plasma process has made some progress, but there are still some problems: high energy consumption, short power supply life and performance to be improved; moreover, the expensive price of plasma equipment, high cost of system operation and maintenance, and complex equipment structure make the technology not widely used and restrict its development.

6. Conclusions and Prospect

Along with the huge global emission of NOx, which continues to deteriorate planet earth environment and human health in the Anthropocene, it becomes progressively significant to collect and summarize the efficient approaches of governance and treatment of NOx. Optimizing the above four distinct denitrification methods and controlling the emission of NOx are the main challenges for future atmospheric environmental treatment. Finally, in order to solve the problem of air pollution that plagues all countries in the world, the feasibility of four main technologies in the field of flue gas denitrification and industrial application of flue gas tail purify is evaluated.
Adsorption treatment of NOx has the features of environmental friendliness, energy-saving, and high efficiency. The process is simple to operate, controllable, for low concentration poisonous gases, efficient capture can be achieved. However, due to their limited adsorption capacity, frequent regeneration of adsorbents, and huge size and the large-scale investment required, the application of adsorbents is limited. Oxidation process is a relatively mature method in the field of wet denitrification technology because of its simple route, easy operation and outstanding denitrification effect. Some of them have been industrialized. However, the oxidation tower has the problems of expensive raw materials and equipment corrosion, which limits its development due to its cost and safety. Reduction is the most efficient and mature denitrification technology with the most extensive application as well as the most exceptional means to control NOx pollution. But at the same time, it has some shortcomings, high investment, high cost of catalyst regeneration, NH3 escaping from secondary pollution, and poor sulfur and water resistance of catalyst and so on. The removal of NOx by plasma technology is remarkable, which can rapidly improve the denitrification performance in a short time, and promote the stability and low temperature activity of the catalyst. However, the required equipment covers a large area, consumes high energy, and costs high investment, operation and maintenance, which is limited in practical application.
Future, the general requirement of denitrification technology is low cost, high efficiency and green. The overall development trend of technology is to realize the coordinated removal of multiple pollutants by coupling multiple technologies. Different regions and industries are suitable for different denitrification means. In order to decrease the emission level of NOx, cut down the expense of treatment and recovery, and improve the economic performance, different regions and industries should choose appropriate denitrification technology according from the resource situation to product usage. Therefore, it is of great practical significance to develop denitrification methods with high efficiency, low energy consumption, low secondary pollution, and low investment.

Author Contributions

F.Y. and B.D. provided framework. Z.L. summarizes the content of the article. C.M., J.D. and J.L. collected information.

Funding

The work was supported by Major Scientific and Technological Project of Bingtuan (No. 2018AA002), the National Natural Science Foundation of China (21663022), and Science and Technology Innovation Talents Program of Bingtuan (No. 2019CB025)

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Seitzinger, S.P.; Phillips, L. Nitrogen stewardship in the Anthropocene. Science 2017, 357, 350–351. [Google Scholar] [CrossRef] [PubMed]
  2. Tvw, J.; Pnr, V. A molecular dance to cleaner air. Science 2017, 357, 866–867. [Google Scholar]
  3. Boyle, E. Nitrogen pollution knows no bounds. Science 2017, 356, 700. [Google Scholar] [CrossRef] [PubMed]
  4. Anenberg, S.C.; Miller, J.; Minjares, R.; Du, L.; Henze, D.K.; Lacey, F.; Malley, C.S.; Emberson, L.; Franco, V.; Klimont, Z. Impacts and mitigation of excess diesel-related NOx emissions in 11 major vehicle markets. Nature 2017, 545, 467. [Google Scholar] [CrossRef] [PubMed]
  5. Lelieveld, J.; Evans, J.S.; Fnais, M.; Giannadaki, D.; Pozzer, A. The contribution of outdoor air pollution sources to premature mortality on a global scale. Nature 2015, 525, 367–371. [Google Scholar] [CrossRef]
  6. McDonald, B.C.; de Gouw, J.A.; Gilman, J.B.; Jathar, S.H.; Akherati, A.; Cappa, C.D.; Jimenez, J.L.; Lee-Taylor, J.; Hayes, P.L.; McKeen, S.A.; et al. Volatile chemical products emerging as largest petrochemical source of urban organic emissions. Science 2018, 359, 760. [Google Scholar] [CrossRef] [PubMed]
  7. Zhao, C.N.; Xu, Z.; Wu, G.C.; Mao, Y.M.; Liu, L.N.; Qian, W.; Dan, Y.L.; Tao, S.S.; Zhang, Q.; Sam, N.B.; et al. Emerging role of air pollution in autoimmune diseases. Autoimmun. Rev. 2019, 18, 507–614. [Google Scholar] [CrossRef]
  8. Schraufnagel, D.E.; Balmes, J.R.; Cowl, C.T.; De Matteis, S.; Jung, S.-H.; Mortimer, K.; Perez-Padilla, R.; Rice, M.B.; Riojas-Rodriguez, H.; Sood, A.; et al. Air Pollution and Noncommunicable Diseases: A Review by the Forum of International Respiratory Societies’ Environmental Committee, Part 2: Air Pollution and Organ Systems. Chest 2019, 155, 417–426. [Google Scholar] [CrossRef]
  9. Guo, T.; Wang, Y.; Zhang, H.; Zhang, Y.; Zhao, J.; Wang, Q.; Shen, H.; Wang, Y.; Xie, X.; Wang, L.; et al. The association between ambient PM2.5 exposure and the risk of preterm birth in China: A retrospective cohort study. Sci. Total Environ. 2018, 633, 1453–1459. [Google Scholar] [CrossRef]
  10. Zeng, Q.; Darboux, F.; Man, C.; Zhu, Z.; An, S. Soil aggregate stability under different rain conditions for three vegetation types on the Loess Plateau (China). CATENA 2018, 167, 276–283. [Google Scholar] [CrossRef] [Green Version]
  11. Du, E.; Dong, D.; Zeng, X.; Sun, Z.; Jiang, X.; de Vries, W. Direct effect of acid rain on leaf chlorophyll content of terrestrial plants in China. Sci. Total Environ. 2017, 605, 764–769. [Google Scholar] [CrossRef] [PubMed]
  12. Bang, H.Q.; Nguyen, H.D.; Vu, K.; Hien, T.T. Photochemical Smog Modelling Using the Air Pollution Chemical Transport Model (TAPM-CTM) in Ho Chi Minh City, Vietnam. Environ. Model. Assess. 2019, 24, 295–310. [Google Scholar] [CrossRef]
  13. Lewis, A.C. The changing face of urban air pollution. Science 2018, 359, 744. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, C.; Wang, W.; Sardans, J.; An, W.; Zeng, C.; Abid, A.A.; Peñuelas, J. Effect of simulated acid rain on CO2, CH4 and N2O fluxes and rice productivity in a subtropical Chinese paddy field. Environ. Pollut. 2018, 243, 1196–1205. [Google Scholar] [CrossRef] [PubMed]
  15. Song, Y.; Guo, S.; Zhang, M. Assessing customers’ perceived value of the anti-haze cosmetics under haze pollution. Sci. Total Environ. 2019, 685, 753–762. [Google Scholar] [CrossRef] [PubMed]
  16. Chong, X.; Wang, Y.; Liu, R.; Zhang, Y.; Zhang, Y.; Zheng, W. Pollution characteristics and source difference of gaseous elemental mercury between haze and non-haze days in winter. Sci. Total Environ. 2019, 678, 671–680. [Google Scholar] [CrossRef] [PubMed]
  17. Li, C.; Wang, Y.; Jia, B.; Roskilly, A.P. Application of Miller cycle with turbocharger and ethanol to reduce NOx and particulates emissions from diesel engine—A numerical approach with model validations. Appl. Therm. Eng. 2019, 150, 904–911. [Google Scholar] [CrossRef]
  18. Gu, Y.; Epling, W.S. Passive NOx adsorber: An overview of catalyst performance and reaction chemistry. Appl. Catal. Gen. 2019, 570, 1–14. [Google Scholar] [CrossRef]
  19. Huang, R.; Zhang, S.; Ding, J.; Meng, Y.; Zhong, Q.; Kong, D.; Gu, C. Effect of adsorption properties of phosphorus-doped TiO2 nanotubes on photocatalytic NO removal. J. Colloid Interface Sci. 2019, 553, 647–654. [Google Scholar] [CrossRef] [PubMed]
  20. Hirai, T.; Okoshi, M.; Ishikawa, A.; Nakai, H. Temperature- and pressure-dependent adsorption configuration of NO molecules on Rh(111) surface: A theoretical study. Surf. Sci. 2019, 686, 58–62. [Google Scholar] [CrossRef]
  21. Wang, T.; Ding, L.; Song, Y.; Cuiqing, L.I.; Wang, H.; Ren, X. Performance of simultaneous denitrification and desulfurization for Y2O3/AC catalysts. Chin. J. Environ. Eng. 2016, 4, 132–139. [Google Scholar]
  22. Liu, Y.; Zhang, X.; Ding, J. Chemical effect of NO on CH4 oxidation during combustion in O2/NO environments. Chem. Phys. Lett. 2019, 727, 59–65. [Google Scholar] [CrossRef]
  23. Zhang, T.; Li, H.; Yang, Z.; Cao, F.; Li, L.; Chen, H.; Liu, H.; Xiong, K.; Wu, J.; Hong, Z.; et al. Electrospun YMn2O5 nanofibers: A highly catalytic activity for NO oxidation. Appl. Catal. Environ. 2019, 247, 133–141. [Google Scholar] [CrossRef]
  24. Jo, D.; Park, G.T.; Ryu, T.; Hong, S.B. Economical synthesis of high-silica LTA zeolites: A step forward in developing a new commercial NH3-SCR catalyst. Appl. Catal. Environ. 2019, 243, 212–219. [Google Scholar] [CrossRef]
  25. Chitpakdee, C.; Junkaew, A.; Maitarad, P.; Shi, L.; Promarak, V.; Kungwan, N.; Namuangruk, S. Understanding the role of Ru dopant on selective catalytic reduction of NO with NH3 over Ru-doped CeO2 catalyst. Chem. Eng. J. 2019, 369, 124–133. [Google Scholar] [CrossRef]
  26. Yang, W.; Gao, Z.; Liu, X.; Ma, C.; Ding, X.; Yan, W. Directly catalytic reduction of NO without NH3 by single atom iron catalyst: A DFT calculation. Fuel 2019, 243, 262–270. [Google Scholar] [CrossRef]
  27. Wang, D.; Cheng, J.; Wang, B.; Lou, J.; Li, Y.; Li, X.; Li, Z.; Liu, X.; Meng, Q.; Gao, P.; et al. Plasma-catalytic high-efficiency oxidation of NO over Co-Mn/Ti catalysts using surface dielectric barrier discharge plasma. Vacuum 2019, 167, 249–254. [Google Scholar] [CrossRef]
  28. Wang, J.; Yi, H.; Tang, X.; Zhao, S.; Gao, F. Exceptional adsorptive capacity and kinetic of γ-Al2O3 for selective adsorption of NO assisted by nonthermal plasma. Chem. Eng. J. 2019, 378, 122095. [Google Scholar] [CrossRef]
  29. Xie, S.; Ren, W.; Qiao, C.; Tong, K.; Sun, J.; Zhang, M.; Liu, X.; Zhang, Z. An electrochemical adsorption method for the reuse of waste water-based drilling fluids. Nat. Gas Ind. B 2018, 5, 508–512. [Google Scholar] [CrossRef]
  30. Ikeda, H.; Koike, Y.; Shiratori, K.; Ueda, K.; Shirahata, N.; Isegawa, K.; Toyoshima, R.; Masuda, S.; Mase, K.; Nito, T.; et al. Adsorption state of NO on Ir (111) surfaces under excess O2 coexisting condition. Surf. Sci. 2019, 685, 1–6. [Google Scholar] [CrossRef]
  31. Zhao, Y.; Wang, H.; Wang, T. Adsorption of NO from flue gas by molecularly imprinted adsorbents. Chem. Eng. J. 2016, 306, 832–839. [Google Scholar] [CrossRef]
  32. Hastürk, E.; Ernst, S.J.; Janiak, C. Recent advances in adsorption heat transformation focusing on the development of adsorbent materials. Curr. Opin. Chem. Eng. 2019, 24, 26–36. [Google Scholar] [CrossRef]
  33. Qu, W.; Yuan, T.; Yin, G.; Xu, S.; Zhang, Q.; Su, H. Effect of properties of activated carbon on malachite green adsorption. Fuel 2019, 249, 45–53. [Google Scholar] [CrossRef]
  34. Szczęśniak, B.; Choma, J.; Jaroniec, M. Ultrahigh benzene adsorption capacity of graphene-MOF composite fabricated via MOF crystallization in 3D mesoporous graphene. Microporous Mesoporous Mater. 2019, 279, 387–394. [Google Scholar] [CrossRef]
  35. Sherino, B.; Halim, S.N.A.; Manan, N.S.A.; Sarip, R.; Al’Abri, A.M.; Mohamad, S. Facile synthesis and characterization of novel dicarboxylate-Cu based MOFs materials. Inorg. Chim. Acta 2019, 491, 59–66. [Google Scholar] [CrossRef]
  36. Qin, Y.H.; Huang, L.; Zhang, L.; He, H. One-step synthesis of confined ion Agx-Cu-BTC for selective catalytic reduction of NO with CO. Inorg. Chem. Commun. 2019, 102, 130–133. [Google Scholar] [CrossRef]
  37. Khan, A.H.; Peikert, K.; Fröba, M.; Bertmer, M. NO adsorption in amino-modified Cu3(btc)2-type MOFs studied by solid-state NMR. Microporous Mesoporous Mater. 2015, 216, 111–117. [Google Scholar] [CrossRef]
  38. Kaur, R.; Kaur, A.; Umar, A.; Anderson, W.A.; Kansal, S.K. Metal organic framework (MOF) porous octahedral nanocrystals of Cu-BTC: Synthesis, properties and enhanced adsorption properties. Mater. Res. Bull. 2019, 109, 124–133. [Google Scholar] [CrossRef]
  39. Meng, G.; Song, X.; Ji, M.; Hao, J.; Shi, Y.; Ren, S.; Qiu, J.; Hao, C. Molecular simulation of adsorption of NO and CO2 mixtures by a Cu-BTC metal organic framework. Curr. Appl. Phys. 2015, 15, 1070–1074. [Google Scholar] [CrossRef]
  40. Qin, Y.H.; Huang, L.; Zhang, D.L.; Sun, L.G. Mixed-node A-Cu-BTC and porous carbon based oxides derived from A-Cu-BTC as low temperature NO–CO catalyst. Inorg. Chem. Commun. 2016, 66, 64–68. [Google Scholar] [CrossRef]
  41. Li, Z.; Yu, Q.; Ying, C.; Fei, X.; Li, W.; Li, Y.; Chen, M. Adsorption properties of activated carbon from reed with a high adsorption capacity. Ecol. Eng. 2017, 102, 443–450. [Google Scholar]
  42. Vanraes, P.; Ghodbane, H.; Davister, D.; Wardenier, N.; Nikiforov, A.; Verheust, Y.P.; Swh, V.H.; Hamdaoui, O.; Vandamme, J.; Van, D.J. Removal of several pesticides in a falling water film DBD reactor with activated carbon textile: Energy efficiency. Water Res. 2017, 116, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Beijer, K.; Björlenius, B.; Shaik, S.; Lindberg, R.H.; Brunström, B.; Brandt, I. Removal of pharmaceuticals and unspecified contaminants in sewage treatment effluents by activated carbon filtration and ozonation: Evaluation using biomarker responses and chemical analysis. Chemosphere 2017, 176, 342–351. [Google Scholar] [CrossRef] [PubMed]
  44. Qu, G.; Kou, L.; Wang, T.; Liang, D.; Hu, S. Evaluation of activated carbon fiber supported nanoscale zero-valent iron for chromium (VI) removal from groundwater in a permeable reactive column. J. Environ. Manag. 2017, 201, 378. [Google Scholar] [CrossRef] [PubMed]
  45. Shimabuku, K.K.; Paige, J.M.; Lunaaguero, M.; Summers, R.S. Simplified Modeling of Organic Contaminant Adsorption by Activated Carbon and Biochar in the Presence of Dissolved Organic Matter and Other Competing Adsorbates. Environ. Sci. Technol. 2017, 51, 10031–10040. [Google Scholar] [CrossRef] [PubMed]
  46. Li, Y.; Duan, Y.; Hui, W.; Zhao, S.; Chen, M.; Meng, L.; Wei, H. Effects of Acidic Gases on Mercury Adsorption by Activated Carbon in Simulated Oxy-Fuel Combustion Flue Gas. Energy Fuels 2017, 31, 9745–9751. [Google Scholar] [CrossRef]
  47. Sawant, S.Y.; Munusamy, K.; Somani, R.S.; John, M.; Newalkar, B.L.; Bajaj, H.C. Precursor suitability and pilot scale production of super activated carbon for greenhouse gas adsorption and fuel gas storage. Chem. Eng. J. 2017, 315, 415–425. [Google Scholar] [CrossRef]
  48. Fałtynowicz, H.; Hodurek, P.; Kaczmarczyk, J.; Kułażyński, M.; Łukaszewicz, M. Hydrolysis of surfactin over activated carbon. Bioorg. Chem. 2019, 13, 145–149. [Google Scholar] [CrossRef]
  49. Sajjadi, S.A.; Mohammadzadeh, A.; Tran, H.N.; Anastopoulos, I.; Dotto, G.L.; Lopičić, Z.R.; Sivamani, S.; Rahmani-Sani, A.; Ivanets, A.; Hosseini-Bandegharaei, A. Efficient mercury removal from wastewater by pistachio wood wastes-derived activated carbon prepared by chemical activation using a novel activating agent. J. Environ. Manag. 2018, 223, 1001–1009. [Google Scholar] [CrossRef]
  50. Ouhammou, M.; Lahnine, L.; Mghazli, S.; Hidar, N.; Bouchdoug, M.; Jaouad, A.; Mandi, L.; Mahrouz, M. Valorisation of cellulosic waste basic cactus to prepare activated carbon. J. Saudi Soc. Agric. Sci. 2019, 18, 133–140. [Google Scholar] [CrossRef]
  51. Björklund, K.; Li, L.Y. Adsorption of organic stormwater pollutants onto activated carbon from sewage sludge. J. Environ. Manag. 2017, 197, 490–497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Benstoem, F.; Becker, G.; Firk, J.; Kaless, M.; Wuest, D.; Pinnekamp, J.; Kruse, A. Elimination of micropollutants by activated carbon produced from fibers taken from wastewater screenings using hydrothermal carbonization. J. Environ. Manag. 2018, 211, 278–286. [Google Scholar] [CrossRef] [PubMed]
  53. Pirsaheb, M.; Mohamadi, S.; Rahmatabadi, S.; Hossini, H.; Motteran, F. Simultaneous wastewater treatment and biogas production using integrated anaerobic baffled reactor granular activated carbon from baker’s yeast wastewater. Environ. Technol. 2018, 39, 2724–2735. [Google Scholar] [CrossRef] [PubMed]
  54. Jahandar, L.M.; Atkinson, J.D.; Hashisho, Z.; Phillips, J.H.; Anderson, J.E.; Nichols, M. The role of beaded activated carbon’s surface oxygen groups on irreversible adsorption of organic vapors. J. Hazard. Mater. 2016, 317, 284–294. [Google Scholar] [CrossRef] [PubMed]
  55. Fu, Y.; Zhang, Y.; Li, G.; Zhang, J.; Guo, Y. NO removal activity and surface characterization of activated carbon with oxidation modification. J. Energy Inst. 2017, 90, 813–823. [Google Scholar] [CrossRef]
  56. Ren, S.; Guo, F.; Yang, J.; Yao, L.; Zhao, Q.; Kong, M. Selection of carbon materials and modification methods in low-temperature sintering flue gas denitrification. Chem. Eng. Res. Des. 2017, 126, 278–285. [Google Scholar] [CrossRef]
  57. Jie, W.; Lin, H.; Xue, H.; Ying, Z.; Wei, C. Physicochemical studies of adsorptive denitrogenation by oxidized activated carbons. Ind. Eng. Chem. Res. 2017, 56, 5033–5041. [Google Scholar]
  58. Chong-jiu, L.; Zhao, R.; Meng-qi, P.; Liu, H.; Gang, Y.U.; Xia, D.S. Study on desulfurization and denitrification by modified activated carbon fibers with visible-light photocatalysis. J. Fuel Chem. Technol. 2015, 43, 1516–1522. [Google Scholar]
  59. Arcibar-Orozco, J.A.; Acosta-Herrera, A.A.; Rangel-Mendez, J.R. Simultaneous desulfuration and denitrogenation of model diesel fuel by Fe-Mn microwave modified activated carbon: Iron crystalline habit influence on adsorption capacity. J. Clean. Prod. 2019, 218, 69–82. [Google Scholar] [CrossRef]
  60. You, F.T.; Yu, G.W.; Xing, Z.J.; Li, J.; Xie, S.Y.; Li, C.X.; Wang, G.; Ren, H.Y.; Wang, Y. Enhancement of NO catalytic oxidation on activated carbon at room temperature by nitric acid hydrothermal treatment. Appl. Surf. Sci. 2019, 471, 633–644. [Google Scholar] [CrossRef]
  61. Sun, X.; Sun, L.; Liu, Y.; Li, K.; Wang, C.; Song, X.; Ning, P. Research on reaction conditions and mechanism for simultaneous removal of NO and Hg0 over CuFe modified activated carbon at low temperature. J. Energy Inst. 2019, 6, 223–228. [Google Scholar]
  62. Wei, Z.S.; Zeng, G.H.; Xie, Z.R.; Ma, C.Y.; Liu, X.H.; Sun, J.L.; Liu, L.H. Microwave catalytic NOx and SO2 removal using FeCu/zeolite as catalyst. Fuel 2011, 90, 1599–1603. [Google Scholar] [CrossRef]
  63. Li, G.; Wang, B.; Sun, Q.; Xu, W.Q.; Ma, Z.; Wang, H.; Zhang, D.; Zhou, J. Novel synthesis of fly-ash-derived Cu-loaded SAPO-34 catalysts and their use in selective catalytic reduction of NO with NH3. Green Energy Environ. 2019, 32, 234–239. [Google Scholar] [CrossRef]
  64. Agote-Arán, M.; Lezcano-González, I.; Greenaway, A.G.; Hayama, S.; Díaz-Moreno, S.; Kroner, A.B.; Beale, A.M. Operando HERFD-XANES/XES studies reveal differences in the activity of Fe-species in MFI and CHA structures for the standard selective catalytic reduction of NO with NH3. Appl. Catal. A: Gen. 2019, 570, 283–291. [Google Scholar] [CrossRef]
  65. Chen, L.; Wang, X.; Cong, Q.; Ma, H.; Li, S.; Li, W. Design of a hierarchical Fe-ZSM-5@CeO2 catalyst and the enhanced performances for the selective catalytic reduction of NO with NH3. Chem. Eng. J. 2019, 369, 957–967. [Google Scholar] [CrossRef]
  66. Baran, R.; Averseng, F.; Wierzbicki, D.; Chalupka, K.; Krafft, J.-M.; Grzybek, T.; Dzwigaj, S. Effect of postsynthesis preparation procedure on the state of copper in CuBEA zeolites and its catalytic properties in SCR of NO with NH3. Appl. Catal. A: Gen. 2016, 523, 332–342. [Google Scholar] [CrossRef]
  67. Ma, Y.; Li, J.; Liu, E.; Wan, J.; Hu, X.; Fan, J. High efficiency for H2 evolution and NO removal over the Ag nanoparticles bridged g-C3N4 and WS2 heterojunction photocatalysts. Appl. Catal. B: Environ. 2017, 219, 467–478. [Google Scholar] [CrossRef]
  68. Imai, S.; Miura, H.; Shishido, T. Selective catalytic reduction of NO with CO and C3H6 over Rh/NbOPO4. Catal. Today 2019, 332, 267–271. [Google Scholar] [CrossRef]
  69. Xiao, P.; Davis, R.C.; Ouyang, X.; Li, J.; Thomas, A.; Scott, S.L.; Zhu, J. Mechanism of NO reduction by CO over Pt/SBA-15. Catal. Commun. 2014, 50, 69–72. [Google Scholar] [CrossRef]
  70. Wang, L.; Cheng, X.; Wang, Z.; Ma, C.; Qin, Y. Investigation on Fe-Co binary metal oxides supported on activated semi-coke for NO reduction by CO. Appl. Catal. B: Environ. 2017, 201, 636–651. [Google Scholar] [CrossRef]
  71. Chen, W.; Li, Z.; Hu, F.; Qin, L.; Han, J.; Wu, G. In-situ DRIFTS investigation on the selective catalytic reduction of NO with NH3 over the sintered ore catalyst. Appl. Surf. Sci. 2018, 439, 75–81. [Google Scholar] [CrossRef]
  72. Chen, Q.; Zheng, J.; Wen, L.; Yang, C.; Zhang, L. A multi-functional-group modified cellulose for enhanced heavy metal cadmium adsorption: Performance and quantum chemical mechanism. Chemosphere 2019, 224, 509–518. [Google Scholar] [CrossRef] [PubMed]
  73. 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: Gen. 2018, 561, 158–167. [Google Scholar] [CrossRef]
  74. Ibusuki, T.; Takeuchi, K. Removal of low concentration nitrogen oxides through photoassisted heterogeneous catalysis. J. Mol. Catal. 1994, 88, 93–102. [Google Scholar] [CrossRef]
  75. Huang, J.; Huang, H.; Liu, L.; Jiang, H. Revisit the effect of manganese oxidation state on activity in low-temperature NO-SCR. Mol. Catal. 2018, 446, 49–57. [Google Scholar] [CrossRef]
  76. Jiang, L.; Liu, Q.; Ran, G.; Kong, M.; Ren, S.; Yang, J.; Li, J. V2O5-modified Mn-Ce/AC catalyst with high SO2 tolerance for low-temperature NH3-SCR of NO. Chem. Eng. J. 2019, 370, 810–821. [Google Scholar] [CrossRef]
  77. Lin, F.; Shao, J.; Tang, H.; Li, Y.; Wang, Z.; Chen, G.; Yuan, D.; Cen, K. Enhancement of NO oxidation activity and SO2 resistance over LaMnO3+δ perovskites catalysts with metal substitution and acid treatment. Appl. Surf. Sci. 2019, 479, 234–246. [Google Scholar] [CrossRef]
  78. Salman, A.U.R.; Enger, B.C.; Auvray, X.; Lødeng, R.; Menon, M.; Waller, D.; Rønning, M. Catalytic oxidation of NO to NO2 for nitric acid production over a Pt/Al2O3 catalyst. Appl. Catal. A: Gen. 2018, 564, 142–146. [Google Scholar] [CrossRef]
  79. Jia, J.; Ran, R.; Guo, X.; Wu, X.; Chen, W.; Weng, D. Enhanced low-temperature NO oxidation by iron-modified MnO2 catalysts. Catal. Commun. 2019, 119, 139–143. [Google Scholar] [CrossRef]
  80. Yuan, P.; Egedy, A.; Miskolczi, N.; Shen, B.; Wang, J.; Zhou, W.; Pan, Y.; Zhang, H. Oxidation removal of NO by in situ Fenton system: Factors and optimization. Fuel 2018, 233, 519–528. [Google Scholar] [CrossRef]
  81. Zhao, Y.; Hao, R.; Wang, T.; Yang, C. Follow-up research for integrative process of pre-oxidation and post-absorption cleaning flue gas: Absorption of NO2, NO and SO2. Chem. Eng. J. 2015, 273, 55–65. [Google Scholar] [CrossRef]
  82. Guo, L.; Han, C.; Zhang, S.; Zhong, Q.; Ding, J.; Zhang, B.; Zeng, Y. Enhancement effects of O2− and OH radicals on NOX removal in the presence of SO2 by using an O3/H2O2 AOP system with inadequate O3 (O3/NO molar ratio = 0.5). Fuel 2018, 233, 769–777. [Google Scholar] [CrossRef]
  83. Fei, L.; Lin, Q.; Bi, X.; Chan, L.Y.; Chen, S.; Yan, X.; Liu, Y.; Wang, X.; Zhang, Y.; Cen, C. Effect of lightning activities on surface atmospheric NO, O3 and submicron particles based on artificially triggered lightning technology: A case study. Atmos. Pollut. Res. 2019, 10, 105–110. [Google Scholar] [CrossRef]
  84. Hultén, A.H.; Nilsson, P.; Samuelsson, M.; Ajdari, S.; Normann, F.; Andersson, K. First evaluation of a multicomponent flue gas cleaning concept using chlorine dioxide gas—Experiments on chemistry and process performance. Fuel 2017, 210, 885–891. [Google Scholar] [CrossRef]
  85. Sun, W.Y.; Ding, S.L.; Zeng, S.S.; Su, S.J.; Jiang, W.J. Simultaneous absorption of NOx and SO2 from flue gas with pyrolusite slurry combined with gas-phase oxidation of NO using ozone. J. Hazard. Mater. 2011, 192, 124–130. [Google Scholar] [CrossRef] [PubMed]
  86. Mok, Y.S.; LEE, H.J. Removal of sulfur dioxide and nitrogen oxides by using ozone injection and absorption-reduction technique. Fuel Process. Technol. 2006, 87, 591–597. [Google Scholar] [CrossRef]
  87. Lin, F.; Wang, Z.; Ma, Q.; Yang, Y.; Whiddon, R.; Zhu, Y.; Cen, K. Catalytic deep oxidation of NO by ozone over MnOx loaded spherical alumina catalyst. Appl. Catal. B: Environ. 2016, 198, 100–111. [Google Scholar] [CrossRef]
  88. Han, C.; Zhang, S.; Guo, L.; Zeng, Y.; Li, X.; Shi, Z.; Zhang, Y.; Zhang, B.; Zhong, Q. Ehanced catalytic ozonation of NO over black-TiO2 catalyst under inadequate ozone (O3/NO molar ratio = 0.6). Chem. Eng. Res. Des. 2018, 136, 219–229. [Google Scholar] [CrossRef]
  89. Hao, R.; Mao, X.; Wang, Z.; Zhao, Y.; Wang, T.; Sun, Z.; Yuan, B.; Li, Y. A novel method of ultraviolet/NaClO2-NH4OH for NO removal: Mechanism and kinetics. J. Hazard. Mater. 2019, 368, 234–242. [Google Scholar] [CrossRef]
  90. Zhao, Y.; Guo, T.X.; Chen, Z.Y.; Du, Y.R. Simultaneous removal of SO2 and NO using M/NaClO2 complex absorbent. Chem. Eng. J. 2010, 160, 42–47. [Google Scholar] [CrossRef]
  91. Deshwal, B.R.; Si, H.L.; Jung, J.H.; Shon, B.H.; Lee, H.K. Study on the removal of NO_x from simulated flue gas using acidic NaClO_2 solution. J. Environ. Sci. 2008, 20, 33–38. [Google Scholar] [CrossRef]
  92. Hao, R.; Yang, S.; Bo, Y.; Yi, Z. Simultaneous desulfurization and denitrification through an integrative process utilizing NaClO2/Na2S2O8. Fuel Process. Technol. 2017, 159, 145–152. [Google Scholar] [CrossRef]
  93. Yang, S.; Pan, X.; Han, Z.; Zheng, D.; Yu, J.; Xia, P.; Liu, B.; Yan, Z. Nitrogen Oxide Removal from Simulated Flue Gas by UV-Irradiated Sodium Chlorite Solution in a Bench-Scale Scrubbing Reactor. Ind. Eng. Chem. Res. 2017, 56, 3671–3678. [Google Scholar] [CrossRef]
  94. Wang, W.; Guo, R.; Pan, W.; Hu, G. Low temperature catalytic oxidation of NO over different-shaped CeO2. J. Rare Earths 2018, 36, 588–593. [Google Scholar] [CrossRef]
  95. Zhao, S.; 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]
  96. Shang, H.; Li, Y.; Liu, J.; Tang, X.; Yang, J.; Li, J. CH4/N2 separation on methane molecules grade diameter channel molecular sieves with a CHA-type structure. Chin. J. Chem. Eng. 2018, 1, 102–110. [Google Scholar] [CrossRef]
  97. Yamane, Y.; Tanaka, H.; Miyahara, M.T. In silico synthesis of carbon molecular sieves for high-performance air separation. Carbon 2019, 141, 626–634. [Google Scholar] [CrossRef]
  98. Liu, X.; Jiang, S.; Li, H.; Yang, J.; Yang, Z.; Zhao, J.; Peng, H.; Shih, K. Elemental mercury oxidation over manganese oxide octahedral molecular sieve catalyst at low flue gas temperature. Chem. Eng. J. 2019, 356, 142–150. [Google Scholar] [CrossRef]
  99. Geng, Q.; Wang, L.J.; Yang, C.; Zhang, H.Y.; Zhao, Y.R.; Fan, H.L.; Huo, C. Room-temperature hydrogen sulfide removal with zinc oxide nanoparticle/molecular sieve prepared by melt infiltration. Fuel Process. Technol. 2019, 185, 26–37. [Google Scholar] [CrossRef]
  100. Dharmarathna, S.; King’ondu, C.K.; Pahalagedara, L.; Kuo, C.H.; Zhang, Y.; Suib, S.L. Manganese octahedral molecular sieve (OMS-2) catalysts for selective aerobic oxidation of thiols to disulfides. Appl. Catal. B: Environ. 2014, 147, 124–131. [Google Scholar] [CrossRef]
  101. Park, D.; Ju, Y.; Kim, J.H.; Ahn, H.; Lee, C.H. Equilibrium and kinetics of nitrous oxide, oxygen and nitrogen adsorption on activated carbon and carbon molecular sieve. Sep. Purif. Technol. 2019, 223, 63–80. [Google Scholar] [CrossRef] [Green Version]
  102. Sarmah, B.; Srivastava, R. Selective two-step synthesis of 2,5-diformylfuran from monosaccharide, disaccharide, and polysaccharide using H-Beta and octahedral MnO2 molecular sieves. Mol. Catal. 2019, 462, 92–103. [Google Scholar] [CrossRef]
  103. Hamaguchi, T.; Tanaka, T.; Takahashi, N.; Tsukamoto, Y.; Takagi, N.; Shinjoh, H. Low-temperature NO-adsorption properties of manganese oxide octahedral molecular sieves with different potassium content. Appl. Catal. B: Environ. 2016, 193, 234–239. [Google Scholar] [CrossRef]
  104. Tolaymat, T.; El Badawy, A.; Genaidy, A.; Abdelraheem, W.; Sequeira, R. Analysis of metallic and metal oxide nanomaterial environmental emissions. J. Clean. Prod. 2017, 143, 401–412. [Google Scholar] [CrossRef]
  105. Jõgi, I.; Erme, K.; Raud, J.; Laan, M. Oxidation of NO by ozone in the presence of TiO2 catalyst. Fuel 2016, 173, 45–51. [Google Scholar] [CrossRef]
  106. Wang, H.; Chen, H.; Wang, Y.; Lyu, Y.K. Performance and mechanism comparison of manganese oxides at different valence states for catalytic oxidation of NO. Chem. Eng. J. 2019, 361, 1161–1172. [Google Scholar] [CrossRef]
  107. Ma, S.J.; Wang, X.W.; Chen, T.; Yuan, Z.H. Effect of surface morphology on catalytic activity for NO oxidation of SmMn2O5 nanocrystals. Chem. Eng. J. 2018, 354, 191–196. [Google Scholar] [CrossRef]
  108. Cai, W.; Zhao, Y.; Chen, M.; Jiang, X.; Wang, H.; Ou, M.; Wan, S.; Zhong, Q. The formation of 3D spherical Cr-Ce mixed oxides with roughness surface and their enhanced low-temperature NO oxidation. Chem. Eng. J. 2017, 333, 414–422. [Google Scholar] [CrossRef]
  109. Wang, Z.; Huang, Y.; Ho, W.; Cao, J.; Shen, Z.; Lee, S.C. Fabrication of Bi2O2CO3/g-C3N4 heterojunctions for efficiently photocatalytic NO in air removal: In-situ self-sacrificial synthesis, characterizations and mechanistic study. Appl. Catal. B: Environ. 2016, 199, 123–133. [Google Scholar] [CrossRef]
  110. Ângelo, J.; Andrade, L.; Madeira, L.M.; Mendes, A. An overview of photocatalysis phenomena applied to NOx abatement. J. Environ. Manag. 2013, 129, 522–539. [Google Scholar] [CrossRef] [PubMed]
  111. Duan, Y.; Luo, J.; Zhou, S.; Mao, X.; Shah, M.W.; Wang, F.; Chen, Z.; Wang, C. TiO2-supported Ag nanoclusters with enhanced visible light activity for the photocatalytic removal of NO. Appl. Catal. B: Environ. 2018, 234, 206–212. [Google Scholar] [CrossRef]
  112. He, D.; Li, Y.; Wang, I.; Wu, J.; Yang, Y.; An, Q. Carbon wrapped and doped TiO2 mesoporous nanostructure with efficient visible-light photocatalysis for NO removal. Appl. Surf. Sci. 2017, 391, 318–325. [Google Scholar] [CrossRef]
  113. Yuan, Y.; Zhang, J.; Li, H.; Li, Y.; Zhao, Y.; Zheng, C. Simultaneous removal of SO2, NO and mercury using TiO2-aluminum silicate fiber by photocatalysis. Chem. Eng. J. 2012, 192, 21–28. [Google Scholar] [CrossRef]
  114. Jin, S.; Dong, G.; Luo, J.; Ma, F.; Wang, C. Improved photocatalytic NO removal activity of SrTiO3 by using SrCO3 as a new co-catalyst. Appl. Catal. B: Environ. 2018, 227, 24–34. [Google Scholar] [CrossRef]
  115. Huy, T.H.; Bui, D.P.; Kang, F.; Wang, Y.F.; Liu, S.H.; Thi, C.M.; You, S.J.; Chang, G.M.; Pham, V.V. SnO2/TiO2 nanotube heterojunction: The first investigation of NO degradation by visible light-driven photocatalysis. Chemosphere 2019, 215, 323–332. [Google Scholar] [CrossRef] [PubMed]
  116. Su, X.; Li, X.; Ma, L.; Fan, J. Formation and transformation of schwertmannite in the classic Fenton process. J. Environ. Sci. 2019, 82, 145–154. [Google Scholar] [CrossRef]
  117. Zhang, Y.; Liang, J.; Zhou, W.; Xiao, N. Comparison of Fenton and bismuth ferrite Fenton-like pretreatments of sugarcane bagasse to enhance enzymatic saccharification. Bioresour. Technol. 2019, 285, 121343. [Google Scholar] [CrossRef]
  118. Hao, R.; Mao, Y.; Mao, X.; Wang, Z.; Gong, Y.; Zhang, Z.; Zhao, Y. Cooperative removal of SO2 and NO by using a method of UV-heat/H2O2 oxidation combined with NH4OH-(NH4)2SO3 dual-area absorption. Chem. Eng. J. 2019, 365, 282–290. [Google Scholar] [CrossRef]
  119. Song, Z.; Wang, B.; Yu, J.; Ma, C.; Chen, T.; Yang, W.; Liu, S.; Sun, L. Effect of Ti doping on heterogeneous oxidation of NO over Fe3O4 (111) surface by H2O2: A density functional study. Chem. Eng. J. 2018, 354, 517–524. [Google Scholar] [CrossRef]
  120. Wu, B.; Zhang, S.; He, S.; Xiong, Y. Follow-up mechanism study on NO oxidation with vaporized H2O2 catalyzed by Fe2O3 in a fixed-bed reactor. Chem. Eng. J. 2019, 356, 662–672. [Google Scholar] [CrossRef]
  121. Wang, Y.; Liu, Y.; Liu, Y. Elimination of Nitric Oxide Using New Fenton Process Based on Synergistic Catalysis: Optimization and Mechanism. Chem. Eng. J. 2019, 372, 92–98. [Google Scholar] [CrossRef]
  122. Yi, Z.; Wen, X.; Guo, T.; Zhou, J.; Fuproc, J.; Process, W. Desulfurization and denitrogenation from flue gas using Fenton reagent. Fuel Process. Technol. 2014, 128, 54–60. [Google Scholar]
  123. Cheng, L.; Huang, J.; Ni, F. Generation kinetics of hydroxyl radicals by Fenton’s reagent. Tech. Equip. Environ. Pollut. Control 2003, 12, 189–195. [Google Scholar]
  124. Guo, R.T.; Pan, W.G.; Zhang, X.B.; Ren, J.X.; Qiang, J.; Xu, H.J.; Jiang, W. Removal of NO by using Fenton reagent solution in a lab-scale bubbling reactor. Fuel 2011, 90, 3295–3298. [Google Scholar] [CrossRef]
  125. Cui, R.; Yang, B.; Li, S.; Wang, J.; Ma, S. Heterogeneous Fenton catalysts prepared from modified-fly ash for NOx removal with H2O2. Catal. Commun. 2019, 119, 180–184. [Google Scholar] [CrossRef]
  126. Zhao, Y.; Han, Y.; Wang, T.; Sun, Z.; Fang, C. Simultaneous removal of SO2 and NO from flue gas using iron-containing polyoxometalates as heterogeneous catalyst in UV-Fenton-like process. Fuel 2019, 250, 42–51. [Google Scholar] [CrossRef]
  127. Yuan, P.; Mei, X.; Shen, B.; Lu, F.; Zhou, W.; Si, M.; Chakraborty, S. Oxidation of NO by in situ Fenton reaction system with dual ions as reagents. Chem. Eng. J. 2018, 351, 660–667. [Google Scholar] [CrossRef]
  128. Góra-Marek, K.; Brylewska, K.; Tarach, K.A.; Rutkowska, M.; Jabłońska, M.; Choi, M.; Chmielarz, L. IR studies of Fe modified ZSM-5 zeolites of diverse mesopore topologies in the terms of their catalytic performance in NH3-SCR and NH3-SCO processes. Appl. Catal. B: Environ. 2015, 179, 589–598. [Google Scholar] [CrossRef]
  129. Schiavoni, M.; Campisi, S.; Gervasini, A. Effect of Cu deposition method on silico aluminophosphate catalysts in NH3-SCR and NH3-SCO reactions. Appl. Catal. A: Gen. 2017, 543, 162–172. [Google Scholar] [CrossRef]
  130. Shi-Gu, Q.; Wang, Y.; Li, X. Effect of praseodymium substitution on La1-xPrxMnO3 (x = −0.4) perovskites and catalytic activity for NO oxidation. J. Phys. Chem. Solids 2019, 133, 52–58. [Google Scholar]
  131. Wang, H.; Xu, R.; Jin, Y.; Zhang, R. Zeolite structure effects on Cu active center, SCR performance and stability of Cu-zeolite catalysts. Catal. Today 2019, 327, 295–307. [Google Scholar] [CrossRef]
  132. Yu, S.; Lu, Y.; Gao, F.; Dong, L. Study on the Crystal Plane Effect of CuO/TiO2 Catalysts in NH3-SCR Reaction. Catal. Today 2019, 327, 235–245. [Google Scholar] [CrossRef]
  133. Ling, L.; Cao, Y.; Zhao, Z.; Liu, P.; Wang, B.; Zhang, R.; Li, D. Density functional theory calculations and analysis for the reduction of NO by H2 on Pd6/TiO2. Comput. Mater. Sci. 2018, 149, 182–190. [Google Scholar] [CrossRef]
  134. Jung, Y.; Pyo, Y.D.; Jang, J.; Kim, G.C.; Cho, C.P.; Yang, C. NO, NO2 and N2O emissions over a SCR using DOC and DPF systems with Pt reduction. Chem. Eng. J. 2019, 369, 1059–1067. [Google Scholar] [CrossRef]
  135. Locci, C.; Vervisch, L.; Farcy, B.; Domingo, P.; Perret, N. Selective Non-catalytic Reduction (SNCR) of Nitrogen Oxide Emissions: A Perspective from Numerical Modeling. Flow Turbul. Combust. 2018, 100, 301–340. [Google Scholar] [CrossRef]
  136. Hao, J.; Yu, W.; Lu, P.; Zhang, Y.; Zhu, X. The effects of Na/K additives and flyash on NO reduction in a SNCR process. Chemosphere 2015, 122, 213–218. [Google Scholar] [CrossRef] [PubMed]
  137. Fu, S.L.; Song, Q.; Yao, Q. Influence of CaO on urea pyrolysis in the selective non-catalytic reduction deNOx process. J. Anal. Appl. Pyrolysis 2017, 126, 397–404. [Google Scholar] [CrossRef]
  138. Chen, H.; Chen, D.Z.; Fan, S.; Hong, L.; Wang, D. SNCR De-NOx within a moderate temperature range using urea-spiked hydrazine hydrate as reductant. Chemosphere 2016, 161, 208–218. [Google Scholar] [CrossRef]
  139. Kang, Z.; Yuan, Q.; Zhao, L.; Dai, Y.; Sun, B.; Wang, T. Study of the performance, simplification and characteristics of SNCR de-NOx in large-scale cyclone separator. Appl. Therm. Eng. 2017, 123, 635–645. [Google Scholar] [CrossRef]
  140. Fu, S.L.; Song, Q.; Yao, Q. Mechanism study on the adsorption and reactions of NH3, NO, and O2 on the CaO surface in the SNCR deNOx process. Chem. Eng. J. 2016, 285, 137–143. [Google Scholar] [CrossRef]
  141. Fu, S.l.; Song, Q.; Yao, Q. Study on the catalysis of CaCO3 in the SNCR deNOx process for cement kilns. Chem. Eng. J. 2015, 262, 9–17. [Google Scholar] [CrossRef]
  142. Tan, L.; Guo, Y.; Liu, Z.; Feng, P.; Li, Z. An investigation on the catalytic characteristic of NOx reduction in SCR systems. J. Taiwan Inst. Chem. Eng. 2019, 65, 236–246. [Google Scholar]
  143. Odenbrand, C.U.I. CaSO4 deactivated V2O5-WO3/TiO2 SCR catalyst for a diesel power plant. Characterization and simulation of the kinetics of the SCR reactions. Appl. Catal. B: Environ. 2018, 234, 365–377. [Google Scholar] [CrossRef]
  144. Lu, J.; Zhou, Z.; Zhang, H.; Yang, Z. Influenced factors study and evaluation for SO2/SO3 conversion rate in SCR process. Fuel 2019, 245, 528–533. [Google Scholar] [CrossRef]
  145. Yu, S.; Lu, Y.; Cao, Y.; Wang, J.; Sun, B.; Gao, F.; Tang, C.; Dong, L. Composite catalytic systems: A strategy for developing the low temperature NH3-SCR catalysts with satisfactory SO2 and H2O tolerance. Catal. Today 2019, 327, 235–245. [Google Scholar] [CrossRef]
  146. Gao, F.; Tang, X.; Yi, H.; Zhao, S.; Wang, J.; Gu, T. Improvement of activity, selectivity and H2O&SO2-tolerance of micro-mesoporous CrMn2O4 spinel catalyst for low-temperature NH3-SCR of NOx. Appl. Surf. Sci. 2019, 466, 411–424. [Google Scholar]
  147. Guo, M.; Liu, Q.; Zhao, P.; Han, J.; Li, X.; Ha, Y.; Fu, Z.; Song, C.; Ji, N.; Liu, C.; et al. Promotional effect of SO2 on Cr2O3 catalysts for the marine NH3-SCR reaction. Chem. Eng. J. 2019, 361, 830–838. [Google Scholar] [CrossRef]
  148. Niu, C.; Niu, J.; Wang, S.; Wang, Z.; Dong, S.; Fan, H.; Hong, Y.; Liu, D. Synergistic effect in one-stage dielectric barrier discharge plasma and Ag/Al2O3 catalytic systems on C2H2-SCR of NOx. Catal. Commun. 2019, 123, 49–53. [Google Scholar] [CrossRef]
  149. Sun, X.; Guo, R.T.; Liu, S.W.; Liu, J.; Pan, W.G.; Shi, X.; Qin, H.; Wang, Z.Y.; Qiu, Z.Z.; Liu, X.Y. The promoted performance of CeO2 catalyst for NH3-SCR reaction by NH3 treatment. Appl. Surf. Sci. 2018, 462, 187–193. [Google Scholar] [CrossRef]
  150. Li, X.; Li, X.; Yang, R.T.; Mo, J.; Li, J.; Hao, J. The poisoning effects of calcium on V2O5-WO3/TiO2 catalyst for the SCR reaction: Comparison of different forms of calcium. Mol. Catal. 2017, 434, 16–24. [Google Scholar] [CrossRef]
  151. Jung, Y.; Shin, Y.J.; Pyo, Y.D.; Cho, C.P.; Jang, J.; Kim, G. NOx and N2O emissions over a Urea-SCR system containing both V2O5-WO3/TiO2 and Cu-zeolite catalysts in a diesel engine. Chem. Eng. J. 2017, 326, 853–862. [Google Scholar] [CrossRef]
  152. Liang, Q.M.; Li, J.; He, H.; Liang, W.J.; Zhang, T.J.; Fan, X. Effects of SO2 on the low temperature selective catalytic reduction of NO by NH3 over CeO2-V2O5-WO3/TiO2 catalysts. Front. Environ. Sci. Eng. 2017, 11, 4. [Google Scholar] [CrossRef]
  153. Li, P.; Liu, Q.; Liu, Z. Behaviors of NH4HSO4 in SCR of NO by NH3 over different cokes. Chem. Eng. J. 2012, 181, 169–173. [Google Scholar] [CrossRef]
  154. Ye, D.; Qu, R.; Zheng, C.; Cen, K.; Gao, X. Mechanistic investigation of enhanced reactivity of NH4HSO4 and NO on Nb- and Sb-doped VW/Ti SCR catalysts. Appl. Catal. A: Gen. 2018, 549, 310–319. [Google Scholar] [CrossRef]
  155. Zhao, X.; Mao, L.; Dong, G. Mn-Ce-V-WOx/TiO2 SCR Catalysts: Catalytic Activity, Stability and Interaction among Catalytic Oxides. Catalysts 2018, 8, 76. [Google Scholar] [CrossRef]
  156. Tarot, M.L.; Barreau, M.; Duprez, D.; Lauga, V.; Iojoiu, E.; Courtois, X.; Can, F. Influence of the Sodium Impregnation Solvent on the Deactivation of Cu/FER-Exchanged Zeolites Dedicated to the SCR of NOx with NH3. Catalysts 2018, 8, 3. [Google Scholar] [CrossRef]
  157. Dong, L.; Fan, Y.; Ling, W.; Yang, C.; Huang, B. Effect of Ce/Y Addition on Low-Temperature SCR Activity and SO2 and H2O Resistance of MnOx/ZrO2/MWCNTs Catalysts. Catalysts 2017, 7, 181. [Google Scholar] [CrossRef]
  158. Shi, C.; Chang, H.; Wang, C.; Zhang, T.; Peng, Y.; Li, M.; Wang, Y.; Li, J. Improved Activity and H2O Resistance of Cu-Modified MnO2 Catalysts for NO Oxidation. Ind. Eng. Chem. Res. 2018, 57, 920–926. [Google Scholar] [CrossRef]
  159. Wang, C.; Yu, F.; Zhu, M.; Wang, X.; Dan, J.; Zhang, J.; Cao, P.; Dai, B. Microspherical MnO2-CeO2-Al2O3 mixed oxide for monolithic honeycomb catalyst and application in selective catalytic reduction of NOx with NH3 at 50–150 °C. Chem. Eng. J. 2018, 346, 182–192. [Google Scholar] [CrossRef]
  160. Gao, C.; Shi, J.-W.; Fan, Z.; Gao, G.; Niu, C. Sulfur and Water Resistance of Mn-Based Catalysts for Low-Temperature Selective Catalytic Reduction of NOx: A Review. Catalysts 2018, 8, 11. [Google Scholar] [CrossRef]
  161. Zhang, K.; Yu, F.; Zhu, M.; Dan, J.; Wang, X.; Zhang, J.; Dai, B. Enhanced Low Temperature NO Reduction Performance via MnOx-Fe2O3/Vermiculite Monolithic Honeycomb Catalysts. Catalysts 2018, 8, 100. [Google Scholar] [CrossRef]
  162. Yu, T.; Hao, T.; Fan, D.; Wang, J.; Shen, M.; Li, W. Recent NH3-SCR mechanism research over Cu/SAPO-34 catalyst. J. Phys. Chem. C 2014, 118, 6565–6575. [Google Scholar] [CrossRef]
  163. Woo, J.; Leistner, K.; Bernin, D.; Ahari, H.; Shost, M.; Zammit, M.; Olsson, L. Effect of various structure directing agents (SDAs) on low-temperature deactivation of Cu/SAPO-34 during NH3-SCR reaction. Catal. Sci. Technol. 2018, 8, 3–4. [Google Scholar] [CrossRef]
  164. Ryu, T.; Kim, H.; Hong, S.B. Nature of active sites in Cu-LTA NH3-SCR catalysts: A comparative study with Cu-SSZ-13. Appl. Catal. B: Environ. 2019, 245, 513–521. [Google Scholar] [CrossRef]
  165. Campisi, S.; Galloni, M.G.; Bossola, F.; Gervasini, A. Comparative performance of copper and iron functionalized hydroxyapatite catalysts in NH3-SCR. Catal. Commun. 2019, 123, 79–85. [Google Scholar] [CrossRef]
  166. Yang, Y.; Liu, J.; Liu, F.; Wang, Z.; Ding, J.; Huang, H. Reaction mechanism for NH3-SCR of NOx over CuMn2O4 catalyst. Chem. Eng. J. 2019, 361, 578–587. [Google Scholar] [CrossRef]
  167. Wu, X.; Wang, R.; Du, Y.; Zou, C.; Meng, H.; Xie, X. Performance enhancement of NH3-SCR via employing hydrotalcite-like precursor to induce the decoration of NiO by TiO2 phase. Mol. Catal. 2019, 467, 150–160. [Google Scholar] [CrossRef]
  168. Wu, W.; Zeng, Z.; Lu, P.; Xing, Y.; Wei, J.; Yue, H.; Li, R. Simultaneous oxidation of Hg0 and NH3-SCR of NO by nanophase CexZryMnzO2 at low temperature: The interaction and mechanism. Environ. Sci. Pollut. Res. 2018, 25, 14471–14485. [Google Scholar] [CrossRef]
  169. Xu, H.; Zan, Q.; Zong, C.; Quan, F.; Jian, M.; Yan, N. Catalytic oxidation and adsorption of Hg0 over low-temperature NH3-SCR LaMnO3 perovskite oxide from flue gas. Appl. Catal. B Environ. 2016, 186, 30–40. [Google Scholar] [CrossRef]
  170. Tian, J.; Wang, C.; Yu, F.; Zhou, X.; Zhang, J.; Yang, S.; Dan, J.; Cao, P.; Dai, B.; Wang, Q.; et al. Mn-Ce-Fe-Al mixed oxide nanoparticles via a high shear mixer facilitated coprecipitation method for low temperature selective catalytic reduction of NO with NH3. Appl. Catal. A: Gen. 2019, 586, 117237. [Google Scholar] [CrossRef]
  171. Wang, C.; Yu, F.; Zhu, M.Y.; Tang, C.J.; Dong, L.; Dai, B. Both powdered and performed of MnOx-CeO2-Al2O3 catalysts synthesized by self-propagating high-temperature synthesis for the selective catalystic reduction of NOx with NH3. ACS Omega 2018, 3, 5692–5703. [Google Scholar] [CrossRef] [PubMed]
  172. Wang, C.; Xia, W.; Zhao, Y. New insight into hydroxyl-mediated NH3 formation on the Rh-CeO2 catalyst surface during catalytic reduction of NO by CO. Chin. J. Catal. 2017, 38, 1399–1405. [Google Scholar] [CrossRef]
  173. Zhang, S.; Zhao, Y.; Yang, J.; Zhang, J.; Zheng, C. Fe-modified MnOx/TiO2 as the SCR catalyst for simultaneous removal of NO and mercury from coal combustion flue gas. Chem. Eng. J. 2018, 348, 618–629. [Google Scholar] [CrossRef]
  174. Wang, C.; Yu, F.; Zhu, M.; Shi, Y.; Dan, J.; Lv, Y.; Guo, X.; Dai, B. Up-scaled flash nano-precipitation production route to develop a MnOx–CeO2–Al2O3 catalyst with enhanced activity and H2O resistant performance for NOx selective catalytic reduction with NH3. Chem. Eng. Res. Des. 2018, 134, 476–486. [Google Scholar] [CrossRef]
  175. Xiang, J.; Wang, L.; Cao, F.; Qian, K.; Su, S.; Hu, S.; Wang, Y.; Liu, L. Adsorption properties of NO and NH3 over MnOx based catalyst supported on γ-Al2O3. Chem. Eng. J. 2016, 302, 570–576. [Google Scholar] [CrossRef]
  176. Wang, X.; Li, X.; Zhao, Q.; Sun, W.; Tade, M.; Liu, S. Improved activity of W-modified MnOx–TiO2 catalysts for the selective catalytic reduction of NO with NH3. Chem. Eng. J. 2016, 288, 216–222. [Google Scholar] [CrossRef]
  177. Zhang, P.; Pan, W.G.; Guo, R.T.; Zhu, X.B.; Liu, J.; Qin, L.; She, X.L. The Mo modified Ce/TiO2 catalyst for simultaneous Hg0 oxidation and NO reduction. J. Energy Inst. 2018, 44, 224–254. [Google Scholar] [CrossRef]
  178. Zhao, D.; Wang, C.; Yu, F.; Shi, Y.; Cao, P.; Dan, J.; Chen, K.; Lv, Y.; Guo, X.; Dai, B. Enhanced Oxygen Vacancies in a Two-Dimensional MnAl-Layered Double Oxide Prepared via Flash Nanoprecipitation Offers High Selective Catalytic Reduction of NOx with NH3. Nanomaterials 2018, 8, 620. [Google Scholar] [CrossRef]
  179. Wang, C.; Yu, F.; Zhu, M.; Tang, C.; Zhang, K.; Zhao, D.; Dong, L.; Dai, B. Highly selective catalytic reduction of NOx by MnOx–CeO2–Al2O3 catalysts prepared by self-propagating high-temperature synthesis. J. Environ. Sci. 2019, 75, 124–135. [Google Scholar] [CrossRef]
  180. Liu, Z.; Liu, Y.; Yuan, L.; Hang, S.; Ma, L. WO3 promoted Mn–Zr mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Acs Appl Mater. Interfaces 2016, 283, 1044–1050. [Google Scholar]
  181. Shin, B.; Chun, H.H.; Cha, J.S.; Shin, M.C.; Lee, H. Physico-Chemical Property and Catalytic Activity of a CeO2-Doped MnO(x)-TiO2 Catalyst with SO2 Resistance for Low-Temperature NH3-SCR of NO(x). J. Nanosci. Nanotechnol. 2016, 16, 4370. [Google Scholar] [CrossRef] [PubMed]
  182. Gao, F.; Tang, X.; Yi, H.; Li, J.; Zhao, S.; Wang, J.; Chao, C.; Li, C. Promotional mechanisms of activity and SO2 tolerance of Co- or Ni-doped MnOx-CeO2 catalysts for SCR of NOx with NH3 at low temperature. Chem. Eng. J. 2017, 317, 20–31. [Google Scholar] [CrossRef]
  183. Cao, F.; Su, S.; Xiang, J.; Wang, P.Y.; Hu, S.; Sun, L.S.; Zhang, A.C. The activity and mechanism study of Fe-Mn-Ce/gamma-Al2O3 catalyst for low temperature selective catalytic reduction of NO with NH3. Fuel 2015, 139, 232–239. [Google Scholar] [CrossRef]
  184. Zhang, Q.; Qiu, C.; Xu, H.; Tao, L.; Lin, Z.; Gong, M.; Chen, Y. Low-temperature selective catalytic reduction of NO with NH3 over monolith catalyst of MnOx/CeO2–ZrO2–Al2O3. Catal. Today 2012, 175, 171–176. [Google Scholar] [CrossRef]
  185. Tian, J.; Zhang, K.; Wang, W.; Wang, F.; Dan, J.; Yang, S.; Zhang, J.; Dai, B.; Yu, F. Enhanced selective catalytic reduction of NO with NH3 via porous micro-spherical aggregates of Mn–Ce–Fe–Ti mixed oxide nanoparticles. Green Energy Environ. 2019, 4, 311–321. [Google Scholar] [CrossRef]
  186. Cheng, X.; Cheng, Y.; Wang, Z.; Ma, C. Comparative study of coal based catalysts for NO adsorption and NO reduction by CO. Fuel 2018, 214, 230–241. [Google Scholar] [CrossRef]
  187. Shen, B.; Zhu, S.; Zhang, X.; Chi, G.; Patel, D.; Si, M.; Wu, C. Simultaneous removal of NO and Hg0 using Fe and Co co-doped Mn-Ce/TiO2 catalysts. Fuel 2018, 224, 241–249. [Google Scholar] [CrossRef]
  188. Singh, S.; Bhatia, D. Modeling the role of CO and C3H6 in NOx reduction on a Cu-CHA SCR catalyst. Chem. Eng. J. 2018. [Google Scholar] [CrossRef]
  189. Zheng, Y.; Harold, M.P.; Luss, D. Effects of CO, H2 and C3H6 on Cu-SSZ-13 catalyzed NH3-SCR. Catal. Today 2016, 264, 44–54. [Google Scholar] [CrossRef]
  190. Liu, T.; Qian, J.; Yao, Y.; Shi, Z.; Han, L.; Liang, C.; Li, B.; Dong, L.; Fan, M.; Zhang, L. Research on SCR of NO with CO over the Cu0.1La0.1Ce0.8O mixed-oxide catalysts: Effect of the grinding. Mol. Catal. 2017, 430, 43–53. [Google Scholar] [CrossRef]
  191. Boningari, T.; Pavani, S.M.; Ettireddy, P.R.; Chuang, S.S.C.; Smirniotis, P.G. Mechanistic investigations on NO reduction with CO over Mn/TiO2 catalyst at low temperatures. Mol. Catal. 2018, 451, 33–42. [Google Scholar] [CrossRef]
  192. Panahi, P.N.; Salari, D.; Niaei, A.; Mousavi, S.M. NO reduction over nanostructure M-Cu/ZSM-5 (M: Cr, Mn, Co and Fe) bimetallic catalysts and optimization of catalyst preparation by RSM. J. Ind. Eng. Chem. 2013, 19, 1793–1799. [Google Scholar] [CrossRef]
  193. Yang, D.; Fan, X.; Zhao, D.; An, Y.; Hu, Y.; Luo, Z. Sc2CO2 and Mn-doped Sc2CO2 as gas sensor materials to NO and CO: A first-principles study. Phys. E: Low-Dimens. Syst. Nanostruct. 2019, 111, 84–90. [Google Scholar] [CrossRef]
  194. Chen, X.; Lyu, Y.; Nwabara, U.; Schwank, J.W. Reactivity study of CO+NO reaction over Pd/Al2O3 and Pd/CeZrO2 catalysts. Catal. Today 2019, 323, 148–158. [Google Scholar] [CrossRef]
  195. Cheng, G.; Tan, X.; Song, X.; Chen, X.; Dai, W.; Yuan, R.; Fu, X. Visible light assisted thermocatalytic reaction of CO + NO over Pd/LaFeO3. Appl. Catal. B: Environ. 2019, 251, 130–142. [Google Scholar] [CrossRef]
  196. Lopes, D.; Zotin, F.; Palacio, L.A. Copper-nickel catalysts from hydrotalcite precursors: The performance in NO reduction by CO. Appl. Catal. B: Environ. 2018, 237, 327–338. [Google Scholar] [CrossRef]
  197. Bánsági, T.; Zakar, T.S.; Solymosi, F. An FTIR study on the formation of NCO surface complexes over Rh/CeO2. Appl. Catal. B: Environ. 2006, 66, 147–150. [Google Scholar] [CrossRef]
  198. Wang, L.; Cheng, X.; Wang, Z.; Huang, H.; Ma, C.; Qin, Y. Effect of the NO + CO reaction on the consumption of carbon supports: An in situ TG-FTIR analysis. Chem. Eng. J. 2018, 352, 90–102. [Google Scholar] [CrossRef]
  199. Yamashita, S.; Yamamoto, Y.; Kawabata, H.; Niwa, Y.; Katayama, M.; Inada, Y. Dynamic chemical state conversion of nickel species supported on silica under CO–NO reaction conditions. Catal. Today 2018, 303, 33–39. [Google Scholar] [CrossRef]
  200. Yao, X.; Li, L.; Zou, W.; Yu, S.; An, J.; Li, H.; Yang, F.; Dong, L. Preparation, characterization, and catalytic performance of high efficient CeO2-MnOx-Al2O3 catalysts for NO elimination. Chin. J. Catal. 2016, 37, 1369–1380. [Google Scholar] [CrossRef]
  201. Jin, Q.; Shen, Y.; Sui, G.; Tao, X.; Pan, Y.; Zhu, S. Synergistic catalytic removals of NO, CO and HC over CeO2 modified Mn-Mo-W-Ox/TiO2-SiO2 catalyst. J. Rare Earths 2018, 36, 148–155. [Google Scholar] [CrossRef]
  202. Xu, G.; Ma, J.; Wang, L.; Xie, W.; Liu, J.; Yu, Y.; He, H. Insight into the origin of sulfur tolerance of Ag/Al2O3 in the H2-C3H6-SCR of NOx. Appl. Catal. B: Environ. 2019, 244, 909–918. [Google Scholar] [CrossRef]
  203. Thomas, C. On an additional promoting role of hydrogen in the H2-assisted C3H6-SCR of NOx on Ag/Al2O3: A lowering of the temperature of formation–decomposition of the organo-NOx intermediates? Appl. Catal. B: Environ. 2015, 162, 454–462. [Google Scholar] [CrossRef]
  204. Xu, G.; Ma, J.; He, G.; Yu, Y.; He, H. An alumina-supported silver catalyst with high water tolerance for H2 assisted C3H6-SCR of NOx. Appl. Catal. B: Environ. 2017, 207, 60–71. [Google Scholar] [CrossRef]
  205. Chaieb, T.; Delannoy, L.; Costentin, G.; Louis, C.; Casale, S.; Chantry, R.L.; Li, Z.Y.; Thomas, C. Insights into the influence of the Ag loading on Al2O3 in the H2-assisted C3H6-SCR of NOx. Appl. Catal. B: Environ. 2014, 156, 192–201. [Google Scholar] [CrossRef]
  206. Yang, X.; Su, Y.X.; Qian, W.Y.; Yuan, M.H.; Zhou, H.; Deng, W.Y.; Zhao, B.T. Experimental study on selective catalytic reduction of NO by C3H6 over Fe-Ag/Al2O3 catalysts. J. Fuel Chem. Technol. 2017, 45, 1365–1375. [Google Scholar] [CrossRef]
  207. Wang, X.; Wang, X.; Yu, H.; Wang, X. The functions of Pt located at different positions of HZSM-5 in H2-SCR. Chem. Eng. J. 2019, 355, 470–477. [Google Scholar] [CrossRef]
  208. Xue, Y.; Sun, W.; Wang, Q.; Cao, L.; Yang, J. Sparsely loaded Pt/MIL-96(Al) MOFs catalyst with enhanced activity for H2-SCR in a gas diffusion reactor under 80 °C. Chem. Eng. J. 2018, 335, 612–620. [Google Scholar] [CrossRef]
  209. Zhang, X.; Wang, X.; Zhao, X.; Xu, Y.; Liu, Y.; Yu, Q. Promotion effect of tungsten on the activity of Pt/HZSM-5 for H2-SCR. Chem. Eng. J. 2015, 260, 419–426. [Google Scholar] [CrossRef]
  210. Väliheikki, A.; Petallidou, K.C.; Kalamaras, C.M.; Kolli, T.; Huuhtanen, M.; Maunula, T.; Keiski, R.L.; Efstathiou, A.M. Selective catalytic reduction of NOx by hydrogen (H2-SCR) on WOx-promoted CezZr1-zO2 solids. Appl. Catal. B: Environ. 2014, 156, 72–83. [Google Scholar] [CrossRef]
  211. Yang, S.; Wang, X.; Chu, W.; Song, Z.; Zhao, S. An investigation of the surface intermediates of H2-SCR of NOx over Pt/H-FER. Appl. Catal. B: Environ. 2011, 107, 380–385. [Google Scholar] [CrossRef]
  212. Huai, L.Y.; He, C.Z.; Wang, H.; Wen, H.; Yi, W.C.; Liu, J.Y. NO dissociation and reduction by H2 on Pd(1 1 1): A first-principles study. J. Catal. 2015, 322, 73–83. [Google Scholar] [CrossRef]
  213. Liu, G.; Cui, Y.; Ji, J.; Shen, D.; Wang, Q.; Li, C.; Luo, K.H. A technical method to improve NOx/NH3 mixing ratio in SCR system and its engineering applications. J. Energy Inst. 2018, 43, 243–249. [Google Scholar] [CrossRef]
  214. Kikugawa, M.; Yamazaki, K.; Kato, A.; Uyama, T.; Takahashi, N.; Shinjoh, H. Silver sulfate catalyst for soot oxidation with high resistance to sulfur poisoning. Appl. Catal. A: Gen. 2019, 576, 32–38. [Google Scholar] [CrossRef]
  215. Tian, Y.; Yang, J.; Liu, L.; Liu, Q.; Kong, B.; Lin, F.; Kong, M.; Hu, G. Insight into regeneration mechanism with sulfuric acid for arsenic poisoned commercial SCR catalyst. J. Energy Inst. 2019, 76, 214–223. [Google Scholar] [CrossRef]
  216. Wang, A.; Wang, Y.; Walter, E.D.; Washton, N.M.; Guo, Y.; Lu, G.; Peden, C.H.F.; Gao, F. NH3-SCR on Cu, Fe and Cu+Fe exchanged beta and SSZ-13 catalysts: Hydrothermal aging and propylene poisoning effects. Catal. Today 2019, 320, 91–99. [Google Scholar] [CrossRef]
  217. Hammershøi, P.S.; Jensen, A.D.; Janssens, T.V.W. Impact of SO2-poisoning over the lifetime of a Cu-CHA catalyst for NH3-SCR. Appl. Catal. B: Environ. 2018, 238, 104–110. [Google Scholar] [CrossRef]
  218. Li, H.; Wang, S.; Wang, X.; Hu, J. Activity of CuCl2-modified cobalt catalyst supported on Ti-Ce composite for simultaneous catalytic oxidation of Hg0 and NO in a simulated pre-sco process. Chem. Eng. J. 2017, 316, 1103–1113. [Google Scholar] [CrossRef]
  219. Salazar, M.; Hoffmann, S.; Singer, V.; Becker, R.; Grünert, W. Hybrid catalysts for the selective catalytic reduction (SCR) of NO by NH3. On the role of fast SCR in the reaction network. Appl. Catal. B: Environ. 2016, 199, 433–438. [Google Scholar] [CrossRef]
  220. Han, S.; Cheng, J.; Ye, Q.; Cheng, S.; Kang, T.; Dai, H. Ce doping to Cu-SAPO-18: Enhanced catalytic performance for the NH3-SCR of NO in simulated diesel exhaust. Microporous Mesoporous Mater. 2019, 276, 133–146. [Google Scholar] [CrossRef]
  221. Xu, Y.; Wu, X.; Lin, Q.; Hu, J.; Ran, R.; Weng, D. SO2 promoted V2O5-MoO3/TiO2 catalyst for NH3-SCR of NOx at low temperatures. Appl. Catal. A: Gen. 2019, 570, 42–50. [Google Scholar] [CrossRef]
  222. Langmuir, I. Oscillations in Ionized Gases. Proc. Natl. Acad. Sci. USA 1928, 14, 627–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Shahmansouri, M.; Misra, A.P. Surface plasmon oscillations in a semi-bounded semiconductor plasma. Plasma Sci. Technol. 2018, 20, 025001. [Google Scholar] [CrossRef] [Green Version]
  224. Parastaev, A.; Hoeben, W.F.L.M.; van Heesch, B.E.J.M.; Kosinov, N.; Hensen, E.J.M. Temperature-programmed plasma surface reaction: An approach to determine plasma-catalytic performance. Appl. Catal. B: Environ. 2018, 239, 168–177. [Google Scholar] [CrossRef]
  225. Shumyantseva, V.V.; Bulko, T.V.; Sigolaeva, L.V.; Kuzikov, A.V.; Pogodin, P.V.; Archakov, A.I. Molecular imprinting coupled with electrochemical analysis for plasma samples classification in acute myocardial infarction diagnostic. Biosens. Bioelectron. 2018, 99, 216–222. [Google Scholar] [CrossRef] [PubMed]
  226. Shabbir, A.; Hornung, G.; Noterdaeme, J.-M.; Verdoolaege, G. Classification of ELM types in Joint European Torus based on global plasma parameters using discriminant analysis. Fusion Eng. Des. 2017, 123, 717–721. [Google Scholar] [CrossRef] [Green Version]
  227. Gott, Y.V. Ion thermal conductivity in a tokamak plasma for the nonneoclassical distribution function (Review Paper). Plasma Phys. Rep. 1995, 21, 621–633. [Google Scholar]
  228. Capitelli, M.; Capitelli, F.; Eletskii, A. Non-equilibrium and equilibrium problems in laser-induced plasmas. Spectrochim. Acta Part. B At. Spectrosc. 2000, 55, 559–574. [Google Scholar] [CrossRef]
  229. Capitelli, M.; Armenise, I.; Bruno, D.; Cacciatore, M.; Celiberto, R.; Colonna, G.; Pascale, O.D.; Diomede, P.; Esposito, F.; Gorse, C. Non-equilibrium plasma kinetics: A state-to-state approach. Plasma Sources Sci. Technol. 2007, 16, S30–S44. [Google Scholar] [CrossRef]
  230. Lourek, I.; Tribeche, M. Dust charging current in non equilibrium dusty plasma in the context of Kaniadakis generalization. Phys. A: Stat. Mech. Appl. 2019, 517, 522–529. [Google Scholar] [CrossRef]
  231. Brzhozovskii, B.; Brovkova, M.; Gestrin, S.; Martynov, V.; Zinina, E. The effect of a combined low-pressure gas discharge on metal surfaces. J. Phys. D Appl. Phys. 2018, 51, 145204. [Google Scholar] [CrossRef]
  232. Li, Y.; Wang, W.; Wang, F.; Di, L.; Yang, S.; Zhu, S.; Yao, Y.; Dai, B.; Ma, C.; Yu, F. Enhanced photocatalytic degradation of organic dyes via defect-rich TiO2 prepared by dielectric barrier discharge plasma. Nanomaterials 2019, 9, 720. [Google Scholar] [CrossRef] [PubMed]
  233. Oda, T.; Kato, T.; Takahashi, T.; Shimizu, K. Nitric oxide decomposition in air by using nonthermal plasma processing with additives and catalyst. IEEE Trans. Ind. Appl. 1998, 34, 268–272. [Google Scholar] [CrossRef]
  234. Baeva, M.; Gier, H.; Pott, A.; Uhlenbusch, J.; H02schele, J.; Steinwandel, J. Erratum: “Studies on Gas Purification by a Pulsed Microwave Discharge at 2.46 GHz in Mixtures of N2/NO/O2 at Atmospheric Pressure” [Plasma Chemistry and Plasma Processing, Vol. 21, No. 2, 2001]. Plasma Chem. Plasma Process. 2002, 22, 197. [Google Scholar] [CrossRef]
  235. Baeva, M.; Gier, H.; Pott, A.; Uhlenbusch, J.; Höschele, J.; Steinwandel, J. Pulsed microwave discharge at atmospheric pressure for NOx decomposition. Plasma Sources Sci. Technol. 2002, 11, 1–9. [Google Scholar] [CrossRef]
  236. Hueso, J.L.; Gonzalez-Elipe, A.R.; Cotrino, J.; Caballero, A. Plasma chemistry of NO in complex gas mixtures excited with a surfatron launcher. J. Phys. Chem. A 2005, 109, 4930–4938. [Google Scholar] [CrossRef]
  237. Hueso, J.L.; González-Elipe, A.R.; Cotrino, J.; Caballero, A. Removal of NO in NO/N2, NO/N2/O2, NO/CH4/N2, and NO/CH4/O2/N2 Systems by Flowing Microwave Discharges. J. Phys. Chem. A 2007, 111, 1057–1065. [Google Scholar] [CrossRef]
  238. Schütz, A.; Lara-Ortega, F.J.; Klute, F.D.; Brandt, S.; Schilling, M.; Michels, A.; Veza, D.; Horvatic, V.; García-Reyes, J.F.; Franzke, J. Soft Argon-Propane Dielectric Barrier Discharge Ionization. Anal. Chem. 2018, 90, 3537–3542. [Google Scholar] [CrossRef]
  239. Brandt, S.; Klute, F.D.; Schütz, A.; Franzke, J. Dielectric barrier discharges applied for soft ionization and their mechanism. Anal. Chim. Acta 2017, 951, 16–31. [Google Scholar] [CrossRef]
  240. Zuliang, W.U.; Hou, P.; Zhao, J.; Wang, J.; Yan, X.U.; Yao, S. Experimental Study of Simultaneous Desulfurization and Denitrification by Corona Discharge Coupling Wet Absorption. High. Volt. Eng. 2016, 42, 398–404. [Google Scholar]
  241. Chen, S.; Wang, T.; Wang, H.; Wu, Z. Insights into the reaction pathways and mechanism of NO removal by SDBD plasma via FT-IR measurements. Fuel Process. Technol. 2019, 186, 125–136. [Google Scholar] [CrossRef]
  242. Zhang, Q.Z.; Bogaerts, A. Propagation of a plasma streamer in catalyst pores. Plasma Sources Sci. Technol. 2018, 27, 035009. [Google Scholar] [CrossRef]
  243. Hong, L.; Chen, D.; Yang, M.; Yin, L.; Wang, D.; Wang, L. Interaction between NO and SO2 removal processes in a pulsed corona discharge plasma (PCDP) reactor and the mechanism. Chem. Eng. J. 2019, 359, 1130–1138. [Google Scholar] [CrossRef]
  244. Zhang, H.; Li, K.; Li, L.; Liu, L.; Meng, X.; Sun, T.; Jia, J.; Fan, M. High efficient styrene mineralization through novel NiO-TiO2-Al2O3 packed pre-treatment/treatment/post-treatment dielectric barrier discharge plasma. Chem. Eng. J. 2018, 343, 759–769. [Google Scholar] [CrossRef]
  245. Huang, H.; Ye, D.; Huang, B.; Wei, Z. Vanadium Supported on Viscose-Based Activated Carbon Fibers Modified By Oxygen Plasma For The Scr Of No. Catal. Today 2008, 139, 100–108. [Google Scholar] [CrossRef]
  246. Liu, L.; Zheng, C.; Wu, S.; Gao, X.; Ni, M.; Cen, K. Manganese-cerium oxide catalysts prepared by non-thermal plasma for NO oxidation: Effect of O2 in discharge atmosphere. Appl. Surf. Sci. 2017, 416, 78–85. [Google Scholar] [CrossRef]
  247. An, J.; Jiang, Y.; Zhang, Z.; Ma, X.; Wang, T.; Shang, K.; Li, J. Oxidation characteristics of mixed NO and Hg0 in coal-fired flue gas using active species injection generated by surface discharge plasma. Chem. Eng. J. 2016, 288, 298–304. [Google Scholar] [CrossRef]
  248. Cui, S.; Hao, R.; Fu, D. Integrated method of non-thermal plasma combined with catalytical oxidation for simultaneous removal of SO2 and NO. Fuel 2019, 246, 365–374. [Google Scholar] [CrossRef]
  249. Zhao, D.; Yu, F.; Zhou, A.; Ma, C.; Dai, B. High-efficiency removal of NOx using dielectric barrier discharge nonthermal plasma with water as an outer electrode. Plasma Sci. Technol. 2018, 20, 014020. [Google Scholar] [CrossRef]
  250. Peng, M.; Zhao, R.; Xia, M.; Li, C.; Gong, X.; Wang, D.; Xia, D. Study on the mechanism of NO removal by plasma-adsorption catalytic process. Fuel 2017, 200, 290–298. [Google Scholar] [CrossRef]
  251. Park, H.W.; Choi, S.; Park, D.W. Simultaneous treatment of NO and SO2 with aqueous NaClO2 solution in a wet scrubber combined with a plasma electrostatic precipitator. J. Hazard. Mater. 2015, 285, 117–126. [Google Scholar] [CrossRef] [PubMed]
  252. Wang, T.; Liu, H.; Zhang, X.; Liu, J.; Zhang, Y.; Guo, Y.; Sun, B. Catalytic conversion of NO assisted by plasma over Mn-Ce/ZSM5-multi-walled carbon nanotubes composites: Investigation of acidity, activity and stability of catalyst in the synergic system. Appl. Surf. Sci. 2018, 457, 187–199. [Google Scholar] [CrossRef]
  253. Li, H.; Jiang, X.; Zheng, X. Plasma-assisted CuO/CeO2/TiO2-γ-Al2O3 catalysts for NO+CH4 reaction and NO temperature programmed desorption studies. Appl. Surf. Sci. 2013, 280, 273–281. [Google Scholar] [CrossRef]
  254. Dang, X.; Qin, C.; Huang, J.; Teng, J.; Huang, X. Adsorbed benzene/toluene oxidation using plasma driven catalysis with gas circulation: Elimination of the byproducts. J. Ind. Eng. Chem. 2016, 37, 366–371. [Google Scholar] [CrossRef]
  255. Tang, X.; Gao, F.; Xiang, Y.; Yi, H.; Zhao, S. Low temperature catalytic oxidation of nitric oxide over the Mn–CoOx catalyst modified by nonthermal plasma. Catal. Commun. 2015, 64, 12–17. [Google Scholar] [CrossRef]
  256. Niu, J.; Yang, X.; Zhu, A.; Shi, L.; Sun, Q.; Xu, Y.; Shi, C. Plasma-assisted selective catalytic reduction of NOx by C2 H2 over Co-HZSM-5 catalyst. Catal. Commun. 2006, 7, 297–301. [Google Scholar] [CrossRef]
  257. Broer, S.; Hammer, T. Selective catalytic reduction of nitrogen oxides by combining a non-thermal plasma and a V2O5-WO3/TiO2 catalyst. Appl. Catal. B: Environ. 2000, 28, 101–111. [Google Scholar] [CrossRef]
  258. Fan, H.Y.; Shi, C.; Li, X.S.; Yang, X.F.; Xu, Y.; Zhu, A.M. Low-temperature NOx Selective Reduction by Hydrocarbons on H-Mordenite Catalysts in Dielectric Barrier Discharge Plasma. Plasma Chem. Plasma Process. 2009, 29, 43–53. [Google Scholar] [CrossRef]
  259. Li, K.; Tang, X.; Yi, H.; Ning, P.; Song, J.; Wang, J. Mechanism of Catalytic Oxidation of NO over Mn–Co–Ce–Ox Catalysts with the Aid of Nonthermal Plasma at Low Temperature. Ind. Eng. Chem. Res. 2011, 50, 11023–11028. [Google Scholar] [CrossRef]
  260. Cho, B.K.; Lee, J.H.; Crellin, C.C.; Olson, K.L.; Hilden, D.L.; Kim, M.K.; Kim, P.S.; Heo, I.; Oh, S.H.; Nam, I.S. Selective catalytic reduction of NOx by diesel fuel: Plasma-assisted HC/SCR system. Catal. Today 2012, 191, 20–24. [Google Scholar] [CrossRef]
  261. Pan, H.; Su, Q.; Wei, J.; Jian, Y. Promotion of Nonthermal Plasma on the SO2 and H2O Tolerance of Co–In/Zeolites for the Catalytic Reduction of NOx by C3H8 at Low Temperature. Plasma Chem. Plasma Process. 2015, 35, 831–844. [Google Scholar] [CrossRef]
  262. Wang, T.; Zhang, X.; Liu, J.; Liu, H.; Wang, Y.; Sun, B. Effects of temperature on NOx removal with Mn-Cu/ZSM5 catalysts assisted by plasma. Appl. Therm. Eng. 2018, 130, 1224–1232. [Google Scholar] [CrossRef]
  263. Tao, W.; Liu, H.; Zhang, X.; Guo, Y.; Zhang, Y.; Yang, W.; Sun, B. A plasma-assisted catalytic system for NO removal over CuCe/ZSM-5 catalysts at ambient temperature. Fuel Process. Technol. 2017, 158, 199–205. [Google Scholar]
  264. Jõgi, I.; Stamate, E.; Irimiea, C.; Schmidt, M.; Brandenburg, R.; Hołub, M.; Bonisławski, M.; Jakubowski, T.; Kääriäinen, M.-L.; Cameron, D.C. Comparison of direct and indirect plasma oxidation of NO combined with oxidation by catalyst. Fuel 2015, 144, 137–144. [Google Scholar]
  265. Li, K.; Tang, X.; Yi, H.; Ning, P.; Xiang, Y.; Wang, J.; Wang, C.; Peng, X. Research on manganese oxide catalysts surface pretreated with non-thermal plasma for NO catalytic oxidation capacity enhancement. Appl. Surf. Sci. 2013, 264, 557–562. [Google Scholar] [CrossRef]
  266. Garraud, O.; Coppo, P. Types of fresh plasma with focus on therapeutic plasma exchange. Transfus. Apher. Sci. 2019, 58, 258–261. [Google Scholar] [CrossRef] [PubMed]
  267. Ibrahimoglu, B.; Yilmazoglu, M.Z. Numerical modeling of a downdraft plasma coal gasifier with plasma reactions. Int. J. Hydrog. Energy 2019, 67, 165–173. [Google Scholar] [CrossRef]
Figure 1. (a) Adsorption process and reaction mechanism of NO on A-Cu-BTC, (b) NO conversion with different metal dropping, (c) XRD images of metal doped Cu-BTC, and (d) SEM images of A-Cu-BTC [40]. Reproduced with the permission from Ref. [40], Copyright 2016, Elsevier.
Figure 1. (a) Adsorption process and reaction mechanism of NO on A-Cu-BTC, (b) NO conversion with different metal dropping, (c) XRD images of metal doped Cu-BTC, and (d) SEM images of A-Cu-BTC [40]. Reproduced with the permission from Ref. [40], Copyright 2016, Elsevier.
Catalysts 09 00771 g001
Figure 2. (a) Adsorption process of NO on activated carbon surface, (b) effect of adding O2 on NO conversion, (c) study on SO2 resistance of Cu-Fe/MCSAC catalyst, and (d) in situ infrared analysis of Cu-Fe/MCSAC after SO2 entering gas pool [61]. Reproduced from Ref. [61], Copyright 2019, Elsevier.
Figure 2. (a) Adsorption process of NO on activated carbon surface, (b) effect of adding O2 on NO conversion, (c) study on SO2 resistance of Cu-Fe/MCSAC catalyst, and (d) in situ infrared analysis of Cu-Fe/MCSAC after SO2 entering gas pool [61]. Reproduced from Ref. [61], Copyright 2019, Elsevier.
Catalysts 09 00771 g002
Figure 3. (a) Reaction mechanism of NOx removal by O3 in black TiO2, (b) de-NOx performance at 60 °C, (c) X-ray photoelectron spectroscopy (XPS) analysis under different conditions, and (d) sulfur resistance of black TiO2 under 1000 ppm SO2 [88]. Reproduced from Ref. [88], Copyright 2018, Elsevier.
Figure 3. (a) Reaction mechanism of NOx removal by O3 in black TiO2, (b) de-NOx performance at 60 °C, (c) X-ray photoelectron spectroscopy (XPS) analysis under different conditions, and (d) sulfur resistance of black TiO2 under 1000 ppm SO2 [88]. Reproduced from Ref. [88], Copyright 2018, Elsevier.
Catalysts 09 00771 g003
Figure 4. (a) Reaction mechanism of NO oxidation on OMS-2, (b) concentration changes of NO and NO2 with K-doping, (c) morphology of OMS-2, and (d) Kubelka–Mung function diagram [103]. Reproduced from Ref. [103], Copyright 2016, Elsevier.
Figure 4. (a) Reaction mechanism of NO oxidation on OMS-2, (b) concentration changes of NO and NO2 with K-doping, (c) morphology of OMS-2, and (d) Kubelka–Mung function diagram [103]. Reproduced from Ref. [103], Copyright 2016, Elsevier.
Catalysts 09 00771 g004
Figure 5. (a) Surface oxidation mechanism of 3D spherical Cr–Ce mixed oxides, (b) NO conversion of hydrothermal synthesis catalysts with different solvents, (c) HRTEM image of Cr–Ce/Et catalysts, and (d) sulfur and water resistance at 250–300 °C [108]. Reproduced from Ref. [108], Copyright 2018, Elsevier.
Figure 5. (a) Surface oxidation mechanism of 3D spherical Cr–Ce mixed oxides, (b) NO conversion of hydrothermal synthesis catalysts with different solvents, (c) HRTEM image of Cr–Ce/Et catalysts, and (d) sulfur and water resistance at 250–300 °C [108]. Reproduced from Ref. [108], Copyright 2018, Elsevier.
Catalysts 09 00771 g005
Figure 6. (a) Reaction mechanism of visible light removal of NOx by SnO2/TiO2 nanotube heterojunction, (b) removal efficiency of NOx under different conditions, (c) transmission electron microscopy of SnO2/TiO2 nanomaterials, and (d) detection process of free radicals under different conditions [115]. Reproduced from Ref. [115], Copyright 2019, Elsevier.
Figure 6. (a) Reaction mechanism of visible light removal of NOx by SnO2/TiO2 nanotube heterojunction, (b) removal efficiency of NOx under different conditions, (c) transmission electron microscopy of SnO2/TiO2 nanomaterials, and (d) detection process of free radicals under different conditions [115]. Reproduced from Ref. [115], Copyright 2019, Elsevier.
Catalysts 09 00771 g006
Figure 7. (a) Reaction mechanism of in situ Fenton system with two-ion reagents, (b) schematic diagram of in situ Fenton system, (c) effect of Ce-doping with different proportion on NO conversion, and (d) XRD was used to further explore the effect of Ce on the dispersion of Fe2+ [127]. Reproduced from Ref. [127], Copyright 2018, Elsevier.
Figure 7. (a) Reaction mechanism of in situ Fenton system with two-ion reagents, (b) schematic diagram of in situ Fenton system, (c) effect of Ce-doping with different proportion on NO conversion, and (d) XRD was used to further explore the effect of Ce on the dispersion of Fe2+ [127]. Reproduced from Ref. [127], Copyright 2018, Elsevier.
Catalysts 09 00771 g007
Figure 8. (a) Reaction mechanism of selective noncatalytic reduction (SNCR) on CaCO3 surface, (b) NH3 conversion and NO selectivity in different mixed gas, (c) FTIR analysis of CaCO3 surface active species, and (d) in situ FTIR analysis of reaction intermediates at 500 °C [140,141]. Reproduced from Ref. [140], Copyright 2016, Elsevier; Reproduced from Ref. [141], Copyright 2015, Elsevier.
Figure 8. (a) Reaction mechanism of selective noncatalytic reduction (SNCR) on CaCO3 surface, (b) NH3 conversion and NO selectivity in different mixed gas, (c) FTIR analysis of CaCO3 surface active species, and (d) in situ FTIR analysis of reaction intermediates at 500 °C [140,141]. Reproduced from Ref. [140], Copyright 2016, Elsevier; Reproduced from Ref. [141], Copyright 2015, Elsevier.
Catalysts 09 00771 g008
Figure 9. (a) Design route for De-NOx via NH3-SCR. (I) high-temperature at high dust content, (II) high-temperature at high SO2 content, (III) low-temperature at high H2O content, and (IV) low temperature with water removal. (b) Room temperature De-NOx via NH3-SCR. (c) Luo’s process for treating CO, NOx, and SOx [159]. Reproduced from Ref. [159], Copyright 2018, Elsevier.
Figure 9. (a) Design route for De-NOx via NH3-SCR. (I) high-temperature at high dust content, (II) high-temperature at high SO2 content, (III) low-temperature at high H2O content, and (IV) low temperature with water removal. (b) Room temperature De-NOx via NH3-SCR. (c) Luo’s process for treating CO, NOx, and SOx [159]. Reproduced from Ref. [159], Copyright 2018, Elsevier.
Catalysts 09 00771 g009
Figure 10. (a) Different preparation processes of Ni–Ti hydrotalcite-like precursors, (b) NO conversion of catalysts obtained by different preparation processes, (c) mechanism of NO removal induced by TiO2 on the surface of NiO modified catalysts, and (d) sulfur resistance of Ni–Ti hydrotalcite-like precursors at 240 °C [167]. Reproduced from Ref. [167], Copyright 2019, Elsevier.
Figure 10. (a) Different preparation processes of Ni–Ti hydrotalcite-like precursors, (b) NO conversion of catalysts obtained by different preparation processes, (c) mechanism of NO removal induced by TiO2 on the surface of NiO modified catalysts, and (d) sulfur resistance of Ni–Ti hydrotalcite-like precursors at 240 °C [167]. Reproduced from Ref. [167], Copyright 2019, Elsevier.
Catalysts 09 00771 g010
Figure 11. (a) The SEM image of nanoporous microspheres Mn-Ce-Fe-Ti mixed oxide catalysts, (b) focused ion beam (FIB) images and (c) SEM image of nanoporous microspheres Mn–Ce-Fe-Ti [185], (d) SEM images of MnOx-CeO2-Al2O3 (CP-MHC), and (e) MnOx-CeO2-Al2O3 (SD-MHC) [159]. Reproduced from Ref. [159], Copyright 2018, Elsevier.
Figure 11. (a) The SEM image of nanoporous microspheres Mn-Ce-Fe-Ti mixed oxide catalysts, (b) focused ion beam (FIB) images and (c) SEM image of nanoporous microspheres Mn–Ce-Fe-Ti [185], (d) SEM images of MnOx-CeO2-Al2O3 (CP-MHC), and (e) MnOx-CeO2-Al2O3 (SD-MHC) [159]. Reproduced from Ref. [159], Copyright 2018, Elsevier.
Catalysts 09 00771 g011
Figure 12. (a) The mechanism of selective catalytic CO removal from CeO2-MnOx-Al2O3 surface, (b) NO conversion of Ce-Mn/Al2O3 catalyst [200] (Reproduced from Ref. [200], Copyright 2016, Elsevier.), and (c,d) Mn-Mo-WOx/TiO2-SiO2 catalyst CO-SCR for sulfur and water resistance [201]. (Reproduced from Ref. [201], Copyright 2018, Elsevier.)
Figure 12. (a) The mechanism of selective catalytic CO removal from CeO2-MnOx-Al2O3 surface, (b) NO conversion of Ce-Mn/Al2O3 catalyst [200] (Reproduced from Ref. [200], Copyright 2016, Elsevier.), and (c,d) Mn-Mo-WOx/TiO2-SiO2 catalyst CO-SCR for sulfur and water resistance [201]. (Reproduced from Ref. [201], Copyright 2018, Elsevier.)
Catalysts 09 00771 g012
Figure 13. (a) The mechanism of C3H6 selective catalytic NO removal on Ag/Al2O3 surface, (b) the effect of Ag loading on NO conversion [205], and (c,d) the effect of Fe doping on the sulfur resistance and water resistance of Ag/Al2O3 catalyst [206]. Reproduced from Ref. [206], Copyright 2017, Elsevier.
Figure 13. (a) The mechanism of C3H6 selective catalytic NO removal on Ag/Al2O3 surface, (b) the effect of Ag loading on NO conversion [205], and (c,d) the effect of Fe doping on the sulfur resistance and water resistance of Ag/Al2O3 catalyst [206]. Reproduced from Ref. [206], Copyright 2017, Elsevier.
Catalysts 09 00771 g013
Figure 14. (a) Reaction pathways of NO + H2 and products on Pd (111), (b) potential energy diagram and geometrical structure of H atom-assisted direct dissociation of NO on Pd (111), and (c,d) relative selectivity of N2, NH3, N2O, and H2O at different temperatures [212]. Reproduced from Ref. [212], Copyright 2015, Elsevier.
Figure 14. (a) Reaction pathways of NO + H2 and products on Pd (111), (b) potential energy diagram and geometrical structure of H atom-assisted direct dissociation of NO on Pd (111), and (c,d) relative selectivity of N2, NH3, N2O, and H2O at different temperatures [212]. Reproduced from Ref. [212], Copyright 2015, Elsevier.
Catalysts 09 00771 g014
Figure 15. (a) Reaction pathway and mechanism of nonthermal plasma for NO removal, (b) effect of plasma treatment under different atmospheres, and (c,d) effect of H2O and SO2 on nonthermal plasma removal of NOx [241]. Reproduced from Ref. [241], Copyright 2019, Elsevier.
Figure 15. (a) Reaction pathway and mechanism of nonthermal plasma for NO removal, (b) effect of plasma treatment under different atmospheres, and (c,d) effect of H2O and SO2 on nonthermal plasma removal of NOx [241]. Reproduced from Ref. [241], Copyright 2019, Elsevier.
Catalysts 09 00771 g015
Figure 16. (a) Mechanism of NOx removal by plasma-modified catalyst Mn–Ce/ZSM5, (b) effect of different Mn–Ce ratios and plasma treatment on NO conversion, (c) denitrification process of catalyst Mn1Ce1/ZSM5 conforms to Eley–Rideal mechanism, and (d) study on sulfur and water resistance of plasma-modified catalyst [252]. Reproduced from Ref. [252], Copyright 2018, Elsevier.
Figure 16. (a) Mechanism of NOx removal by plasma-modified catalyst Mn–Ce/ZSM5, (b) effect of different Mn–Ce ratios and plasma treatment on NO conversion, (c) denitrification process of catalyst Mn1Ce1/ZSM5 conforms to Eley–Rideal mechanism, and (d) study on sulfur and water resistance of plasma-modified catalyst [252]. Reproduced from Ref. [252], Copyright 2018, Elsevier.
Catalysts 09 00771 g016
Figure 17. (a) Schematic diagram of plasma-enhanced catalysis and plasma-driven catalysis, (b) mechanism of low-temperature denitrification of nonthermal plasma-enhanced Mn-CoOx catalyst, (c) effect of plasma voltage and treatment time on NO conversion, and (d) study on thermal stability of thermogravimetric curves [255]. Reproduced from Ref. [255], Copyright 2015, Elsevier.
Figure 17. (a) Schematic diagram of plasma-enhanced catalysis and plasma-driven catalysis, (b) mechanism of low-temperature denitrification of nonthermal plasma-enhanced Mn-CoOx catalyst, (c) effect of plasma voltage and treatment time on NO conversion, and (d) study on thermal stability of thermogravimetric curves [255]. Reproduced from Ref. [255], Copyright 2015, Elsevier.
Catalysts 09 00771 g017
Table 1. Effect of different adsorption materials on NO conversion.
Table 1. Effect of different adsorption materials on NO conversion.
Adsorbent MaterialSynthetic MethodReaction ConditionsTemperature (°C)Conversion (%)Ref.
Ag-Cu-BTCpreassembled method500 ppm NO, GHSV 10,000 h−1, N2238100[36]
ZeoliteWetness impregnation250 mg/m NO, 1000 mg m−3 SO2 GHSV 150 L h−1, N225091.7[62]
Gas-phasemicrowave600 ppm NH3, 3% O2, and GHSV 30,000 h−1 N28050[61]
ACFPrecipitation methodFlue gas 0.1–1%, GHSV 1000 mL/min, N24060[58]
ACNitric acid hydrothermal1000 ppm NO, 20 vol.% O2 GHSV 16,000 h−1 N22556.6[60]
Table 2. Effect of different adsorbents on removal of NOx.
Table 2. Effect of different adsorbents on removal of NOx.
Adsorbent MaterialSynthetic MethodReaction ConditionsNO Conversion %Ref.
Cu/SAPO-34Acid–alkali hydrothermal300 ppm NO, 3% O2 GHSV 12,000 h−1, N290[63]
Fe/H-ZSM-5Hydrothermal5000 ppm NO, 5% O2, GHSV = 35,000 h−1, He63.4[64]
Fe-ZSM-5 @CeO2Dopamine polymerization1000 ppm NO, 5% O2, GHSV 33,600 h−1, N290[65]
Cu-BEAPrecipitation1000 ppm NO, 3.5 vol% O2 1000 mL/min, N2100[66]
g-C3N4/WS2Solvent evaporation500 ppm NO, 20 vol% O2, GHSV 400 mL/min, N272.5[67]
Niobium phosphateImpregnation1000 ppm NO, 1125 ppm O2, GHSV 100 mL/min, He100[68]
Pt/SBA-15Thermal hydrolysis4000 ppm NO, 10 vol% O2, GHSV 50 mL/min, Ar100[69]
Semi-cokeHydrothermal1000 ppm NO, GHSV 6000 h−1, N2100[70]
Sintered oreImpregnation400 mg/m3 NO, 15% O2, GHSV = 1000 h−1, Ar61.6[71]
Table 3. Removal of NOx by oxidation.
Table 3. Removal of NOx by oxidation.
Oxidation MethodReagents MaterialsTemperature (°C)NO Conversion (%)Reference
Gas phaseClO216094[84]
O34082[85]
O36095[88]
Liquid phaseUV/NaClO25098.1[92]
NaClO/NaClO25085[90]
NaClO2/Na2S2O812082.7[92]
Catalytic oxidationK-OMS-250-[103]
MnOx25091.4[106]
SmMn2O530090[107]
Fe0.32MnO225080[79]
Fenton14090[125]
La0.8Pr0.2MnO326091[130]
Table 4. Effects of different synthetic methods on de-NOx activity of NH3-SCR.
Table 4. Effects of different synthetic methods on de-NOx activity of NH3-SCR.
CatalystSynthetic MethodGHSV (hr−1)Temperature (°C)Conversion (%)Ref.
Ce-Mo/TiO2Coprecipitation19,000200–40090[177]
MnAl/LDOFNP60,000150–250100[178]
MnOx-CeO2-Al2O3self-propagating synthesis15,38450–400100[179]
Mn–Ce-TiHydrothermal64,000150–40098[180]
MnOx-CeO2-TiO2Sol-ge10,000100–30090[181]
Co/Ni-CeO2Coprecipitation48,00075–20093[182]
Fe-Mn–Ce/γ-Al2O3Sol-ge10,000100–45095[183]
MnOx/CeO2-ZrO2-Al2O3Impregnation10,00050–30090[184]
Table 5. Effects of different synthetic methods on de-NOx activity of CO-SCR.
Table 5. Effects of different synthetic methods on de-NOx activity of CO-SCR.
CatalystSynthetic MethodReaction ConditionsTemperature (°C)NO Conversion (%)Ref.
Cu0.1La0.1Ce0.8OGrind500 ppm NO, 1000 ppm CO, GHSV 26,000 h−1, Ar25099.2[190]
Mn–Ce/TiO2wetness
impregnation
600 ppm NO, 1200 ppm CO, GHSV 40,000 h−1, N220094.9[187]
Pd/CeZrO2Impregnation5 vol.% NO, 10 vol.% CO, GHSV 24,000 h−1, He300100[194]
Pd/LaFeO3Two-step precipitation5% NO, 10% CO, Water 10%, 24,000 mL h−1, He12098[195]
Cu-Ni/LDHPrecipitation1.5 vol.% CO, 0.2 vol.% NO, O2 0.65 vol.%, 12,000 h−1, He400100[196]
Fe0.8Co0.2/ASCPrecipitation500 ppm NO, 1000 ppm CO, GHSV 40,000 h−1, N220096[198]
Table 6. The performance contrast of plasma-modified catalysts.
Table 6. The performance contrast of plasma-modified catalysts.
CatalystConversion (%)Temperature (°C)Ref.
NiO-TiO2-Al2O3Without plasma < 12180–240[244]
Plasma-modified > 70
V2O5/ACFWithout plasma < 3550–150[245]
Plasma-modified > 55
Table 7. Effect of plasma addition on denitrification activity of different catalysts.
Table 7. Effect of plasma addition on denitrification activity of different catalysts.
CatalystConversion (%)Temperature (°C)Ref.
Co-ZSM-5Plasma only 12150–450[256]
Catalyst only 85300
Plasma assist catalyst 95300
V2O5-WO3/TiO2Plasma only 20150–260[257]
Catalyst only 35180–260
Plasma assist catalyst 80160–260
H-mordenitePlasma only 3.8200[258]
Catalyst only 54200
Plasma assist catalyst 91.4200
Mn-Co-CeO2Plasma only 73150[259]
Catalyst only 70150
Plasma assist catalyst 93100
Ba & Cu (C2H2O3)2Plasma only 8525–400[260]
Catalyst only 8225–400
Plasma assist catalyst 95200
Mn–CoOxPlasma only 3.8200[255]
Catalyst only 54200
Plasma assist catalyst 91.4200
Mn-Cu/ZSM5Plasma only 5025–450[262]
Catalyst only 6025–450
Plasma assist catalyst 9025

Share and Cite

MDPI and ACS Style

Liu, Z.; Yu, F.; Ma, C.; Dan, J.; Luo, J.; Dai, B. A Critical Review of Recent Progress and Perspective in Practical Denitration Application. Catalysts 2019, 9, 771. https://doi.org/10.3390/catal9090771

AMA Style

Liu Z, Yu F, Ma C, Dan J, Luo J, Dai B. A Critical Review of Recent Progress and Perspective in Practical Denitration Application. Catalysts. 2019; 9(9):771. https://doi.org/10.3390/catal9090771

Chicago/Turabian Style

Liu, Zhisong, Feng Yu, Cunhua Ma, Jianming Dan, Jian Luo, and Bin Dai. 2019. "A Critical Review of Recent Progress and Perspective in Practical Denitration Application" Catalysts 9, no. 9: 771. https://doi.org/10.3390/catal9090771

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