Skip to main content

Recent trends in vanadium-based SCR catalysts for NOx reduction in industrial applications: stationary sources

Abstract

Vanadium-based catalysts have been used for several decades in ammonia-based selective catalytic reduction (NH3-SCR) processes for reducing NOx emissions from various stationary sources (power plants, chemical plants, incinerators, steel mills, etc.) and mobile sources (large ships, automobiles, etc.). Vanadium-based catalysts containing various vanadium species have a high NOx reduction efficiency at temperatures of 350–400 °C, even if the vanadium species are added in small amounts. However, the strengthening of NOx emission regulations has necessitated the development of catalysts with higher NOx reduction efficiencies. Furthermore, there are several different requirements for the catalysts depending on the target industry and application. In general, the composition of SCR catalyst is determined by the components of the fuel and flue gas for a particular application. It is necessary to optimize the catalyst with regard to the reaction temperature, thermal and chemical durability, shape, and other relevant factors. This review comprehensively analyzes the properties that are required for SCR catalysts in different industries and the development strategies of high-performance and low-temperature vanadium-based catalysts. To analyze the recent research trends, the catalysts employed in power plants, incinerators, as well as cement and steel industries, that emit the highest amount of nitrogen oxides, are presented in detail along with their limitations. The recent developments in catalyst composition, structure, dispersion, and side reaction suppression technology to develop a high-efficiency catalyst are also summarized. As the composition of the vanadium-based catalyst depends mostly on the usage in stationary sources, various promoters and supports that improve the catalyst activity and suppress side reactions, along with the studies on the oxidation state of vanadium, are presented. Furthermore, the research trends related to the nano-dispersion of catalytically active materials using various supports, and controlling the side reactions using the structure of shaped catalysts are summarized. The review concludes with a discussion of the development direction and future prospects for high-efficiency SCR catalysts in different industrial fields.

1 Introduction

Fine particulate matter has been designated as a Class 1 carcinogen by the World Health Organization and is classified into two types according to its emission source. The particulate matter that comes out as a solid from the source is called primary particulate matter. Secondary particulate matter is generated through a chemical reaction in which sulfur oxides and nitrogen oxides, which are representative gaseous pollutants emitted from industrial sites, combine with water vapor, ozone, or ammonia in the atmosphere [1]. Nitrogen oxide (NOx), a representative air pollutant that acts as a major precursor to particulate matter, is emitted at high concentrations and is mainly generated from fixed sources, such as manufacturing, power plants, industrial boiler, gas turbines, production processes, as well as mobile sources, such as ships and automobiles [2]. The NOx itself is not only harmful to the human body but also causes various environmental problems, such as acid rain, global warming, and smog. The NOx is highly mobile as it is readily carried over long distances by the wind, causing worldwide problems [3].

Methods to reduce the NOx content are classified into pre-combustion control, combustion control, and post-combustion control methods [4]. The pre-combustion control method involves reducing the nitrogen content by refining the nitrogen component in the fuel. Combustion control technology reduces the NOx emission by controlling the temperature, oxygen content, and residence time of gas during the combustion process. Combustion control technologies include low NOx burner and flue gas recirculation. Post-combustion control involves the removal of NOx through an additional after-treatment system; selective catalytic reduction (SCR) and selective non-catalytic reduction (SNCR) are the most widely used technologies for the post-combustion control [4].

However, pre-combustion and combustion control do not remove NOx with sufficient efficiency to comply with the increasingly stringent emission regulations. Therefore, post-combustion control technology is widely used in industry [5].

The SNCR and SCR technologies are the most typical processes used for post-combustion control. In SCR process, combustion gas containing NOx and a reducing agent (such as NH3 or urea) passes through a catalytic layer to selectively reduce NOx to nitrogen and water vapor. In a SNCR process, the reagent converts NOx into nitrogen and water vapor through reactions without catalysts. Because a catalyst is not used, a reducing agent is injected at high temperature (900–1100 °C) to achieve the activation energy required for the reaction [6]. Therefore, SCR technology is widely used as a NOx reduction process to comply with the emission regulations. Several types of SCR catalysts are used including, metal oxide-based, zeolite-based, alkaline-earth metal-based, and rare-earth-based catalysts. Among these, the V–W-based catalysts are the most widely used, as they exhibit high NOx removal efficiency of more than 90% at temperatures over 380 °C [2, 7]. Zeolite-based catalysts are also used in some catalyst applications because they have a high specific surface area and a wide operating temperature range, but they have the disadvantage of exhibiting high activity only when pretreatment is performed in a moisture-free condition [7]. The characteristics of representative SCR catalysts are described in Fig. 1 and Table 1.

Table 1  Operating characteristics of different SCR catalysts [2]

The SCR reactions are usually defined as either “standard SCR” or “fast SCR.” In general, NOx is composed of 95% NO and 5% NO2; therefore, the reduction reaction is followed in the standard condition. However, in the case of fast SCR reaction, in which NO, NO2, and NH3 react in the absence of oxygen, the reaction is completed more rapidly due to the rapid oxidation [5]. Sometimes, an oxidizing device such as a diesel oxidation catalyst (DOC) is connected to the front of the catalyst to oxidize NO to NO2. The catalytic activity in the low-temperature region is enhanced through fast SCR, which is faster than the existing standard reaction. The fast SCR reaction generally shows excellent reactivity when NO and NO2 are present in a 1:1 molar ratio [5].

Oxygen presence condition

$${\text{4NH}}_{{3}} + {\text{ 4NO }} + {\text{ O}}_{{2}} \to {\text{4N}}_{{2}} + {\text{ 6H}}_{{2}} {\text{O }}:{\text{ Standard SCR}}$$
(1)
$${\text{4NH}}_{{3}} + {\text{ 2NO}}_{{2}} + {\text{ O}}_{{2}} \to {\text{3N}}_{{2}} + {\text{ 6H}}_{{2}} {\text{O}}$$
(2)

Oxygen absence condition

$${\text{2NO }} + {\text{ 2NO}}_{{2}} + {\text{ 4NH}}_{{3}} \to {\text{4N}}_{{2}} + {\text{ 6H}}_{{2}} {\text{O }}:{\text{ Fast SCR}}$$
(3)
$${\text{4NH}}_{{3}} + {\text{ 6NO}} \to {\text{5N}}_{{2}} + {\text{ 6H}}_{{2}} {\text{O}}$$
(4)

Side reactions include ammonia oxidation at high temperatures, oxidation of SO2 caused by excess vanadium content, and formation of ammonium salts by the reaction of unreacted ammonia and SO3. The formation of ammonium sulfate corrodes the post-treatment facilities, leading to reduced catalytic activation due to the blocked surface of the catalyst [8].

The SCR catalysts are divided into the three types: honeycomb, plate, and corrugated. Industrial V2O5–WO3/TiO2 catalysts (usually containing 0.5–3 wt% V2O5 and 5–10 wt% WO3) exhibit high de-NOx efficiency, and excellent resistance to sulfur (SO2) and H2O. Nevertheless, they also exhibit several disadvantages, such as high and narrow effective temperature range (300–400 °C), and a tendency to oxidize SO2 to SO3. The SO3 reacts with NH3 to form ammonium sulfate or bisulfate, shortening the lifetime of the catalyst [9]. As a result, in major industrial applications such as thermoelectric power plants and vessels with SCR catalyst, frequent shutdowns are necessary due to operational and equipment limitations. Therefore, SCR catalysts with high de-NOx efficiency at low operating temperatures are required for economic benefits and reduced energy consumption [10]. In addition, operating limitations of the SCR system cause a temporary rise in the system temperature. Currently, in the case of small and medium-sized power generation facilities and general boilers such as the heat recovery system generator boiler, most of the horizontal reactors have a relatively high impact due to high temperature. Therefore, there is a need to develop a SCR catalyst with high activity and durability against thermal shock at high temperatures. Thus, it is necessary to develop a wide temperature range SCR catalyst which is capable of operating at the conventional low temperatures (200–250 °C), exhibits enhanced catalytic properties such as high de-NOx efficiency along with low SOx conversion, and has a long lifetime.

Fig. 1 
figure 1

Three major families of SCR catalyst [2]

Fig. 2 
figure 2

SCR configurations with typical system temperatures: a high-dust system, b low-dust system, and c tail-end system [11]

Fig. 3
figure 3

Potential mercury transformations during coal combustion and in the resulting flue gas [16]

Fig. 4
figure 4

Oxidation of SO2 to SO3 in a boiler and SCR (Electric Power Research Institute) [18]. ESP electrostatic precipitator, FGD flue gas desulfurization system, I.D induced-draft, SCR  selective catalytic reduction system.

Fig. 5
figure 5

NOx removal efficiencies of the synthesized SCR catalysts according to vanadium content measured in the temperature range 150–450 °C a NOx conversion and b N2 selectivity and N2O concentration. Reaction conditions: [NOx] = 300 ppm, [NH3] = 300 ppm, [SO2] = 0 or 300 ppm, [O2] = 5%, balance N2, total flow 500 sccm, and the gas hourly space velocity (GHSV) = 60,000 h−1

2 Applications of NH3-SCR catalysts for stationary sources

In stationary sources, various problems, such as the carbon content in fly ash, sulfur, alkali metals, the concentration of catalytic poisons in fuel, carbon monoxide (CO) emission, and corrosion may occur depending on the application [7]. The SCR facilities in stationary sources are generally classified into high dust configurations, low dust configurations, and tail-end systems, and are configured according to the application requirements (see Fig. 2) [11]. The variety of emission conditions from a stationary source require the NH3-SCR catalyst to operate under varied conditions and catalytic properties. Specifically, the catalysts must resist sulfur poisoning and minimize the oxidation of SO2 to SO3 because the SO2 concentration in the flue gas is relatively high [9].

Power plants generally use a high dust system, and, in this configuration, the SCR units are usually located directly after the boiler. Then an electrostatic precipitator and desulfurizer are installed in that order. This system has the advantage of having a high operating temperature without the need for additional heating devices, as the catalysts are installed at the end of the boiler. However, in facilities using biomass fuels, the catalytic activity decreases significantly due to physical poisoning (cell plugging of the monolith catalyst and wear of the plate catalyst) by alkali (earth) metals, and particulate matter in the exhaust gas significantly decreases the catalyst lifetime [12].

In a low dust configuration, the SCR reactor is located behind the hot-side electrostatic precipitator (H-ESP), reducing catalyst degradation by particulate erosion. However, this configuration requires the installation of an expensive H-ESP and a flue gas reheat system to maintain optimum operating temperatures [11].

Finally, in the tail end system, the catalytic system is placed downstream of the electrostatic precipitator and the desulfurization system. This system is mainly used in Western Europe, and because it removes both dust and SO2 from the catalyst, it is suitable for increasing the catalyst lifetime. However, in this system, NOx removal efficiency is drastically decreased because of the low-temperature regime it operates in. Although V-based catalysts are the most efficient in the temperature range of 300–400 °C, their efficiency rapidly decreases at lower temperatures, and the unreacted NH3 reacts with SO2 or H2O to form ammonia bisulfate. Therefore, additional flue gas reheating equipment is required for V-based catalysts used in this system. Reheating the system to the catalytic activation temperature sharply increases the maintenance cost of the SCR system. Therefore, it is necessary to study the potential of using low-temperature catalysts that exhibit activity at 200 °C or less for this configuration due to the problem of reheating the exhaust gas to the catalyst activation temperature [13]. Due to economic and spatial problems, SCR catalysts are currently positioned as tail-end systems. In particular, a temperature below 220 °C is required in incinerators as the SCR is located behind the baghouse filter [7].

Recently, problems related to the storage and transport of ammonia have led to suggestions that it be subject to environmental regulation in the United States, and emissions of ammonia slippage in the unreacted flue gas are stringently monitored in New Jersey and California. Several studies are underway to replace the reducing agent. Although hydrocarbon-reducing agents have attracted considerable attention as replacements for ammonia, their commercialization is difficult due to the requirements of high reaction temperatures and noble metal catalysts [7]. Studies related to “Enhanced SCR” show high NO reduction efficiency in the low-temperature range of ~ 200–350 °C with the addition of NH4NO3 along with NH3, a reducing agent. The resultant reaction is summarized in Eq. (5) [14].

$${\text{2NH}}_{{3}} + {\text{ 2NO }} + {\text{ NH}}_{{4}} {\text{NO}}_{{3}} \to {\text{3N}}_{{2}} + {\text{ 5H}}_{{2}} {\text{O}}$$
(5)

Variables that promote the enhanced SCR reaction at low temperatures include space velocity, reaction temperature, and injection amount of ammonium nitrate. When these parameters are optimized, a high NOx removal efficiency of 90% can be expected at a low temperature of 180 °C even with the use of a commercial V–W/TiO2 catalyst [14].

2.1 Power plants

Power plants mainly use various fossil fuels, such as coal, heavy oil, and liquefied natural gas (LNG) to generate electricity. However, each fuel source poses a particular set of problems in the SCR plant; the deactivation mechanism of SCR catalysts used in coal-fired power plants is different from that of catalysts used in biomass power plants. In coal-fired power plants, catalyst deactivation by mercury and sulfate is the main factor. However, in biomass power generation, alkali (earth) metals act as the main cause of catalyst life reduction [15].

Coal fuel presents several problems. First, it is a representative anthropogenic source of mercury. Among heavy metals in coal, highly volatile mercury is converted into gaseous elemental mercury under high-temperature conditions (> 1400 °C) during combustion. The Hg present in flue gas exists in three main forms: Hg0 (elemental mercury), Hg2+ (oxidized mercury), and HgP (particulate mercury). First, Hg0(g), which is the gaseous elemental mercury present in the flue gas, reacts with other gaseous components and particulate matter and is converted into either HgP or oxidized mercury (Hg2+) [16]. In Fig. 3, HgP is collected in ESP facilities, and mercury oxide is removed with flue gas desulfurization (FGD) because it is soluble in water; however, gaseous elemental mercury is insoluble in water and is therefore difficult to reduce using the existing facilities.

Fig. 6
figure 6

Proposed mechanism of NO reduction and N2O formation, as well as H2O/SO2 suppression effects, with the participation of a Lewis acid sites and b Brønsted acid sites over the Mn/Ti–Si catalyst [43]

To solve this problem, an SCR catalyst has been developed with elemental mercury oxidation capability, along with a separate oxidation catalyst. Although SCR technology is mainly used for NOx removal, many studies have shown that it can also be used to convert Hg0(g) into Hg2+. The oxidation of elemental mercury by the SCR catalyst is closely related to the chlorine content in the coal. As the chlorine content of coal increases, the oxidation of Hg0 to Hg2+ is promoted by the SCR catalyst, and as a result, the Hg2+ concentration at the end of the SCR process increases [16, 17].

Second, a problem is caused by the sulfur present in coal and heavy oil. During combustion, pyrite and organically-bound sulfur are mostly oxidized to SO2, and a very small amount is converted to SO3 (SO2/SO3 ratio is generally ~ 40:1–80:1). Meanwhile, a portion (0.3–2%) of SO2 is oxidized to SO3 in the passage of the SCR facility (see Fig. 4) [18]. The SO3 in fuel forms various salts, such as H2SO4, NH4HSO4, and (NH4)2SO4 by reacting with H2O or NH3 in the exhaust gas, poisoning the catalyst and corroding the post-treatment facilities such as the air pre-heater and ESP [11].

The co-firing and burning of biomass fuel causes catalyst poisoning due to impurities in the fuel. The analysis of Danish cereal straw and wood chips, which are the most widely-used biomass fuels for power plants (Table 2), indicated the presence of 0.2–1.9 wt%. potassium, which is an alkali metal. Submicron aerosol particles in flue gases from straw combustion consist of almost pure potassium chloride and sulfate with small amounts of sodium, phosphorus, and calcium [19, 20].

Table2 Fuel data for Danish cereal straw and wood chips [20]

Currently, as many countries increase the use of biomass for environmental reasons, the effect on the deactivation of the catalyst by the alkali metals K and Na and the alkaline earth metals Ca and Mg contained in biomass is attracting attention. The P2O5 reduces the pore volume of the catalyst surface and also reduces the catalytic activity. In addition, it reacts with H2O or O2 in the exhaust gas to form H3PO4. With an increase in the P2O5 concentration, the H3PO4 forms poly-, pyro-, tri-, and metaphosphoric acid chains, negatively affecting the catalyst activity [21]. Potassium (K) in fly ash generated during biomass combustion forms KCl and K2SO4 compounds, which are sticky particles with low melting points, and blocks the micropores of the catalyst to reduce its catalytic activity. Zheng et al. reported the need for additional research on catalyst poisoning caused by small amounts of phosphorus (P), calcium (Ca), and sodium (Na) [22]. In the research on catalyst deactivation by the formation of K compounds (e.g., KCl, K2SO4, and K3PO4) following the combustion of K-getter fuel, the polyphosphoric acid poisoning was relatively low compared to K poisoning. However, pore blocking and fouling occurred on the surface of the catalyst, which had a greater effect on deactivation than poisoning by K [22,23,24]. According to a study on the effect of the addition of inorganic substances to the SCR catalyst and poisoning, catalysts doped with alkali (earth) metals exhibit catalyst deactivation due to a reduction in the NH3 storage capacity, and a strong poisoning effect occurs in the order of K > Na > Ca > Mg [25].

Lastly, LNG-fired power generation is an eco-friendly power source to replace coal-fired power in Asia, and many new power plants are being built. As LNG is composed of ~ 72%–95% methane, ~ 3–13% ethane, ~ 1–4% propane, and ~ 1–18% nitrogen, it does not contain toxic substances such as sulfur and alkali (earth) metals, and there is almost no reduction in the catalyst life. However, as the main components are hydrocarbons, the process emits a large number of harmful substances, including carbon monoxide (CO) and unburned hydrocarbons (UHC).

As described above, various factors affect catalyst poisoning in power plants depending on the fuel used. Research on catalysts having anti-poisoning properties is an active field of study. In addition, a large amount of NOx is generated due to incomplete combustion during power plant startup, as the exhaust gas temperature does not reach the catalyst activation temperature and is discharged. Therefore, to control this problem, research on technology for storing NOx generated below the catalyst activation temperature by installing a NOx trap in front of the catalyst is also being conducted [26].

2.2 Incinerators

Incinerators have a high dust configuration, and the SCR catalyst is placed at the rear end of the boiler; however, catalytic efficiency is severely reduced due to physical and chemical deactivation by fly ash present in the flue gas. Catalysts used in municipal waste incineration (MSWI) plants have reduced activity due to decreases in their specific surface area and pore sizes. One of the main causes is the change of the surface acid sites by alkali metals such as Na and K present in the flue gas [27, 28]. According to the findings of Jan et al., when domestic waste and sludge are incinerated simultaneously, the flue gas contains metals such as Ca, Si, Cl, S, K, Na, Pb, Zn, and P [27]. In particular, during co-incineration, a large amount of P is present in domestic sludge, and therefore, a larger amount of P is emitted than that formed during single combustion [29]. On the one hand, Castellino found that when the H3PO4 concentration reached 1000 ppm, the catalyst redox performance deteriorated, and the catalyst was deactivated after 24 h due to a decrease in the number of active vanadium species [23]. On the other hand, Cao et al. suggested that as P inhibits NH3 oxidation at temperatures above 300 °C, it prevents the formation of N2O and NOx, which are side reactants, and improves catalyst efficiency. To solve the problem of catalyst deactivation, the tail-end system is applied, and ESP and FGD are installed in front of the SCR to remove particulate matter, preventing catalyst poisoning. However, as the flue gas temperature rapidly decreases to approximately 160 °C, it is necessary to develop a low-temperature catalyst [30]. In addition, research on SCR catalysts for incinerators is mainly focused on the catalyst deactivation effect and reaction modeling for each element, along with studies on inhibiting the physicochemical poisoning of catalysts by fly ash and catalyst regeneration methods [27, 31].

2.3 Cement, iron, and steel industry

The cement industry mainly uses the SNCR process to reduce NOx, but as air pollutant emission regulations are tightened around the world, hybrid SNCR–SCR and SCR technologies have attracted increased attention from the cement industry [32]. [32] While the amount of dust in the flue gas of a power plant is approximately 25 g/Nm3, the dust concentrations in the precalciner kiln system of a cement plant range from 50 to 100 g/Nm3. Therefore, the high dust configuration or semi dust application is mainly applied to the post-treatment facility. The semi dust application requires the development of a low-temperature catalyst because the temperature of the SCR facility decreases as dust is collected using a cyclone or ESP equipment in front of the SCR unit [33]. Therefore, there is an unfulfilled requirement for the development of catalysts for cement plants. These catalysts should be resistant to physicochemical poisoning by particulate matter such as fly ash, and exhibit high activity at low temperatures [34, 35].

The iron and steel process produces pollutants such as waste gas, water, and slag, the most serious of which is waste gas. Steelworks emit NOx from processes such as sintering, coking, and rolling, and a particularly large number of pollutants are emitted from the sintering process [36]. In the sintering process, limestone is mixed with iron ore in powder form and heated to process it into a sinter in the form of a homogeneous mass [37]. Steelworks generally use a high-dust configuration, and like the previous stationary sources, the demand for efficient low-temperature catalysts that can function below 240 °C is increasing. As SO2 is present in the exhaust gas, the poisoning of SCR catalysts by sulfur is a major problem [38]. In addition, even if a low-temperature catalyst is developed, the problem of catalyst poisoning due to moisture and sulfur remains. Therefore, research on sulfur- and moisture-resistant catalysts through the addition of a co-catalyst or by changing the catalyst support is an active field of investigation [39, 40]. According to Chu and Wang, there is considerable ongoing research related to the selection of various complex adsorbents to improve the efficiency of denitrification and desulfurization through the simultaneous removal of SO2 and NOx [41].

3 Vanadium-based SCR catalysts

Vanadium, which is most widely used as a commercial catalyst, is considered a representative catalyst because it has high activity at ~ 350–400 °C even at ~ 1–2 wt% content of various vanadium chemical species such as monomeric, polymeric, and crystalline V2O5 [42]. Recently, the tail-end system has become more widely used, and the demand for low-temperature catalysts is increasing. However, active materials that can increase activity in the low-temperature region (< 200 °C), such as Mn and Ce, form salts such as MnSO4 or CeSO4 with the SO2 present in the exhaust gas. This results in the poisoning of the active catalyst sites, and the problem of activity deterioration continues to be an issue. Therefore, the catalytic activity temperature range is extended by increasing the vanadium content, making vanadium the major component in low-temperature catalysts. However, vanadium catalysts still have some problems. The operating temperature range is relatively narrow at ~ 350–400 °C, and at high temperatures, ammonia oxidation, SO2 oxidation, and N2O formation occur due to side reactions [5]. Because vanadium exhibits several problems, such as being easily sublimated and generating biological toxicity when used for a long time, it is necessary to minimize the vanadium content.

An analysis of NOx conversion as a function of the vanadium content of a catalyst showed that when the vanadium content was between 0.5 and 1 wt%, the conversion efficiency was 19.57 and 51% at 200 °C, respectively. As the vanadium content increased from 2 to 5 wt%, it exhibited increased efficiency and an extended operating temperature range from 200 to 350 °C. When the temperature was greater than 400 °C, the efficiency decreased rapidly with vanadium content and was closely related to the amount of N2O generated. Figure 5 shows that as the vanadium content increases, the amount of N2O, generated due to one of the side reactions of the SCR reaction, increases rapidly, and N2 selectivity decreases. The main SCR side reactions by vanadium catalysts include N2O generation by NH3 oxidation and SO3 generation by SO2 oxidation. Extensive studies are required on vanadium content conditions and dispersibility improvements that inhibit these side reactions and yield high efficiency. Figure 7a shows the NOx conversion efficacy according to the presence of SO2 (off: solid line, on: dotted line); the condition without SO2 at low temperature (below 200 °C) shows high efficiency across the entire temperature range. There is no significant difference in the range of ~ 250–350 °C; however, the efficiency is increased when SO2 is added at a high temperature.

Fig. 7
figure 7

NOx removal efficiencies of the synthesized SCR catalysts according to tungsten content measured in the temperature range 150–450 °C a NOx conversion and b N2 selectivity and N2O concentration. Reaction conditions: [NOx] = 300 ppm, [NH3] = 300 ppm, [SO2] = 0 or 300 ppm, [O2] = 5%, balance N2, total flow 500 sccm, and the gas hourly space velocity (GHSV) = 60,000 h−1

When SO2 is present at a low temperature, it is more readily adsorbed to the catalyst acid site than ammonia. It blocks the acid sites, which may cause a decrease in activity at low temperatures (Fig. 6) [43]. In addition, in the presence of SO2, the amount of N2O generated may be reduced as the oxidation reaction of NH3 is suppressed. This is because SO2 is first adsorbed to the Brønsted acid site and Lewis acid site, and then the adsorption of NO is inhibited, leading to the preferential adsorption of NH3. The adsorbed NH3 improves the SCR reaction and suppresses side reactions.

Fig. 8
figure 8

Structures of the dehydrated surface vanadate phases on TiO2:a isolated mono-oxo VO4, b oligomeric mono-oxo VO4, and c crystalline V2O5 nanoparticles on top of the surface vanadate monolayer [78]

As a commercial catalyst, tungsten or molybdenum are mainly used as co-catalysts for vanadium-based catalysts. Tungsten and molybdenum play a role in maintaining the structural and thermal stability of the catalyst and have the advantage of resistance to sulfur. Although the choice of tungsten or molybdenum depends on the intended use, molybdenum has the limitation of generating N2O at high temperatures.

Analyzing the catalyst properties according to tungsten content (Fig. 7) at a constant vanadium content shows that as the tungsten content increases, the NOx removal efficiency is improved at low temperatures. When 10 wt% of tungsten is added, the catalysts show the highest efficiency across the entire temperature range, but this decreases sharply at 13 wt% tungsten. The amount of N2O emitted showed almost no change according to the tungsten content. The value of the N2 selectivity calculated based on the amount of N2O generated showed an almost invariant range within ~ 80–87% at 450 °C. As the content of the co-catalyst WO3 was increased, the activity was improved at low temperatures. However, as it increases to 13 wt%, the dispersibility is lowered, and catalytic activity is reduced due to aggregation and crystallization. Therefore, to improve this, it is necessary to improve the dispersion properties of the main/co-catalyst.

Fig. 9
figure 9

Structures of supported a monomeric and b oligomeric VOx units. c Crystal structure of orthorhombic phase V2O5 (Pmnm, No.59). d Raman spectra of bulk V2O5 (V2O5-ox) and TiO2-supported V2O5 catalysts (1, 3, 5 and 9 wt % V2O5/TiO2) [76]

3.1 Addition of promoter and support for V-based catalyst

Table 3 lists the various compositions of vanadium-based SCR catalysts, including the main catalyst and support contents. The activity temperature window is defined as the region where more than 90% of NOx removal activity occurs for each synthesized catalyst, and specific results are summarized. Various components such as W, Mo, Ba, Ce, Mn, and Sb are added as co-catalysts to the vanadium-based SCR catalyst [44,45,46,47,48,49,50,51,52]. The W and Mo are the most widely used commercial co-catalysts and demonstrate thermal and structural stability and resistance to sulfur. In particular, W extends the working temperature of vanadium-based catalysts from low to high-temperature ranges [53], and although molybdenum has arsenic poisoning resistance, it exhibits a drawback of generating N2O at high temperatures. However, as low-temperature catalysts have recently been in the spotlight, the widespread use of Mo is increasing as a co-catalyst [54]. This interest has also led to many studies related to Mn and Ce, particularly related to the effect of Mn on NOx removal activity according to its oxidation states of + 1 to + 4. The Mn shows high NOx removal characteristics even at low temperatures (~ 150–200 °C). However, in the presence of SO2, MnSO4 is formed, and catalyst acid sites are poisoned [55]. Furthermore, Ce is widely used as a co-catalyst because of its high oxygen storage capacity and easy oxygen storage and release due to the Ce4+ ↔ Ce3+ redox shift [56,57,58]. Based on the oxygen vacancies present on the surface, active oxygen chemically adsorbed on the surface moves rapidly in the Ce-based catalyst and promotes the fast SCR reaction [5]. The Zr expands the catalyst operating temperature range by inhibiting particle aggregation and improving acid site dispersibility and has high thermal stability and SO2 resistance; however, its excessive addition promotes ammonia oxidation [59, 60]. When Sb and Nd are added in small amounts, they prevent catalyst poisoning by SO2 and water, and promote the decomposition of ammonium bisulfate (ABS) [61,62,63,64]. As a NOx adsorbent, Ba is used as a co-catalyst to adsorb NOx at low temperatures [26]. In addition, many researches have been conducted in which many transition metals and rare earth metals are used as co-catalysts for SCR catalysts [65,66,67,68,69,70,71].

Table 3 Comparison of different types of vanadium-based SCR catalysts

3.2 Effect of Vanadium oxide states

As a representative transition metal, vanadium contains various valence states, such as + 2, + 3, + 4, and + 5, in the form of VO2, V2O3, and V2O5 [72]. Vanadium is mainly present in the form of + 4 or + 5 oxide states, but the rate at which side reactions such as SO2 oxidation and N2O generation occur and the low-temperature catalytic efficiency change, depend on the ratio of oxidation states present [73]. Therefore, studies on pH control in an aqueous vanadium precursor solution, changing the calcination temperature and maintenance time, and using various precursors are of great interest [54, 74, 75]. Previous studies have shown that the vanadium-oxalate complex undergoes thermal decomposition to V2O5 in the presence of + 4 and + 5 compounds at approximately 270 °C. Therefore, to control this, the amount of V4+ was changed by changing the calcination time under a temperature of 270 °C [76]. Inomata et al. found that the shorter the calcination time (at 270 °C), the higher the fraction of + 4 oxides, yielding an amorphous catalyst with a greenish color. As the calcination time increases, most vanadium exists in the + 5 form, resulting in a highly crystalline yellow catalyst. In addition, it was suggested that the NH3-SCR reaction rate was faster based on the high redox cycle and Lewis acid site of bulk vanadium oxide than that of V2O5 dispersed in the support. Youn et al. explored a V2O3 catalyst containing 5 wt% V, and reported that it had a high SCR efficiency across a wide temperature range and formed the least N2O when it existed in the + 3 oxidation state rather than in the + 4 or + 5 oxidation states, and therefore, the N2 selectivity was high. Additionally, when defective V2O5−x was formed on the surface, it had a lower NH3-SCR reaction energy barrier than that of crystalline V2O5, and therefore this study showed that high activity vanadium catalysts could exist even at low temperatures [77].

3.3 Effect of Vanadium surface density

Inomata et al. reported that VOx exists as monomeric and oligomeric VOx units on the surface of the support, and crystallized vanadium shows an orthorhombic phase V2O5 crystal form [76]. At a low surface concentration (< 2 V atoms/nm2), vanadium exists in the form of monomeric vanadyl without V–O–V bonds. As the vanadium content increases (2–8 V atoms/nm2), it adopts a V–O–V bridge and exists in the form of oligomeric vanadyl. At high surface densities greater than 8 V atoms/nm2, it exists in the form of crystallized V2O5 nanoparticles (Fig. 8) [78].

Fig. 10
figure 10

Type of monolith SCR catalysts: a honeycomb monolith, b plate, and c corrugated type catalyst

The binding forms of vanadium can be confirmed through Raman spectroscopy (Fig. 9), and the binding form changes depending on the vanadium content and the resulting surface density. Therefore, the catalytic efficiency also changes [76].

Fig. 11
figure 11

Catalytic oxidation process of SO2-SO3 under the action of V2O5

He et al. suggested that when the vanadium content is less than 1.3 wt%, 80% or more of the vanadium is monomeric [79]. When the vanadium content is 1.3–3 wt%, the monomeric fraction decreases rapidly, and the polymeric fraction increases to 33%. According to their results, polymeric vanadyl species show higher NH3-SCR efficiency compared to that of monomeric species owing to the shorter reaction pathway for regeneration of catalytic oxidation–reduction reaction and lower reaction energy barrier for the catalytic reaction cycle [79].

4 Dispersion of active catalytic materials

Because catalysis is a surface reaction, it must be supported on a porous support. The choice of support is determined by the need for high dispersibility to increase catalyst efficiency and the need for it to have high thermal stability, specific surface area, and suitable mechanical properties for the intended application [80, 81]. Therefore, many researchers have used various dispersants to improve dispersibility [82,83,84,85,86]. It is important to finely disperse catalytically active materials without aggregation to enhance the catalyst acid sites, achieve a high specific surface area, and prevent sintering at high temperatures [87,88,89]. In addition, significant research has been conducted for improving catalyst dispersibility using supports with abundant functional groups [90] and having high efficiency and durability by dispersing single-atom catalysts on a support [91].

4.1 Nanocomposite selection

Nanoparticles are preferentially reduced precursor materials to form small particle clusters, which can aggregate to form stable particles. Therefore, Ye et al. reported nano-dispersed catalyst materials on the surface by synthesizing the support at the particle formation stage, inhibiting the formation of large size of these particles, and thus exhibiting high NOx reduction efficiency with a smaller catalyst content [92, 93]. Using commercial reduced graphene oxide (rGO) as a support, an MnCe/rGO composite was prepared in which Mn and Ce were nano-dispersed. The catalytic efficiency was increased owing to a high specific surface area without aggregation of the active material. It was sufficiently moldable to be synthesized even with a 1-inch SCR catalyst [94]. In addition, to retain both the advantages of the high specific surface area of rGO and the abundant oxygen functional groups of GO, a catalyst using a GO-r support subjected to thermal reduction after supporting an active material on the GO surface was synthesized [8]. Next, surface-treated graphene was synthesized using N-doped graphene with nitrogen functional groups, which have higher thermal stability than oxygen groups [95]. The N-rGO support exhibited the highest dispersion and particle aggregation inhibition properties. The N-rGO has an appropriate amount of oxygen and N-functional groups on the surface, and thus demonstrates excellent thermal stability at high temperatures and efficient NOx removal characteristics at low temperatures. Surface oxygen functional groups on graphene act as anchoring sites which play an important role in preventing the aggregation of catalytically active materials and dispersing nano-sized particles as the support.

In addition to graphene, studies related to a catalyst containing vanadium and tungsten active materials dispersed using an oxygenated carbon nanotube (O-CNT) support on which an oxygen group is formed through surface acid treatment and having dispersibility and thermal stability of a TiO2 support with nitrogen functional group were conducted [88, 96]. In addition, a study was conducted using hexagonal boron nitride, which has a high melting point of 3000 °C and excellent thermal stability, as a support synthesized h-BN with a porous structure on the surface due to catalytic etching by the transition metal. When the active material is dispersed in this porous structure, the phase is thermally stabilized with high dispersibility, and phase changes and agglomeration are suppressed [97]. As a result, it was possible to nano-disperse the catalytically active material down to 10 nm by utilizing the surface defect anchoring sites.

4.2 Effect of surface modification on catalytic activity (structure and morphology)

Catalytic performance is determined by the composition of active catalytic materials, variations of which could improve catalytic efficiencies. However, studies to minimize the NH3 slip and to develop the SCO catalyst are required owing to the current lack of researches. Therefore, it is difficult to utilize more than a specific amount of active catalytic materials, and it is naturally limited in the linear efficiencies line. To enhance the catalytic performance despite relatively lower contents of active materials, various studies have been conducted concerning surface modification such as functionalization and structure deformation. A phase transition in catalyst structures, such as hollow and kegging structures [98], is also one of the considerations. In the previous literature, active materials were easily deposited on modified and/or functionalized support, nanotubes, APT(Hydroxyapatite), controlled pores/defects, and exchanged sulfated species [99]. Deliberate surface sulfation can increase the active oxygen sites leading to the enhancement of NH3-chemisorption [100, 101]. Zhang et al. [102] reported on the influence of sulfation in a heterogeneous system with an iron-based catalyst. The enhancement of acid sites (Brønsted acidity and Lewis acid strength) and NH3-adsorption show positive effects, not only increasing efficiencies at the relatively higher temperature but also improving the resistance to SO2. In addition, the tolerance to H2O and SO2 is improved because SO4 species formed by sulfation act as acid sites to increase the amount of adsorbed NH3. Similarly, the catalysis surface shows the mentioned effects due to modification by acidification using acids such as HCl, HNO3, H3PO4, and H2SO4 [103, 104]. Chenglong et al. [103] reported an improvement in the catalytic activity, NOx conversion, and N2 selectivity by acidification in the following order: H2SO4 > H3PO4 > HNO3 > HCl. In particular, surface acidity was significantly improved while increasing the Brønsted acid sites (M–O–NH4+), and it naturally yielded reducibility and surface chemical adsorption. Other studies have been conducted using H4PO7 to synthesize a nano-hollow structure, increasing surface oxygen storage capacity and acid sites [104]. In other work [105], the structure, morphology, and size of pores influenced the amount of NH4+ adsorbed for high N2 selectivity. In addition, a study was conducted on the role of pore diffusion in determining the active site in NH3-SCR reaction, and enhanced activity was investigated by controlling the more significant textile properties such as specific surface area, pore size, and structure. This work also explained that particles form the active sites, along with the correlation between the pore diffusion and actual activation energies [106]. The pore size in TiO2 results in a significant impact on active catalytic species [107], and the influence of meso- and micro-pores was also studied.

Consequently, it has been reported that the catalytic activity and physicochemical properties change while the vanadium bond is altered by pore size and structure effect, and the suitability of mesopores for catalytic properties is suggested. Hence, research on modification using silicon (Si) was conducted to enhance H2O resistance and improve catalytic activity by controlling pore textile properties [108]. The pore size and structure positively affect the dispersibility and acidity of catalytically active materials such as vanadium and tungsten caused by inhibiting and decreasing the specific surface area and improving the hydrothermal stability. The larger volume of mesopores can have more surface oxygenation groups, naturally leading to improved catalytic performance. Accordingly, research is being conducted to control pores through the use of porous materials based on carbon (C) [109, 110]. Various researchers [111] have studied carbon-based materials, and the studies have mainly focused on enhancing catalytic performance through chemical surface modification. However, the occurrence of N2O is also promoted, leading to the possibility for re-oxidation to NOx [112].

Activated carbon is used to modify the chemical surfaces to enhance catalytic performance. Incorporating nitrogen species into carbon enhances the catalytic activity and promotes NO2 formation by facilitating chemisorption of surface nitrogen species, which improves catalytic performance [113,114,115,116,117,118,119]. The de-NOX efficiency behavior of catalysts with carbon is expected to increase due to the surface porosity. Most studies of surface functionalization have focused on adding oxygen or nitrogen via chemical or physical methods. Nitrogen functionalities can act as adsorption sites for NOX, and the methods are viscose-based activated carbon fibers (VACF), oxygen plasma, and nitric acid modification [120]. Among them, oxygen plasma treatment can improve and increase the oxygen functional groups on the surface, and it seems to improve pore distribution, dispersion of active catalytic material, and catalytic properties. Doping is widely used as a modification to improve catalytic activity. The S-doping of the catalyst improved its redox property, which was beneficial to the catalytic property. Zhang et al. [121] reported that sulfur could lead to more NH3 adsorption species through S-doping and increase the oxygen vacancy, leading to excellent activity. Royer et al. [122] reported the importance of structural and textural properties. This research explained that crystal morphologies and structures show significant differences in surface area and pore volume. Particles were highly dispersed on a hollow sphere- and monoclinic-structure among diverse types of supports including hollow sphere, star, rod, mesoporous, and crystalline- structures. Additionally, control of the TiO2 crystalline phase uses a cationic surfactant cetyltrimethyl ammonium bromide (CATB). In contrast, the structure and valence state of the active phase is controlled by changing the calcination temperature [123]. In conclusion, anatase TiO2 crystalline is more conducive for improving electron transfer.

5 Commercial SCR catalyst monolith forms

As shown in Fig. 10, SCR catalysts are divided into three types [124]: honeycomb monolith, plate, and corrugated. The honeycomb type is an extruded ceramic structure that is easily regenerated, but it takes a long time for manufacture, and the catalyst is too heavy. The plate type is a metallic substrate. It has high thermal and mechanical durability, making it suitable for fly ash and gases (e.g., erosion, pressure drop) but with a low specific surface area. The corrugated type is based on a glass-fiber substrate; it has a large specific surface area and short manufacture period but poor durability. Catalysts normally use titanium oxides TiO2, alumina oxidesAl2O3, zeolite, and carbon as support. In general, TiO2 is used as a commercial catalytic support, and zeolite is used according to the condition of exhaust gases in the mobile source.

Fig. 12
figure 12

Correlations between catalyst wall thickness and channel weight and SO2-SO3 conversion [124]

5.1 Reducing SO2–SO3 conversion and NH3 slip (oxidation NH3)

The main fuels for coal-fired electrical power plants are bituminous coal or biomass mixing coal. Therefore, reagents to sodium-based absorbents, including calcium- and magnesium-based, are used for sulfur removal. Additionally, ammonia is used as the reducing agent in SCR, but it is oxidized at high temperatures (> 380 °C). Ammonia slip in stationary sources, i.e., coal-fired and chemical power plants, causes equipment corrosion and negative side reactions. Therefore, the SCR system of stationary sources is adjusted to an NH3/NOX ratio to alleviate ammonia slip. However, it leads to a serious problem of reduced denitration efficiency. Therefore, the commercial V–W(Mo)/Ti catalyst is used considering SO2–SO3 conversion. The SO3 is generated from an SCR system in a coal-fired boiler and wet electrostatic precipitator, and it reacts with ammonia, corroding the SCR system and deactivating the catalyst. Equation (2) shows that the SCR catalyst converts SO2 in coal-fired sulfur to SO3, thereby, damaging equipment when it generates sulfuric acid on reaction with water vapor present in combustion gas and forms gypsum from the CaO in fly-ash materials. Ammonium sulfate (AS) and ammonium bisulfate (ABS) are generated by the reaction of SO3 with excess ammonia, i.e., slipped-NH3.

The AS is deposited in catalyst pores and plant equipment, causing problems concerning corrosion, clogging, and performance degradation. The engineers in catalyst application and manufacturers generally require a low SO2–SO3 oxidation rate (less than 2%) [125,126,127,128,129,130,131]

$${\text{SO}}_{{2}} {-}{\text{SO}}_{{3}} {\text{conversion }}\left( \% \right) \, = \, \left( {{\text{SO}}_{{3}} {\text{outlet}}/{\text{SO}}_{{2}} {\text{inlet}}} \right) \, \times { 1}00\%$$
(6)

Vanadium provides the active site of the SCR reaction, but it also causes oxidation of SO2 to SO3. Therefore, many researchers have tried to control the SO2–SO3 conversion, such as the use of low sulfur coal and absorbents (magnesium, calcium), as well as improving the desulfurization system. In addition, attempts have been made to suppress this reaction by changing the composition of the commercial SCR catalysts and modifying the surface.

The SCR reaction occurs on the surface, whereas SO2–SO3 conversion is a diffusion-based reaction (see Fig. 11). The variables that affect SO2–SO3 conversion are catalyst composition, temperature [132], catalyst geometry (pitch, open area, wall thickness) (see Fig. 12) [124, 133,134,135,136], gas composition, and operating conditions (see Fig. 13). Therefore, improved durability can be obtained through the inhibition of oxidation on the surface by preventing SO2-adsorption. The results indicated a linear dependence on catalyst wall thickness and channel [124, 128, 135]. For this reason, corrugated type catalysts show lower SO2–SO3 conversion and are better in this respect than honeycomb and plate-type catalysts. The SO2–SO3 conversion rate indicated a difference with respect to the velocity of the gas passing through the catalyst, and the result indicated easier diffusion, as mentioned above, yielding lower AV [134].

However, as the rationale for catalyst choice cannot be limited to a simple SO2–SO3 conversion property, the optimal catalyst monolith is determined by the exhaust gas conditions for each site. Consequently, further research is required for reducing the SO2–SO3 conversion rate.

Studies on the SO2–SO3 conversion by Saltsburg et al. [137] and Morikawa et al. [138] has been in progress since 1979. Scheffknecht et al. [126] and Beretta et al. [139] studied the influence of each factor mentioned above. Scheffknecht et al. [126] reported the effects of SO2–SO3 conversion on catalyst components (Cu, V, W), the concentration of SO2 and H2O, and flue gas velocity. The SO2–SO3 conversion increased with increasing SO2 concentration [140]. However, a non-linear trend was observed from the saturation point of SO2–SO3 conversion [140]. The SO2–SO3 conversion can also be explained by the Arrhenius and Eyring equations. These equations show that the conversion rate increases with increasing temperature. However, the H2O range is not dependent on the SO2–SO3 conversion reaction. This finding was contradicted by other reports, which have found that the addition of H2O influenced the generation of SO3 by inhibiting SO3 adsorption [128, 129]. Along with influenced factors in the SCR process of SO2–SO3 conversion, Yang et al. [133] evaluated and reported various compositions in flue gases, such as O2, NH3, NOx, SOx, H2O, and CO2 [128, 129], as well as catalyst components TiO2, V2O5, WO3, Al2O3, BaO, and SiO2. According to their report, the gas composition is directly influenced by SO2–SO3 conversion. The NH3 especially affects the SO2–SO3 conversion rate, and when the concentration of NH3 exceeds the NO content, it continues to decrease the SO2/SO3 conversion rate. Wang et al. explained that the presence of NH3 inhibits the formation of SO3, while the oxidation of SO2–SO3 conversion can be controlled by readily reacting surface-oxygen with NH3 [141]. The concentration of O2 in the exhaust gas is not impacted by SO2–SO3 conversion [130]. Furthermore, vanadium content directly influences SO2–SO3 conversion, and vanadium surface coverage and density are related to SO3 oxidation [130, 142]. That is to say, SO2–SO3 conversion is connected to only one surface vanadate site. Commercial catalysts use the V–W/Ti components, i.e., a vanadium-based catalyst with a co-catalyst to improve the SO2–SO3 conversion. Thus, many studies have been conducted on SO2–SO3 conversion, and the main factors included the suppression of SO3 through the use of a co-catalyst and optimized support. In general, vanadium-based catalysts are commercially represented using tungsten oxide (WO3) and molybdenum oxide (MoO3) as promoter. Many studies have shown that WO3 inhibits SO2–SO3 oxidation, including the report by Ismagilov et al. [143], and the rate is affected by the tungsten content [93, 143]. However, based on other reports, a consensus has not yet been reached on the influence of tungsten content on the SO2–SO3 conversion rate [133, 142]. Although molybdenum has the desirable property of being resistant to SO2 by naturally restricting the SO2–SO3 conversion rate, other catalysts have been used more widely in the industry than molybdenum [144]. However even with this composition, because the SO2–SO3 conversion rate is limited, studies are being conducted on the effects of metallic bonds and diversification of catalytic composition by incorporating rare earth and transition metals (Cu, Ba, Mg, Fe, Ge, Zn, Ta, and Y). Wachs et al. [130, 131] studied the turnover frequency of SO2 oxidation by co-catalysts on TiO2 supports. The oxidation rate was observed in the order of: V2O5 > Fe2O3 > Re2O7 > CrO3 > Nb2O5 > MoO3 > WO3 > K2O. The effect of iron (Fe2O3) on SO2–SO conversion has been researched by Wang et al. [145]. With an increase in the Fe2O3 loading, the SO2–SO3 conversion also increased. However, studies have reported that excessive iron loading causes a non-uniform distribution of active components on the catalytic surface and an increased SO2–SO3 conversion [145]. Also, to those mentioned in other reports, the addition of barium (Ba) as a co-catalyst showed low SO2–SO3 conversion [143]. The Ba species are an important factor in the SO2–SO3 conversion rate, affecting surface acidity. Ceria (Ce) has sulfur resistance and inhibits SO2–SO3 oxidation; therefore, it is generally used for Mn-based catalysts vulnerable to sulfur [146]. The rare earth niobium (Nb) showed the greatest resistance to SO2 and inhibited surface reaction to SO2. Therefore, using Nb, a decrease in SO2 oxidation at temperatures below 350 °C was demonstrated [93, 130, 147]. Germanium (Ge) and Zinc (Zn) components are effective promoters for retarding the SO2–SO3 conversion [142]. In addition, research on Cu-containing catalysts has shown that they have the disadvantage of increasing the SO2–SO3 conversion; however, they have the advantage of improving the mercury oxidation rate [122]. Xiang et al. [129] have studied the SO3 generation, and Yang et al. [133] focused on the factors of SO2–SO3 conversion. Furthermore, Hitachi-Zosen reported that SO3 reacted with NH3 generated from ABS and AS, as shown in Fig. 14.

Fig. 13
figure 13

Effect of NH3/NOx ratio and AV to SO2-SO3 conversion [134]

The NH3 oxidation generates NO and N2O, and therefore, it must be suppressed. The NH3 slip is generally adjusted to 2 ppm or less [148]. The basic method for inhibiting NH3 slip is to reduce the NH3/NO ratio. However, in this case, it is necessarily directly related to the decrease in the de-NOx efficiency, and therefore, unfortunately, its application is difficult. For this reason, research has been conducted with a focus on the selective catalytic oxidation (SCO) technology and removal of residual NH3 after reaction. The use of co-catalysts to inhibit NH3 oxidation is also related to slip prediction models. However, NH3 can also be oxidized by major substances in the catalyst, such as vanadium, tungsten, and molybdenum; therefore, this factor should also be considered [148, 149]. As the NH3 oxidation reaction is connected to a decrease in N2 selectivity and inhibited NH3-SCR reaction, it causes a decrease in the catalyst activity [150]. The NH3 oxidation reaction proceeds via the following reaction pathway.

$${\text{4NH}}_{{3}} + {\text{ 4NO }} + {\text{ O}}_{{2}} \to {\text{4N}}_{{2}} + {\text{ 6H}}_{{2}} {\text{O }}\left( {\text{direct oxidation}} \right)$$
(7)
$${\text{NH}}_{{3}} + { 5}/{\text{4O}}_{{2}} \to {\text{NO }} + {3}/{\text{2 H}}_{{2}} {\text{O}}$$
(8)
$${\text{2NH}}_{{3}} + {\text{ 2O}}_{{2}} \to {\text{N}}_{{2}} {\text{O }} + {\text{ 3H}}_{{2}} {\text{O}}$$
(9)

In the case of an indirect reaction from NH3 to N2 and H2O, as shown in Eq. 79, there is a problem related to the generation of by-products, such as NO and N2O. Therefore, such a reaction must be suppressed to prevent activity degradation, as discussed above [151]. Epling et al. [152] and Blanco [153] have investigated the presence of an indirect path or direct oxidation to N2 after the formation of NO [152] or N2O (see Fig. 15) [153] during oxidation from NH3. Indeed, to inhibit the NH3-oxidation, many studies have been conducted on mechanisms concerning oxidation and adsorption to examine the doping effects of W [121, 126, 133], Cu [113, 114], Ce [113, 120], Fe [113, 120], and Ru [122, 127]. The studies confirmed the kinetic scheme of NH3/NOx ratio, temperature, vanadium loading influence (see Fig. 16), and NH3-slip. In fact, the direct factors include the intrinsic rate of NH3 oxidation and the volume of catalyst.. In fact, the direct factors depend on the intrinsic rate of NH3 oxidation and the catalyst volume. Additionally, the catalyst monolith thickness and diffusion of NH3 were studied. In other works reported the NH3 oxidation depended on N2 selectivity. According to the research on NH3 reaction, ammonia oxidation increases with an increase in the reaction temperature and vanadium content, and the N2 selectivity decreases rapidly. Grange et al. [154] reported the relationship between NH3 oxidation and SCR catalyst through the DRIFT study and confirmed the connection with the V = O octagon band in DRIFTS, particularly the temperature of the SCR reaction, the de-NOX efficiency, and oxidation in the vanadium-based catalyst (see Fig. 17). Furthermore, Wang et al. [155] reported on the relations between ammonia storage and slip and gas hourly space velocity (GHSV) and temperature, among other factors. This research indicated that GHSV affects NH3 slip by affecting the temporary NH3 reaction and the heating rate of the catalyst. Epling et al. [152] reported that Cu demonstrates the advantage of easier control of NH3 oxidation than Fe as the storage capacity for NH3 and NOx was higher than that of the Fe catalyst. In the V–Mo-based catalyst using Ru [156, 157], Ru inhibits NH3 slip at 350 °C by increasing NH3 decomposition efficiency and improves N2 selectivity to 97%. Additionally, depending on the presence of SO2, NH3 oxidation was partially inhibited. Moreover, a study on NH3 oxidation inhibition was conducted by doping with CaO, a catalyst poison. This research indicated that the addition of CaO improves NO formation by NH3 oxidation and inhibits N2O formation, thereby changing the NH3 reaction pathway. As a result, doping with CaO improves the de-NOX efficiency and N2 selectivity [150]. However, as CaO is generally considered to be a poison substance in catalysts, more complementary research is necessary. In summary, it is essential to control the vanadium content and the amount of NH3 adsorbed [158]. Additionally, it is important to consider the influence of exhaust gas and flow rate [155], such as NH3/NOx for inhibiting the NH3 oxidation. However, it will be necessary that the ongoing researches focus on minimizing the NH3 slip and developing the SCO catalyst because of the lack of related research at present.

Fig. 14
figure 14

Influence of NH3 concentration on SO3 formation

Fig. 15
figure 15

N2 selectivity to a N2 and b N2O during NH3 oxidation as a function of reaction temperature [153]

Fig. 16
figure 16

Reaction temperature effect of NH3 conversion, operating condition: space velocity at normal conditions (GHSV), 15,000 h−1; linear velocity, 0.1 ms−1; pressure, 0.1 MPa. Feed composition: [NH3] = 500 ppm; [O2] = 3 vol%; [Ar] = balance [153]

Fig. 17
figure 17

NH3 oxidation on vanadium catalyst [154]

6 Conclusions and outlook

Vanadium-based catalysts with W or Mo co-catalysts and TiO2 supports have been used as commercial NH3-SCR catalysts without any major changes to the composition for many years, despite research into new compositions. However, air quality standards are being strengthened worldwide, which has renewed interest in the development of SCR catalysts with high NOx reduction efficiencies and excellent poisoning resistance at low operating temperatures. This review discusses the current status of research on commercial vanadium-based SCR catalysts used to reduce NOx emissions from stationary sources worldwide.

The diverse operating conditions of SCR catalysts in actual industrial sites necessitates the tailoring of catalyst properties for a given application. For example, catalysts with physicochemical poisoning resistance to different poisoning substances (mercury, alkali metals, particulate matter, sulfur, etc.) over a wide range of operating temperatures are required; therefore, several methods have been developed to design suitable catalysts based on the intended application area. In addition, there are several methods for increasing the efficiency of SCR catalysts, including the use of different co-catalysts (W, Mo, Ba, Ce, Sb, Fe, Mn, etc.), supports (TiO2, Al2O3, CeO2, activated carbon, etc.), and vanadium species (polymeric, monomeric, and crystalline V2O5); tailoring the surface density, support application (carbon based materials, two-dimensional materials), and particle structure; and surface modification. In addition, although much research has been conducted on powder-type catalysts, this review discussed several other monolith catalysts (honeycomb-type, plate, corrugated, etc.) that may help to meet the commercial catalyst requirements for a given application. Catalyst monolith forms can be optimized for the specific operating and exhaust conditions of a given application; therefore, different catalyst morphologies are used depending on the place of use. We discussed the correlation among the reaction, geometry, and morphology of monolith catalysts to improve their performance and properties under variable operating conditions.

Environmental regulations are set to become continually stricter in the future. Therefore, improved methods of controlling the amount of N2O and unburned NH3 emitted by SCR side reactions will be required. Currently, as fossil fuels are used in industrial plants, NH3, an unreacted reducing agent, is mainly emitted along with NOx, CO, VOCs, UHC, and particulate matter. Therefore, there is a need to develop oxidation–reduction catalysts or module systems that are capable of reducing these pollutants simultaneously in the SCR facility. In addition, as more governments and enterprises seek to achieve carbon neutrality, new combustion technologies based on low-carbon or non-carbon fuels such as ammonia or hydrogen with conventional LNG are being applied to replace fossil fuels. These new fuels change the composition of the flue gas; however, thermal NOx is still emitted in large quantities from combustion reactions. Consequently, continuous research is needed to ensure SCR catalyst technologies keep up with the changing field of fuel combustion. Importantly, under co-firing conditions, pollutants such as CO and unreacted NH3 may be generated due to incomplete combustion. Therefore, in the future, it will be necessary to not only improve the performance of SCR catalysts, but also to keep pace with new developments by designing functional reduction catalysts that can simultaneously reduce various pollutants together with NOx.

Availability of data and materials

Not applicable.

References

  1. J. Seo, Y.B. Lim, D. Youn, J.Y. Kim, H.C. Jin, Synergistic enhancement of urban haze by nitrate uptake into transported hygroscopic particles in the Asian continental outflow. Atmos. Chem. Phys. 20, 7575–7594 (2020)

    Article  CAS  Google Scholar 

  2. R.M. Heck, Catalytic abatement of nitrogen oxides–stationary applications. Catal. Today 53, 519–523 (1999)

    Article  CAS  Google Scholar 

  3. J.C.C. Yu, V.-H. Nguyen, J. Lasek, S.-W. Chiang, D.X. Li, J.C.S. Wu, NOx abatement from stationary emission sources by photo-assisted SCR: Lab-scale to pilot-scale studies. Appl. Catal. A 523, 294–303 (2016)

    Article  CAS  Google Scholar 

  4. U. Asghar et al., Review on the progress in emission control technologies for the abatement of CO2, SOx and NOx from fuel combustion. J. Environ. Chem. Eng. (2021). https://doi.org/10.1016/j.jece.2021.106064

    Article  Google Scholar 

  5. C. Tang, H. Zhang, L. Dong, Ceria-based catalysts for low-temperature selective catalytic reduction of NO with NH3. Catal. Sci. Technol. 6, 1248–1264 (2016)

    Article  CAS  Google Scholar 

  6. X. Yang, Y. Liao, Y. Wang, X. Chen, X. Ma, Research of coupling technologies on NOx reduction in a municipal solid waste incinerator. Fuel 314, 122769 (2022)

    Article  CAS  Google Scholar 

  7. B. Ramachandran, R.G. Herman, S. Choi, H.G. Stenger, C.E. Lyman, J.W. Sale, Testing zeolite SCR catalysts under protocol conditions for NOx abatement from stationary emission sources. Catal. Today 55, 281–290 (2000)

    Article  CAS  Google Scholar 

  8. B. Ye, M. Lee, B. Jeong, J. Kim, D.H. Lee, J.M. Baik, H.-D. Kim, Partially reduced graphene oxide as a support of Mn-Ce/TiO2 catalyst for selective catalytic reduction of NOx with NH3. Catal. Today 328, 300–306 (2019)

    Article  CAS  Google Scholar 

  9. L. Huang, Y. Zeng, Y. Gao, H. Wang, Y. Zong, Z. Chang, S. Zhang, P. Han, Y. Yu, Promotional effect of phosphorus addition on improving the SO2 resistance of V2O5-MoO3/TiO2 catalyst for NH3-SCR of NO. J. Phys. Chem. Solids 163, 110566 (2022)

    Article  CAS  Google Scholar 

  10. K. Guo, J. Ji, W. Song, J. Sun, C. Tang, L. Dong, Conquering ammonium bisulfate poison over low-temperature NH3-SCR catalysts: a critical review. Appl. Catal. B 297, 120388 (2021)

    Article  CAS  Google Scholar 

  11. B.G. Miller, Emissions control strategies for power plants clean coal engineering technology (Elsevier, Amsterdam, 2011)

    Google Scholar 

  12. A.-C. Larsson, J. Einvall, M. Sanati, Deactivation of SCR catalysts by exposure to aerosol particles of potassium and zinc salts. Aerosol Sci. Technol. 41, 369–379 (2007)

    Article  CAS  Google Scholar 

  13. C.J.G. van der Grift, A.F. Woldhuis, O.L. Maaskant, The shell DENOX system for low temperature NOx removal. Catal. Today 27, 23–27 (1996)

    Article  Google Scholar 

  14. P. Forzatti, I. Nova, E. Tronconi, A. Kustov, J.R. Thøgersen, Effect of operating variables on the enhanced SCR reaction over a commercial V2O5–WO3/TiO2 catalyst for stationary applications. Catal. Today 184, 153–159 (2012)

    Article  CAS  Google Scholar 

  15. Y. Yu, C. He, J. Chen, L. Yin, T. Qiu, X. Meng, Regeneration of deactivated commercial SCR catalyst by alkali washing. Catal. Commun. 39, 78–81 (2013)

    Article  CAS  Google Scholar 

  16. K.C. Galbreath, C.J. Zygarlicke, Mercury transformations in coal combustion flue gas. Fuel Process. Technol. 65–66, 289–310 (2000)

    Article  Google Scholar 

  17. B. Li, H. Wang, Y. Xu, J. Xue, Study on mercury oxidation by SCR catalyst in coal-fired power plant. Energy Proced. 141, 339–344 (2017)

    Article  CAS  Google Scholar 

  18. News & Technology for the Global Energy Industry (POWER, 2011), https://www.powermag.com/power-101-improving-the-performance-of-boiler-auxiliaries-part-ii/. Accessed 1 Feb 2011

  19. Y. Zheng, A.D. Jensen, J.E. Johnsson, Deactivation of V2O5-WO3-TiO2 SCR catalyst at a biomass-fired combined heat and power plant. Appl. Catal. B 60, 253–264 (2005)

    Article  CAS  Google Scholar 

  20. B. Sander, Properties of danish biofuels and the requirements for power production. Biomass Bioenergy 12, 177–183 (1997)

    Article  CAS  Google Scholar 

  21. F. Castellino, S.B. Rasmussen, A.D. Jensen, J.E. Johnsson, R. Fehrmann, Deactivation of vanadia-based commercial SCR catalysts by polyphosphoric acids. Appl. Catal. B 83, 110–122 (2008)

    Article  CAS  Google Scholar 

  22. Y. Zheng, A.D. Jensen, J.E. Johnsson, J.R. Thøgersen, Deactivation of V2O5-WO3-TiO2 SCR catalyst at biomass fired power plants: elucidation of mechanisms by lab- and pilot-scale experiments. Appl. Catal. B 83, 186–194 (2008)

    Article  CAS  Google Scholar 

  23. F. Castellino, A.D. Jensen, J.E. Johnsson, R. Fehrmann, Influence of reaction products of K-getter fuel additives on commercial vanadia-based SCR catalysts. Appl. Catal. B 86, 196–205 (2009)

    Article  CAS  Google Scholar 

  24. F. Castellino, A.D. Jensen, J.E. Johnsson, R. Fehrmann, Influence of reaction products of K-getter fuel additives on commercial vanadia-based SCR catalysts. Appl. Catal. B 86, 206–215 (2009)

    Article  CAS  Google Scholar 

  25. M. Klimczak, P. Kern, T. Heinzelmann, M. Lucas, P. Claus, High-throughput study of the effects of inorganic additives and poisons on NH3-SCR catalysts—Part I: V2O5–WO3/TiO2 catalysts. Appl. Catal. B 95, 39–47 (2010)

    Article  CAS  Google Scholar 

  26. J.W. Kim, B. Ye, M. Lee, B. Jung, D.H. Lee, H. Lee, H.-D. Kim, Adsorption/desorption performance and NOx removal efficiency of BaO loaded V2O5-WO3/TiO2 selective catalytic reduction catalyst. Korean J. Met. Mater. 57, 270–277 (2019)

    Article  CAS  Google Scholar 

  27. J.G.M. Brandin, C.U.I. Odenbrand, Deactivation and characterization of SCR catalysts used in municipal waste incineration applications. Catal. Lett. 148, 312–327 (2017)

    Article  Google Scholar 

  28. J.G.M. Brandin, C.U.I. Odenbrand, Poisoning of SCR catalysts used in municipal waste incineration applications. Top. Catal. 60, 1306–1316 (2017)

    Article  CAS  Google Scholar 

  29. J. Cao et al., Influence of phosphorus on the NH3-SCR performance of CeO2-TiO2 catalyst for NOx removal from co-incineration flue gas of domestic waste and municipal sludge. J. Colloid Interface Sci. 610, 463–473 (2022)

    Article  CAS  Google Scholar 

  30. H. Jung, E. Park, M. Kim, J. Jurng, Pilot-scale evaluation of a novel TiO2-supported V2O5 catalyst for DeNOx at low temperatures at a waste incinerator. Waste Manage. 61, 283–287 (2017)

    Article  CAS  Google Scholar 

  31. S. Cimino, C. Ferone, R. Cioffi, G. Perillo, L. Lisi, A case study for the deactivation and regeneration of a V2O5-WO3/TiO2 catalyst in a tail-end scr unit of a municipal waste incineration plant. Catalysts (2019). https://doi.org/10.3390/catal9050464

    Article  Google Scholar 

  32. Y. Guo, L. Luo, Y. Zheng, J. Wang, T. Zhu, Low-medium temperature application of selective catalytic reduction denitration in cement flue gas through a pilot plant. Chemosphere 276, 130182 (2021)

    Article  CAS  Google Scholar 

  33. H. Wu, J. Cai, Q. Ren, C. Shi, A. Zhao, Q. Lyu, A thermal and chemical fuel pretreatment process for NOx reduction from cement kiln. Fuel Process. Technol. (2020). https://doi.org/10.1016/j.fuproc.2020.106556

    Article  Google Scholar 

  34. J.H. Kim, J.H. Kim, H.S. Kim, H.J. Kim, S.H. Kang, J.H. Ryu, S.S. Shim, Reduction of NOx emission from the cement industry in south korea: a review. Atmosphere (2022). https://doi.org/10.3390/atmos13010121

    Article  Google Scholar 

  35. D. Wang, J. Luo, Q. Yang, J. Yan, K. Zhang, W. Zhang, Y. Peng, J. Li, J. Crittenden, Deactivation mechanism of multipoisons in cement furnace flue gas on selective catalytic reduction catalysts. Environ. Sci. Technol. 53, 6937–6944 (2019)

    Article  CAS  Google Scholar 

  36. X. Ding et al., Unexpectedly increased particle emissions from the steel industry determined by wet/semidry/dry flue gas desulfurization technologies. Environ. Sci. Technol. 53, 10361–10370 (2019)

    Article  CAS  Google Scholar 

  37. E. Mousa, C. Wang, J. Riesbeck, M. Larsson, Biomass applications in iron and steel industry: an overview of challenges and opportunities. Renew. Sustain. Energy Rev. 65, 1247–1266 (2016)

    Article  Google Scholar 

  38. Y.-J. Shi, H. Shu, Y.-H. Zhang, H.-M. Fan, Y.-P. Zhang, L.-J. Yang, Formation and decomposition of NH4HSO4 during selective catalytic reduction of NO with NH3 over V2O5-WO3/TiO2 catalysts. Fuel Process. Technol. 150, 141–147 (2016)

    Article  CAS  Google Scholar 

  39. D.W. Kwon, K.H. Park, S.C. Hong, Enhancement of SCR activity and SO2 resistance on VOx /TiO2 catalyst by addition of molybdenum. Chem. Eng. J. 284, 315–324 (2016)

    Article  CAS  Google Scholar 

  40. C. Gao, J.-W. Shi, Z. Fan, G. Gao, C. Niu, Sulfur and water resistance of Mn-based catalysts for low-temperature selective catalytic reduction of NOx: a review. Catalysts (2018). https://doi.org/10.3390/catal8010011

    Article  Google Scholar 

  41. J. Wang, W. Zhong, Simultaneous desulfurization and denitrification of sintering flue gas via composite absorbent. Chin. J. Chem. Eng. 24, 1104–1111 (2016)

    Article  CAS  Google Scholar 

  42. W. Si et al., Sn-doped rutile TiO2 for vanadyl catalysts: Improvements on activity and stability in SCR reaction. Appl. Catal. B 269, 118797 (2020)

    Article  CAS  Google Scholar 

  43. B. Zhang, M. Liebau, W. Suprun, B. Liu, S. Zhang, R. Gläser, Suppression of N2O formation by H2O and SO2 in the selective catalytic reduction of NO with NH3 over a Mn/Ti–Si catalyst. Catal. Sci. Technol. 9, 4759–4770 (2019)

    Article  CAS  Google Scholar 

  44. Y. Peng, C. Wang, J. Li, Structure–activity relationship of VOx/CeO2 nanorod for NO removal with ammonia. Appl. Catal. B 144, 538–546 (2014)

    Article  CAS  Google Scholar 

  45. D. Ye, R. Qu, H. Song, C. Zheng, X. Gao, Z. Luo, M. Ni, K. Cen, Investigation of the promotion effect of WO3 on the decomposition and reactivity of NH4HSO4 with NO on V2O5–WO3/TiO2 SCR catalysts. RSC Adv. 6, 55584–55592 (2016)

    Article  CAS  Google Scholar 

  46. N.V. Economidis, D.A. Pena, P.G. Smirniotis, Comparison of TiO2-based oxide catalysts for the selective catalytic reduction of NO: effect of aging the vanadium precursor solution. Appl. Catal. B Environ. 23, 123–134 (1999)

    Article  CAS  Google Scholar 

  47. J. Duehansen, S. Boghosian, A. Kustov, P. Fristrup, G. Tsilomelekis, K. Stahl, C. Christensen, R. Fehrmann, Vanadia-based SCR catalysts supported on tungstated and sulfated zirconia: Influence of doping with potassium. J. Catal. 251, 459–473 (2007)

    Article  CAS  Google Scholar 

  48. C. Chen, Y. Cao, S. Liu, J. Chen, W. Jia, Review on the latest developments in modified vanadium-titanium-based SCR catalysts. Chin. J. Catal. 39, 1347–1365 (2018)

    Article  CAS  Google Scholar 

  49. M.A. Larrubia, G. Ramis, G. Busca, An FT-IR study of the adsorption of urea and ammonia over V2O5–MoO3–TiO2 SCR catalysts. Appl. Catal. B 27, L145–L151 (2000)

    Article  CAS  Google Scholar 

  50. T. Liuqing, Y. Daiqi, L. Hong, Catalytic performance of a novel ceramic-supported vanadium oxide catalyst for NO reduction with NH3. Catal. Today 78, 159–170 (2003)

    Article  Google Scholar 

  51. L. Lietti, I. Nova, P. Forzatti, Selective catalytic reduction (SCR) of NO by NH3 over TiO2-supported V2O5-WO3 and V2O5-MoO3 cataylsts. Top. Catal. 11–12, 111–122 (2000)

    Article  Google Scholar 

  52. S.T. Choo, S.D. Yim, I.-S. Nam, S.-W. Ham, J.-B. Lee, Effect of promoters including WO3 and BaO on the activity and durability of V2O5/sulfated TiO2 catalyst for NO reduction by NH3. Appl. Catal. B 44, 237–252 (2003)

    Article  CAS  Google Scholar 

  53. L. Lietti, I. Nova, G. Ramis, L. Dall’Acqua, G. Busca, E. Giamello, P. Forzatti, F. Bregani, Characterization and reactivity of V2O5–MoO3/TiO2 de-NOx SCR catalysts. J. Catal. 187, 419–435 (1999)

    Article  CAS  Google Scholar 

  54. L. Huang, Y. Zong, H. Wang, Q. Li, T. Chen, L. Dong, W. Zou, K. Guo, Influence of calcination temperature on the plate-type V2O5–MoO3/TiO2 catalyst for selective catalytic reduction of NO. React. Kinet. Mech. Catal. 124, 603–617 (2018)

    Article  CAS  Google Scholar 

  55. S. Zhang, B. Zhang, B. Liu, S. Sun, A review of Mn-containing oxide catalysts for low temperature selective catalytic reduction of NOx with NH3: reaction mechanism and catalyst deactivation. RSC Adv. 7, 26226–26242 (2017)

    Article  CAS  Google Scholar 

  56. X. Liu, Z. Zhao, R. Ning, Y. Qin, T. Zhu, F. Liu, Ce-Doped V2O5-WO3/TiO2 with Low Vanadium Loadings as SCR Catalysts and the Resistance of H2O and SO2. Catal. Lett. 150, 375–383 (2019)

    Article  Google Scholar 

  57. Z. Liu, S. Zhang, J. Li, J. Zhu, L. Ma, Novel V2O5–CeO2/TiO2 catalyst with low vanadium loading for the selective catalytic reduction of NOx by NH3. Appl. Catal. B 158–159, 11–19 (2014)

    Article  Google Scholar 

  58. Y. Zhang, X. Zhu, K. Shen, H. Xu, K. Sun, C. Zhou, Influence of ceria modification on the properties of TiO2-ZrO2 supported V2O5 catalysts for selective catalytic reduction of NO by NH3. J. Colloid Interface Sci. 376, 233–238 (2012)

    Article  CAS  Google Scholar 

  59. Z. Zhu, Z. Liu, S. Liu, H. Niu, Catalytic NO reduction with ammonia at low tempretatures on V2O5;AC catalysts; effect of metal oxides addition and SO2. Appl. Catal. B 30, 67–276 (2001)

    Article  Google Scholar 

  60. X. Zhao, L. Huang, H. Li, H. Hu, X. Hu, L. Shi, D. Zhang, Promotional effects of zirconium doped CeVO4 for the low-temperature selective catalytic reduction of NOx with NH3. Appl. Catal. B 183, 269–281 (2016)

    Article  CAS  Google Scholar 

  61. D. Ye, R. Qu, C. Zheng, K. Cen, X. Gao, Mechanistic investigation of enhanced reactivity of NH4HSO4 and NO on Nb- and Sb-doped VW/Ti SCR catalysts. Appl. Catal. A 549, 310–319 (2018)

    Article  CAS  Google Scholar 

  62. D.W. Kwon, D.H. Kim, S. Lee, J. Kim, H.P. Ha, A dual catalytic strategy by the nature of the functionalization effect as well as active species on vanadium-based catalyst for enhanced low temperature SCR. Appl. Catal. B Environ. (2021). https://doi.org/10.1016/j.apcatb.2021.120032

    Article  Google Scholar 

  63. D.W. Kwon, K.B. Nam, S.C. Hong, The role of ceria on the activity and SO2 resistance of catalysts for the selective catalytic reduction of NOx by NH3. Appl. Catal B 166–167, 37–44 (2015)

    Article  Google Scholar 

  64. P.A. Kumar, Y.E. Jeong, H.P. Ha, Low temperature NH3 -SCR activity enhancement of antimony promoted vanadia-ceria catalyst. Catal. Today 293–294, 61–72 (2017)

    Article  Google Scholar 

  65. E. García-Bordejé, A. Monzón, M.J. Lázaro, R. Moliner, Promotion by a second metal or SO2 over vanadium supported on mesoporous carbon-coated monoliths for the SCR of NO at low temperature. Catal. Today 102–103, 177–182 (2005)

    Article  Google Scholar 

  66. Q.-M. Zhang, C.-L. Song, G. Lv, F. Bin, H.-T. Pang, J.-O. Song, Effect of metal oxide partial substitution of V2O5 in V2O5–WO3/TiO2 on selective catalytic reduction of NO with NH3. J. Ind. Eng. Chem. 24, 79–86 (2015)

    Article  Google Scholar 

  67. D. Zhang, Z. Ma, B. Wang, Q. Sun, W. Xu, T. Zhu, Effects of MO (M=Mn, Cu, Sb, La) on V-Mo–Ce/Ti selective catalytic reduction catalysts. J. Rare Earths 38, 157–166 (2020)

    Article  CAS  Google Scholar 

  68. M.A. Centeno, I. Carrizosa, J.A. Odriozola, NH3 adsorptino over lanthanide doped V2O5/Al2O3 catalysts. J. Alloy. Compd. 323–324, 597–600 (2001)

    Article  Google Scholar 

  69. M. Casanova, K. Schermanz, J. Llorca, A. Trovarelli, Improved high temperature stability of NH3-SCR catalysts based on rare earth vanadates supported on TiO2-WO3-SiO2. Catal. Today 184, 227–236 (2012)

    Article  CAS  Google Scholar 

  70. D.W. Kwon, S. Lee, J. Kim, K.-Y. Lee, H.P. Ha, Influence of support composition on enhancing the performance of Ce-V on TiO2 comprised tungsten-silica for NH3-SCR. Catal. Today 359, 112–123 (2021)

    Article  CAS  Google Scholar 

  71. M. Casanova, J. Llorca, A. Sagar, K. Schermanz, A. Trovarelli, Mixed iron–erbium vanadate NH3-SCR catalysts. Catal. Today 241, 159–168 (2015)

    Article  CAS  Google Scholar 

  72. Y. Zhang, W. Xiong, W. Chen, Y. Zheng, Recent progress on vanadium dioxide nanostructures and devices: fabrication, properties applications and perspectives. Nanomaterials (2021). https://doi.org/10.3390/nano11020338

    Article  Google Scholar 

  73. M. Kobayashi, R. Kuma, S. Masaki, N. Sugishima, TiO2-SiO2 and V2O5/TiO2-SiO2 catalyst: physico-chemical characteristics and catalytic behavior in selective catalytic reduction of NO by NH3. Appl. Catal. B 60, 173–179 (2005)

    Article  CAS  Google Scholar 

  74. S. Youn, S. Jeong, D.H. Kim, Effect of oxidation states of vanadium precursor solution in V2O5/TiO2 catalysts for low temperature NH3 selective catalytic reduction. Catal. Today 232, 185–191 (2014)

    Article  CAS  Google Scholar 

  75. M.S. Lee, S.I. Kim, M.J. Lee, B. Ye, T. Kim, H.D. Kim, J.W. Lee, D.H. Lee, Effect of catalyst crystallinity on V-based selective catalytic reduction with ammonia. Nanomaterials (2021). https://doi.org/10.3390/nano11061452

    Article  Google Scholar 

  76. Y. Inomata et al., Bulk vanadium oxide versus conventional V2O5/TiO2: NH3–SCR catalysts Working at a low temperature below 150°C. ACS Catal. 9, 9327–9331 (2019)

    Article  CAS  Google Scholar 

  77. I.E. Wachs, The generality of surface vanadium oxide phases in mixed oxide catalysts. Appl. Catal. A 391, 36–42 (2011)

    Article  CAS  Google Scholar 

  78. J.-K. Lai, I.E. Wachs, A Perspective on the selective catalytic reduction (SCR) of NO with NH3 by supported V2O5–WO3/TiO2 catalysts. ACS Catal. 8, 6537–6551 (2018)

    Article  CAS  Google Scholar 

  79. G. He, Z. Lian, Y. Yu, Y. Yang, K. Liu, X. Shi, Z. Yan, W. Shan, H. He, Polymeric vanadyl species determine the low-temperature activity of V-based catalysts for the SCR of NOx with NH3. Sci. Adv. (2018). https://doi.org/10.1126/sciadv.aau4637

    Article  Google Scholar 

  80. M. Foix, C. Guyon, M. Tatoulian, P. Da Costa, Study of the use of fluidized bed plasma reactors for the treatment of alumina supported palladium catalyst: application for SCR NOx by CH4 in stationary sources. Catal. Commun. 12, 20–24 (2010)

    Article  CAS  Google Scholar 

  81. J.Y. Lee, S.H. Hong, S.P. Cho, S.C. Hong, The study of deNOx catalyst in low temperature using nano-sized supports. Curr. Appl. Phys. 6, 996–1001 (2006)

    Article  Google Scholar 

  82. Z. Zhu, Z. Liu, S. Liu, H. Niu, A novel carbon-supported vanadium oxide catalyst for NO reduction with NH3 at low temperatures. Appl. Catal. B Environ. 23, L229–L233 (1999)

    Article  CAS  Google Scholar 

  83. Z. Zhu, Z. Liu, H. Niu, S. Liu, T. Hu, T. Liu, Y. Xie, Mechanism of SO2 promotion for NO reduction with NH3 over activated carbon-supported vanadium oxide catalyst. J. Catal. 197, 6–16 (2001)

    Article  CAS  Google Scholar 

  84. Z. Zhu, Z. Liu, H. Niu, S. Liu, Promoting effect of SO2 on activated carbon-supported vanadia catalyst for NO reduction by NH3 at low temperatures. J. Catal. 187, 245–248 (1999)

    Article  CAS  Google Scholar 

  85. B. Huang, R. Huang, D. Jin, D. Ye, Low temperature SCR of NO with NH3 over carbon nanotubes supported vanadium oxides. Catal. Today 126, 279–283 (2007)

    Article  CAS  Google Scholar 

  86. E. Garciabordeje, J. Pinilla, M. Lazaro, R. Moliner, J. Fierro, Role of sulphates on the mechanism of NH3-SCR of NO at low temperatures over presulphated vanadium supported on carbon-coated monoliths. J. Catal. 233, 166–175 (2005)

    Article  CAS  Google Scholar 

  87. P. Wauthoz, M. Ruwet, T. Machej, P. Grange, Influence of the preparation method on the V2O5/TiO2/SiO2 cataylsts in selective catalytic reduction of nitric oxide with ammonia. Appl. Catal. 69, 149–167 (1991)

    Article  CAS  Google Scholar 

  88. B. Ye, S.-I. Kim, M. Lee, M. Ezazi, H.-D. Kim, G. Kwon, D.H. Lee, Synthesis of oxygen functionalized carbon nanotubes and their application for selective catalytic reduction of NOx with NH3. RSC Adv. 10, 16700–16708 (2020)

    Article  CAS  Google Scholar 

  89. J.Y. Choi, B. Jeong, E.S. Kim, H.Y. Chun, D.W. Shin, D.H. Kim, H.D. Kim, Synthesis of vanadium oxide supported on reduced graphene oxide for De-Nox. J Nanosci Nanotechnol 15, 9083–9087 (2015)

    Article  CAS  Google Scholar 

  90. G. Lee, B. Ye, M.-J. Lee, S.-Y. Chun, B. Jeong, H.-D. Kim, J. Jae, T. Kim, Selective catalytic reduction of NO by NH3 over V2O5-WO3 supported by titanium isopropoxide (TTIP)-treated TiO2. J. Ind. Eng. Chem. 109, 422–430 (2022)

    Article  CAS  Google Scholar 

  91. H. Jeong, O. Kwon, B.-S. Kim, J. Bae, S. Shin, H.-E. Kim, J. Kim, H. Lee, Highly durable metal ensemble catalysts with full dispersion for automotive applications beyond single-atom catalysts. Nat. Catal. 3, 368–375 (2020)

    Article  CAS  Google Scholar 

  92. M. Lee, B. Ye, B. Jeong, H.-Y. Chun, D.H. Lee, S.-S. Park, H. Lee, H.-D. Kim, Reduced graphene oxide supported V2O5-WO3-TiO2 catalysts for selective catalytic reduction of NOx. Korean J. Chem. Eng. 35, 1988–1993 (2018)

    Article  CAS  Google Scholar 

  93. B. Jeong, B. Ye, E.-S. Kim, H.-D. Kim, Characteristics of selective catalytic reduction (SCR) catalyst adding graphene-tungsten nanocomposite. Catal. Commun. 93, 15–19 (2017)

    Article  CAS  Google Scholar 

  94. B. Ye, B. Jeong, M. Lee, H.-D. Kim, J.M. Baik, Low-temperature De-NOx extruded monolithic catalysts based on highly dispersive Mn-Ce Oxide nanoparticles of low Ce content. Adv. Mater. Technol. (2019). https://doi.org/10.1002/admt.201800462

    Article  Google Scholar 

  95. B. Ye, J. Kim, M.-J. Lee, S.-Y. Chun, B. Jeong, T. Kim, D.H. Lee, H.-D. Kim, Mn-Ce oxide nanoparticles supported on nitrogen-doped reduced graphene oxide as low-temperature catalysts for selective catalytic reduction of nitrogen oxides. Microporous Mesoporous Mater. (2021). https://doi.org/10.1016/j.micromeso.2020.110588

    Article  Google Scholar 

  96. B. Ye, M.-J. Lee, S.-Y. Chun, G. Lee, J. Kim, B. Jeong, T. Kim, H.-D. Kim, Promotional effect of surface treated N-TiO2 as support for VOx-based catalysts on the selective catalytic reduction of NO using NH3. Appl. Surf. Sci. (2021). https://doi.org/10.1016/j.apsusc.2021.149934

    Article  Google Scholar 

  97. M.-J. Lee, B. Ye, B. Jeong, S.-Y. Chun, T. Kim, D.-H. Kim, H. Lee, H.-D. Kim, MnOx–CeOx nanoparticles supported on porous hexagonal boron nitride nanoflakes for selective catalytic reduction of nitrogen oxides. ACS. Appl. Nano Mater. 3, 11254–11265 (2020)

    Article  CAS  Google Scholar 

  98. R. Wu, N. Zhang, X. Liu, L. Li, L. Song, W. Qiu, H. He, The keggin structure: an important factor in governing NH3–SCR activity over the V2O5–MoO3/TiO2 catalyst. Catal. Lett. 148, 1228–1235 (2018)

    Article  CAS  Google Scholar 

  99. H. Kang, M. Wu, S. Li, C. Wei, X. Chen, J. Chen, F. Jing, W. Chu, Y. Liu, Converting poisonous sulfate species to an active promoter on TiO2 predecorated MnOx catalysts for the NH3-SCR reaction. ACS. Appl. Mater. Interfaces 13, 61237–61247 (2021)

    Article  CAS  Google Scholar 

  100. T. Gu, Y. Liu, X. Weng, H. Wang, Z. Wu, The enhanced performance of ceria with surface sulfation for selective catalytic reduction of NO by NH3. Catal. Commun. 12, 310–313 (2010)

    Article  CAS  Google Scholar 

  101. L. Zhang, H. Qu, T. Du, W. Ma, Q. Zhong, H2O and SO2 tolerance, activity and reaction mechanism of sulfated Ni–Ce–La composite oxide nanocrystals in NH3-SCR. Chem. Eng. J. 296, 122–131 (2016)

    Article  CAS  Google Scholar 

  102. F. Liu, K. Asakura, H. He, W. Shan, X. Shi, C. Zhang, Influence of sulfation on iron titanate catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B 103, 369–377 (2011)

    Article  CAS  Google Scholar 

  103. X. Li, Z. Han, X. Wang, S. Yang, G. Liu, Y. Gao, C. Li, Acid treatment of ZrO2-supported CeO2 catalysts for NH3-SCR of NO: influence on surface acidity and reaction mechanism. J. Taiwan Inst. Chem. Eng. (2022). https://doi.org/10.1016/j.jtice.2021.11.011

    Article  Google Scholar 

  104. P. Gong, J. Xie, D. Fang, F. He, F. Li, K. Qi, Enhancement of the NH3-SCR property of Ce-Zr-Ti by surface and structure modification with P. Appl. Surf. Sci. (2020). https://doi.org/10.1016/j.apsusc.2019.144641

    Article  Google Scholar 

  105. P.G. Blakeman, E.M. Burkholder, H.-Y. Chen, J.E. Collier, J.M. Fedeyko, H. Jobson, R.R. Rajaram, The role of pore size on the thermal stability of zeolite supported Cu SCR catalysts. Catal. Today 231, 56–63 (2014)

    Article  CAS  Google Scholar 

  106. X. Hu, M. Yang, D. Fan, G. Qi, J. Wang, J. Wang, T. Yu, W. Li, M. Shen, The role of pore diffusion in determining NH3 SCR active sites over Cu/SAPO-34 catalysts. J. Catal. 341, 55–61 (2016)

    Article  CAS  Google Scholar 

  107. S. Youn, I. Song, H. Lee, S.J. Cho, D.H. Kim, Effect of pore structure of TiO2 on the SO2 poisoning over V2O5/TiO2 catalysts for selective catalytic reduction of NOx with NH3. Catal. Today. 303, 19–24 (2018)

    Article  CAS  Google Scholar 

  108. S. Yu, N. Jiang, W. Zou, L. Li, C. Tang, L. Dong, A general and inherent strategy to improve the water tolerance of low temperature NH3-SCR catalysts via trace SiO2 deposition. Catal. Commun. 84, 75–79 (2016)

    Article  CAS  Google Scholar 

  109. E. Garcı́a-Bordejé, L. Calvillo, M.J. Lázaro, R. Moliner, Vanadium supported on carbon-coated monoliths for the SCR of NO at low temperature: effect of pore structure. Appl. Catal. B Environ. 50, 235–242 (2004)

    Article  Google Scholar 

  110. E. Garciabordeje, Structure of vanadium oxide supported on mesoporous carbon-coated monoliths and relationship with its catalytic performance in the SCR of NO at low temperatures. J. Catal. 223, 395–403 (2004)

    Article  CAS  Google Scholar 

  111. M. Ouzzine, G.A. Cifredo, J.M. Gatica, S. Harti, T. Chafik, H. Vidal, Original carbon-based honeycomb monoliths as support of Cu or Mn catalysts for low-temperature SCR of NO: effects of preparation variables. Appl. Catal A 342, 150–158 (2008)

    Article  CAS  Google Scholar 

  112. B.R. Stanmore, V. Tschamber, J.F. Brilhac, Oxidation of carbon by NOx, with particular reference to NO2 and N2O. Fuel 87, 131–146 (2008)

    Article  CAS  Google Scholar 

  113. G.S. Szymański, T. Grzybek, H. Papp, Influence of nitrogen surface functionalities on the catalytic activity of activated carbon in low temperature SCR of NO with NH3. Catal. Today 90, 51–59 (2004)

    Article  Google Scholar 

  114. B. Samojeden, T. Grzybek, The influence of the promotion of N-modified activated carbon with iron on NO removal by NH3-SCR (selective catalytic reduction). Energy 116, 1484–1491 (2016)

    Article  CAS  Google Scholar 

  115. L. Singoredjo, F. Kapteijn, J.A. Moulijn, J.-M. Martín-Martínez, H.-P. Boehm, Modified activated carbons for the selective catalytic reduction of NO with NH3. Carbon 31, 213–222 (1993)

    Article  CAS  Google Scholar 

  116. J. Muniz, G. Marban, A.B. Fuertes, Low temperature selective catalytic reduction of NO over polyarylamide-based carbon fibres. Appl. Catal. B 23, 25–35 (1999)

    Article  CAS  Google Scholar 

  117. J. Muniz, G. Marban, A.B. Fuertes, Low temperature selective catalytic reduction of NO over modified activated carbon fibres. Appl. Catal. B 27, 27–36 (2000)

    Article  CAS  Google Scholar 

  118. B.J. Ku, J.K. Lee, D. Park, H.-K. Rhee, Treatment of activated carbon to enhance catalytic activity for reduction of nitric oxide with ammonia. Ind. Eng. Chem. Res. 33, 2868–2874 (1994)

    Article  CAS  Google Scholar 

  119. M.-C. Huang, H. Teng, Nitrogen-containing carbons from phenol–formaldehyde resins and their catalytic activity in NO reduction with NH3. Carbon 41, 951–957 (2003)

    Article  CAS  Google Scholar 

  120. P. Zhang, R.T. Guo, L.J. Wu, W.G. Pan, The enhancement of NH3-SCR performance for CeO2 catalyst by CO pretreatment. Environ. Sci. Pollut. Res. Int. 27, 13617–13636 (2020)

    Article  CAS  Google Scholar 

  121. W. Zhao, Q. Zhong, Y. Pan, R. Zhang, Systematic effects of S-doping on the activity of V2O5/TiO2 catalyst for low-temperature NH3-SCR. Chem. Eng. J. 228, 815–823 (2013)

    Article  CAS  Google Scholar 

  122. S. Liu, H. Wang, Y. Wei, R. Zhang, S. Royer, Morphology-oriented ZrO2-supported vanadium oxide for the NH3-SCR process: importance of structural and textural properties. ACS. Appl. Mater. Interfaces 11, 22240–22254 (2019)

    Article  CAS  Google Scholar 

  123. Y. Ma et al., Improved hydrothermal durability of Cu-SSZ-13 NH3-SCR catalyst by surface Al modification: affinity and passivation. J. Catal. 405, 199–211 (2022)

    Article  CAS  Google Scholar 

  124. T. Schwämmle, F. Bertsche, A. Hartung, J. Brandenstein, B. Heidel, G. Scheffknecht, Influence of geometrical parameters of honeycomb commercial SCR-DeNOx-catalysts on DeNOx-activity, mercury oxidation and SO2/SO3-conversion. Chem. Eng. J. 222, 274–281 (2013)

    Article  Google Scholar 

  125. Z. Yang, C. Zheng, X. Zhang, H. Zhou, A.A. Silva, C. Liu, B. Snyder, Y. Wang, X. Gao, Challenge of SO3 removal by wet electrostatic precipitator under simulated flue gas with high SO3 concentration. Fuel 217, 597–604 (2018)

    Article  CAS  Google Scholar 

  126. T. Schwaemmle, B. Heidel, K. Brechtel, G. Scheffknecht, Study of the effect of newly developed mercury oxidation catalysts on the DeNOx-activity and SO2–SO3-conversion. Fuel 101, 179–186 (2012)

    Article  CAS  Google Scholar 

  127. R. Spörl, J. Maier, G. Scheffknecht, Sulphur oxide emissions from dust-fired oxy-fuel combustion of coal. Energy Proced. 37, 1435–1447 (2013)

    Article  Google Scholar 

  128. J. Svachula, L.J. Alemany, N. Ferlazzo, P. Forzatti, E. Tronconi, F. Bregani, Oxidation of SO2 to SO3 over honeycomb DeNoxing catalysts. Ind. Eng. Chem. Res. 32, 826–834 (1993)

    Article  CAS  Google Scholar 

  129. M. Qing et al., Effects of H2O and CO2 on the catalytic oxidation property of V/W/Ti catalysts for SO3 generation. Fuel 237, 545–554 (2019)

    Article  CAS  Google Scholar 

  130. J.P. Dunn, H.G. Stenger Jr., I.E. Wachs, Oxidation of SO2 over Supported Metal Oxide Catalysts. J. Catal. 181, 233–243 (1999)

    Article  CAS  Google Scholar 

  131. J.P. Dunn, P.R. Koppula, H.G. Stenger, I.E. Wachs, Oxidation of sulfur dioxide to sulfur trioxide over supported vanadia catalysts. Appl. Catal. B Environ. 19, 103–117 (1998)

    Article  CAS  Google Scholar 

  132. H. Xiao, C. Dou, H. Shi, J. Ge, L. Cai, Influence of sulfur-containing sodium salt poisoned V2O5–WO3/TiO2 catalysts on SO2–SO3 conversion and NO removal. Catalysts (2018). https://doi.org/10.3390/catal8110541

    Article  Google Scholar 

  133. J. Lu, Z. Zhou, H. Zhang, Z. Yang, Influenced factors study and evaluation for SO2/SO3 conversion rate in SCR process. Fuel 245, 528–533 (2019)

    Article  CAS  Google Scholar 

  134. I. Nova, A. Beretta, G. Groppi, L. Lietti, E. Tronconi, P. Forzatti, Monolithic catalysts for NOx removal from stationary sources, structured catalyst and reactors (Marcel Dekker, New York, 2006), pp.171–214

    Google Scholar 

  135. C. Zheng, L. Xiao, R. Qu, S. Liu, Q. Xin, P. Ji, H. Song, W. Wu, X. Gao, Numerical simulation of selective catalytic reduction of NO and SO2 oxidation in monolith catalyst. Chem. Eng. J. 361, 874–884 (2019)

    Article  CAS  Google Scholar 

  136. J.L. Sorrels, D.D. Randall, K.S. Schaffner, C.R. Fry, Selective catalytic reduction. EPA air pollution control cost manual. 7 (2019)

  137. H.M.S.C.G. Vayenas, Chemistry at catalytic surfaces the SO2 oxidation on Noble metals. J. Catal. 57, 296–314 (1979)

    Article  CAS  Google Scholar 

  138. Y. H, M. Shigeru, T. Kazumasa, K. Sator, Improvement of V2O5_TiO2 catalyst for NOx reduction with NH3 in flue gases. Chem. Lett. 10, 251–254 (1981)

    Article  Google Scholar 

  139. I.N. Pio Forzatti, A. Beretta, Catalytic properties in deNOx and SO2-SO3 reaction. Catal. Today. 56, 431–441 (2000)

    Article  Google Scholar 

  140. L.L. Carlo Orsenigo, E. Tronconi, P. Forzatti, F. Bregani, Dynamic investigation of the role of the surface sulfates in NOx reduction and SO2 oxidation over V2O5_WO3 TiO2 catalysts. Ind. Eng. Chem. Res. 37, 2350–2359 (1998)

    Article  Google Scholar 

  141. T. Wang, J. Liu, Y. Yang, Z. Sui, Y. Zhang, J. Wang, W.P. Pan, Catalytic conversion of mercury over Ce doped Mn/SAPO-34 catalyst: sulphur tolerance and SO2/SO3 conversion. J. Hazard. Mater. 381, 120986 (2020)

    Article  CAS  Google Scholar 

  142. H. Hiroyuki Kamata, K. Takahashi, A. Yukimura, Y. Seo, SO2 oxidation over the V2O5 TiO2 SCR catalyst. Catal. Lett. 73, 79–83 (2001)

    Article  Google Scholar 

  143. L.T.T.N.N. Sazonova, A.V. Simakov, G.B. Barannik, Z.R. Ismagilov, Relationship between sulfur dioxide oxidation and selective catalytic NO reduction by ammonia on V2O5-TiO2 catalysts doped with WO3 and Nb2O5. React. Kinet. Catal. Lett. 52, 101–106 (1994)

    Article  CAS  Google Scholar 

  144. D.W. Kwon, K.H. Park, H.P. Ha, S.C. Hong, The role of molybdenum on the enhanced performance and SO2 resistance of V/Mo-Ti catalysts for NH3-SCR. Appl. Surf. Sci. 481, 1167–1177 (2019)

    Article  CAS  Google Scholar 

  145. H. Xiao, C. Dou, J. Li, Z. Yuan, H. Lv, Experimental study on SO2-to-SO3 conversion over Fe-modified Mn/ZSM-5 catalysts during the catalytic reduction of NOx. Catal. Surv. Asia 23, 332–343 (2019)

    Article  CAS  Google Scholar 

  146. Z. Ma, X. Wu, Y. Feng, Z. Si, D. Weng, L. Shi, Low-temperature SCR activity and SO2 deactivation mechanism of Ce-modified V2O5–WO3/TiO2 catalyst. Prog. Nat. Sci. Mater. Int. 25, 342–352 (2015)

    Article  CAS  Google Scholar 

  147. B. Wang, Y. Bian, S. Feng, S.-Q. Wang, B.-X. Shen, Modification of the V2O5-WO3/TiO2 catalyst with Nb to reduce its activity for SO2 oxidation during the selective catalytic reduction of NO. J Fuel Chem Technol 50, 503–512 (2022)

    Article  Google Scholar 

  148. N. Usberti, M. Jablonska, M.D. Blasi, P. Forzatti, L. Lietti, A. Beretta, Design of a “high-efficiency” NH3-SCR reactor for stationary applications. A kinetic study of NH3 oxidation and NH3-SCR over V-based catalysts. Appl. Catal. B Environ. 179, 185–195 (2015)

    Article  CAS  Google Scholar 

  149. F.O.R.T. Curtin, C. Deconinck, N. Knüttle, B.K. Hodnett, The catalytic oxidation of ammonia influence of water and sulfur on selectivity to nitrogen over promoted copper oxide alumina catalysts. Catal. Today 55, 189–195 (2000)

    Article  CAS  Google Scholar 

  150. Y. Zheng, Y. Guo, J. Wang, L. Luo, T. Zhu, Ca doping effect on the competition of NH3–SCR and NH3 oxidation reactions over vanadium-based catalysts. J. Phys. Chem. C 125, 6128–6136 (2021)

    Article  CAS  Google Scholar 

  151. Q. Liu, Z. Liu, C. Li, Adsorption and activation of NH3 during selective catalytic reduction of NO by NH3. Chin. J. Catal. 27, 636–646 (2006)

    Article  Google Scholar 

  152. F. Piubello, Y. Jangjou, I. Nova, W.S. Epling, Study of NO formation during NH3 oxidation reaction over a Cu-SAPO-34 SCR catalyst. Catal. Lett. 146, 1552–1561 (2016)

    Article  CAS  Google Scholar 

  153. M. Yates, J.A. Martín, M.Á. Martín-Luengo, S. Suárez, J. Blanco, N2O formation in the ammonia oxidation and in the SCR process with V2O5-WO3 catalysts. Catal. Today 107–108, 120–125 (2005)

    Article  Google Scholar 

  154. P.G. Seong Moon Jung, DRIFTS investigation of V=O behavior and its relations with the reactivity of ammonia oxidation and selective catalytic reduction of NO over V2O5 catalyst. Appl. Catal. B Environ. 36, 325–332 (2002)

    Article  Google Scholar 

  155. Y. Zhao, J. Hu, L. Hua, S. Shuai, J. Wang, Ammonia storage and slip in a urea selective catalytic reduction catalyst under steady and transient conditions. Ind. Eng. Chem. Res. 50, 11863–11871 (2011)

    Article  CAS  Google Scholar 

  156. N. Yan, W. Chen, J. Chen, Z. Qu, Y. Guo, S. Yang, J. Jia, Significance of RuO2 modified SCR catalyst for elemental mercury oxidation in coal-fired flue gas. Environ. Sci. Technol. 45, 5725–5730 (2011)

    Article  CAS  Google Scholar 

  157. W. Chen, Y. Ma, N. Yan, Z. Qu, S. Yang, J. Xie, Y. Guo, L. Hu, J. Jia, The co-benefit of elemental mercury oxidation and slip ammonia abatement with SCR-Plus catalysts. Fuel 133, 263–269 (2014)

    Article  CAS  Google Scholar 

  158. A. Ko, Y. Woo, J. Jang, Y. Jung, Y. Pyo, H. Jo, O. Lim, Y.J. Lee, Availability of NH3 adsorption in vanadium-based SCR for reducing NOx emission and NH3 slip. J. Ind. Eng. Chem. 78, 433–439 (2019)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable

Funding

This work was supported by the Ministry of Trade, Industry and Energy (MOTIE) grant funded by Korea government with code numbers (20005721) and (20015619). This research was supported by the National Research Foundation of Korea (grant number: 2019R1C1C1009427), Technology Innovation Program (grant number: 20015991) funded by the Ministry of Trade, Industry and Energy (MOTIE, Korea), the GRRC program of Gyeonggi province (grant number: GRRC 2020-B03).

Author information

Authors and Affiliations

Authors

Contributions

BY and BJ contributed equally to the writing of the manuscript. All authors have contributed to the writing of the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Seunghyun Lee or Hong-Dae Kim.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ye, B., Jeong, B., Lee, Mj. et al. Recent trends in vanadium-based SCR catalysts for NOx reduction in industrial applications: stationary sources. Nano Convergence 9, 51 (2022). https://doi.org/10.1186/s40580-022-00341-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40580-022-00341-7

Keywords