Next Article in Journal
Innovative Turbine Intake Air Cooling Systems and Their Rational Designing
Previous Article in Journal
Three-Dimensional Imaging and Quantification of Gas Storativity in Nanoporous Media via X-rays Computed Tomography
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Deactivation of V2O5−WO3/TiO2 DeNOx Catalyst under Commercial Conditions in Power Production Plant

by
Maciej Zyrkowski
1,2,*,
Monika Motak
1,
Bogdan Samojeden
1,* and
Krzysztof Szczepanek
2
1
Faculty of Energy and Fuels, AGH University of Science and Technology, al. Mickiewicza 30, 30-059 Kraków, Poland
2
PGE Energia Ciepla S.A., ul. Ciepłownicza 1, 31-587 Kraków, Poland
*
Authors to whom correspondence should be addressed.
Energies 2020, 13(23), 6200; https://doi.org/10.3390/en13236200
Submission received: 2 November 2020 / Revised: 16 November 2020 / Accepted: 23 November 2020 / Published: 25 November 2020

Abstract

:
Nitrogen dioxide is one of the most dangerous air pollutants, because its high concentration in air can be directly harmful to human health. It is also responsible for photochemical smog and acid rains. One of the most commonly used techniques to tackle this problem in large combustion plants is selective catalytic reduction (SCR). Commercial SCR installations are often equipped with a V2O5−WO3/TiO2 catalyst. In power plants which utilize a solid fuel boiler, catalysts are exposed to unfavorable conditions. In the paper, factors responsible for deactivation of such a catalyst are comprehensively reviewed where different types of deactivation mechanism, like mechanical, chemical or thermal mechanisms, are separately described. The paper presents the impact of sulfur trioxide and ammonia slip on the catalyst deactivation as well as the problem of ammonium bisulfate formation. The latter is one of the crucial factors influencing the loss of catalytic activity. The majority of issues with fast catalyst deactivation occur when the catalyst work in off-design conditions, in particular in too high or too low temperatures.

1. Introduction

Fossil fuel power plants are subjected to increasingly rigorous NOx emission limits. In Europe there are still plenty of coal-fired power generation units that produce significant amounts of nitrogen oxides. Two major types of NOx removal technology are used in such plants: selective non-catalytic reduction (SNCR) and selective catalytic reduction (SCR). As the name suggests, the latter takes advantage of a catalyst. The decision of choosing one over the other comes down to economic and technical aspects. While SNCR technology is much more affordable for power plant operators, it may struggle to deliver the required NOx emission reduction nowadays. The emission limits in current European Union (EU) regulations are 150 mg/Nm3 for an existing large coal-fired boiler, 180 mg/Nm3 for a medium-sized boiler and 270 mg/Nm3 for small ones. For newly built units, limits are respectively 85, 100 and 150 mg/Nm3 [1]. These limits favor utilization of SCR technology for medium and large existing units, as well as for all newly built ones.
Catalysts are a well-known technology in a variety of industries, and there is still a lot of ongoing research regarding the development of new catalysts and improvement of existing ones. Despite many advantages, catalysts tend to lose their activity over time. Kiełtyka et al. [2] indicated, that in a typical coal-fired power plant with biomass co-firing, V2O5−WO3/TiO2 commercial catalyst can lose 10% of its activity already after 2000 h of operation, and 20% after nearly 5000 h. According to Zheng et al. [3], the rate of deactivation of V2O5−WO3/TiO2 catalyst installed on a straw-fired boiler can be as high as 1% per day of operation. Eventually catalytic layers need to be replaced every 2–3 years (in commercial installations), and new ones are relatively expensive [4,5].
Commercial catalysts installed on coal-fired boilers are subjected to rapidly changing and rather unfavorable conditions. The reactor is usually placed at the end of the boiler second pass (high-dust configuration). There are two main factors that are responsible for SCR catalyst deactivation—flue gas physical parameters and flue gas composition. The former is derived from various process parameters, boiler construction, as well as boiler load. The latter is derived mainly from fuel composition and the combustion process itself. The composition of flue gas is responsible for chemical deactivation of the catalyst (poisoning, plugging) as well as mechanical deactivation (erosion caused by coarse particles). Flue gas physical parameters are responsible for thermal deactivation (sintering due to the high temperature) and mechanical deactivation (uneven flue gases distribution) [5,6]. During the boiler operation, it is feasible to control both factors to a certain extent, depending on the boiler. One of the most commonly used catalysts in the power industry are vanadium-based catalysts, with 1 ÷ 2.5 wt.% of V2O5 as an active material, 8 wt.% of WO3 as a promoter and anatase TiO2 as a supporter. They can be a plate-type or have the shape of a honeycomb [7]. Vanadium concentration is crucial for the NOx reduction [8], however its high content may cause loss of the selectivity and thermal stability of the catalyst, as well as increased conversion of SO2 to SO3 [9,10,11,12]. The major drawback of V2O5-WO3/TiO2 catalyst is its toxicity and the poor stability of vanadium, reduced activity at low temperature (below 250 °C) as well as a relatively narrow temperature window [10,13,14]. However, the catalyst is known for its high activity and resistance to SO2 [5,15,16,17,18]. According to literature, V2O5−WO3/TiO2 catalysts work best in temperatures 300 ÷ 400 °C [19,20]. Operation at higher temperatures can cause sintering, while operation at lower temperatures may result in catalyst plugging due to ammonium bisulfate formation [21,22].
This paper gives detailed overview of the commercial V2O5−WO3/TiO2 catalysts used in DeNOx process in industrial boilers, with focus on the mechanisms of its deactivation. The mechanisms of SO2 oxidation, as well as ammonia slip from SCR reactors, are also described, as they are vital to understand the catalyst deactivation process. Furthermore, the authors describe the effect of boiler operation conditions on the overall catalyst performance.

2. Catalyst Performance

Overall, an SCR catalyst’s performance can be simply described by its ability to reduce concentration of NOx from flue gas stream. This is a derivative of the catalyst’s activity, which for heterogeneous catalysts can be approximated as the amount of active centers that the catalyst has. Active sites are regions of the solid catalyst, where free (unbalanced) chemical bonds exist allowing the adsorption process. In practice, active centers can be found in all distorted spots of the catalyst’s surface, such as cracks, peaks and corners. The higher the surface area of the catalyst the more active centers it may comprise. Active centers can be also easily damaged by a variety of factors [23].
According to Odenbrand et al. and Li et al. catalyst deactivation is a mainly chemical phenomenon [24,25]. The role of active component in the catalyst is to provide sufficient amount of active sites. For the V2O5−WO3/TiO2 catalyst, most important are V-OH acid sites and V5+ = O sites [15,26]. Vanadium oxide content is, therefore, responsible for catalyst activity, and its reduction during the SCR reactor’s operation results in activity loss. For the V2O5−WO3/TiO2 catalyst, the role of WO3 supporter is also considerable, as it interacts with TiO2 to enhance electron transfer as well as it helps with NO oxidation to NO3− and further decomposition of NO3− to NO2. This is beneficial for catalyst activity and to widen the temperature window for the reaction [27,28].
The main reactions occurring in SCR installation in the presence of a catalyst, are:
4NO + 4NH3 + O2 → 4N2 + 6H2O
2NO2 + 4NH3 + O2 → 3N2 + 6H2O
Ammonia reduction over the V2O5−WO3/TiO2 catalyst is highly selective, which means the NO is reduced before the O2. Therefore reactions (1) and (2) proceed at the highest rate. Side reactions occurring in the SCR reactors are the following [29]:
4NH3 + 3O2 → 2N2 + 6H2O
4NH3 + 4O2 → 2N2O + 6H2O
4NH3 + 5O2 → 4NO + 6H2O
4NH3 + 7O2 → 4NO2 + 6H2O
4NH3 + 4NO + 3O2 → 4N2O + 6H2O
Side reactions (4)–(7) are undesired, as they lead to formation of nitrogen oxides (N2O, NO, NO2) instead of elementary nitrogen. Most often it is possible to detect some N2O downstream the SCR reactor, and it is a strong greenhouse gas but it is hard to detect NO or NO2 in flue gas. However, increased consumption of reagent (particularly when SCR operating temperature exceeds 380 °C) can be an indication of their formation, according to Media et al. [29]. All the above equations indicate the key role of ammonia in the process [30]. To ensure the best performance of the SCR reactor, the proper NH3/NOx molar ratio is required. The highest NOx reduction efficiency can be achieved for NH3/NOx molar ratio between 1.0 and 1.1. For a value of NH3/NOx higher than this, reduction efficiency remains constant but the ammonia slip strongly increases [31]. To start SCR reactions, ammonia must first be activated, which can occur either through Lewis-type interaction or over a Brønsted acidic site, of which the second mechanism is more favorable [32]. According to adsorption theory, a desired reaction is boosted by a catalyst wherever ammonia molecules (in Lewis-type interactions) or ammonia ions (in Brønsted acidic site) can adsorb on the surface [33,34]. Also the presence of sulfates increase Lewis and Brønsted acidity, which is beneficial for NOx reduction capabilities as well as the catalyst lifetime [35,36]. Some authors refer to the catalyst storage capacity as the total amount of NH3 that can be adsorbed on the catalyst surface [37,38]. Another factor, important from the DeNOx reaction point of view, is a proper contact time between flue gas and the catalyst surface, therefore, too high a flue gas velocity can result in decreased NOx removal efficiency [31].
Referring to the basics of the catalysis process, some more interesting facts can be derived. According to literature, the idea of the catalytic process is to reduce energy of activation required for a given reaction. Without the catalyst (i.e., in the SNCR method), the temperature in DeNOx reactions is equal to about 950 °C [39,40]. For vanadium-based catalysts, maximum efficiency can be achieved in about 350 °C [41]. However, NO conversion decreases from 100% at about 270 ÷ 300 °C to only 20% at about 120 °C [27]. According to Song et al. [42], who tested a V2O5−MoO3/TiO2 catalyst, NO conversion can drop significantly also when temperature higher than 350 °C is reached.
In pulverized coal boilers, it is difficult to maintain constant flue gas temperature and distribution. The temperature at which flue gas enters the reactor depends on multiple boiler parameters, such as for instance a coal mills configuration, the amount of air supplied to combustion and to over-fire nozzles, the cleanliness level of heating surfaces, the load of the boiler or feed water inlet temperature. Volatile matter content in coal can also influence flue gas temperature. If there are some regions of the SCR reactor that are exposed to lower flue gas flow, it may lead to faster deactivation of those regions due too. On the other hand, many of the above parameters can be controlled during the boiler operation, which gives room for optimization of SCR working conditions.
Because the catalyst’s performance declines over time, it is important to measure its ability to reduce NOx from flue gas by means of an absolute factor. The performance of SCR catalysts, can be expressed as follows [23]:
X   =   1     C N O o u t + C N O 2 o u t C N O i n +   C N O 2 i n
Varea   =   V A
K   =   Varea   ln 1 X
where:
Varea—area velocity [m3/m2h]
A—total surface area of the catalyst [m2]
V—flue gas volumetric flow through the catalyst [m3/h]
X, X0—efficiency of the catalyst (used, fresh)
C—concentration of inlet and outlet NO/NO2
K, K0—catalyst activity coefficient (used, fresh)
By dividing the activity coefficient K of an used catalyst by K0 of a fresh catalyst, overall loss of activity can be calculated for the catalyst. As long as both the fresh and used catalysts are exposed to exactly the same volumetric flue gas flow and have exactly the same total surface area A (not the surface area measured by BET isotherm method), area velocity Varea (Equations (9) and (10)) can be neglected. When a set of catalyst layers is under consideration, K values for each layer should be added to obtain the activity coefficient for a whole installation. Equation (10) can only be applied when NH3/NOx ratio is between 1.00 and 1.02, while Equation (8) can be used for a general assessment of the catalyst’s performance [23].
Most desired improvements that are under research for vanadium-based catalyst are aimed to ensure its appropriate operation in the temperature range 200–250 °C [43]. According to Koebel et al. [44], rate of reaction at temperatures below 300 °C is limited by reoxidation of vanadium sites, which can be partially compensated by introducing NO2. However, almost all of NOx generated in combustion chamber is in the form of NO [45]. The way to improve a vanadium-based catalyst’s performance at lower temperatures is to take advantage of the so-called “fast SCR” reaction. By increasing the fraction of NO2 before the SCR reactor, which can be done by placing additional oxidation catalyst, the “fast SCR” reaction will occur [44,46]:
2NH3 + NO + NO2 → 2N2 + 3H2O
The activity of a vanadium-based catalyst at low temperature can be also improved by doping with elements, such as Co, Ce, Cr, Mo or Ni [47]. According to Li et al. [48], a small addition of silica to the TiO2 supporter can improve catalyst selectivity towards N2. Activity of the vanadium-based catalyst can be increased by the addition of CeO2, done by the impregnation or co-precipitation methods. This allows us to increase activity of the catalyst, widen the temperature window of the reaction, as well as improve resistance to sulfur [49,50].

3. Problem of Ammonia Bisulfate Formation

3.1. Ammonia Slip

In order to abate NOx emission by employing selective-catalytic reduction methods, the flue gas from a power plant is mixed with liquid ammonia (NH3) in the presence of catalyst. Ammonia is injected to the flue gas stream at the inlet of the SCR reactor by means of spray nozzles. Typically, not all of the injected ammonia reacts and a part of it is carried over by the flue gas stream and escapes the SCR reactor. This amount of NH3 is usually called the ammonia slip [51,52]. As already mentioned, ammonia slip appears when NH3/NOx ratio is higher than 1 [31]. There are several reasons why this phenomenon is highly undesired from the power plant operation point of view. Escaped ammonia would react with SO3 in flue gas, creating sticky ammonium bisulfates that can easily collect dust and plug not only SCR catalysts, but also devices such as rotating air preheaters. Besides the formation of sulfates, ammonia can also adsorb on the fly ash. In particular, adsorption occurs when there is amorphous structure in fly ash particles. Therefore, it is also linked with the combustion process and boiler load, as faster flue gas cooling leads to the presence of amorphous phase in the particles [53,54,55,56,57,58]. Escaped ammonia may thus affect the quality of fly ash from the power plant—the material that is typically sold on the market. A strong smell of ammonia may hinder fly ash from being utilized, which would generate additional costs for the power plant. It is worth mentioning that total ammonia adsorbed by fly ash depends not only on the level of ammonia slip from SCR reactors, but also on the fly ash structure and composition. An acceptable value of ammonia slip from SCR reactor is in the range of 1–2 ppmv while slip higher than 5 ppmv should be avoided [59]. A typical reason for ammonia slip that is too large are poor ammonia distribution in flue gas, catalyst age and contamination, low flue gas temperature, inlet NOx level higher than SCR reactor design point as well high flue gas flow.

3.2. Sulfur Trioxide

One of the crucial factors for both boiler and SCR catalysts’ operation is the concentration of sulfur trioxide (SO3). Its presence in the flue gas stream is derived from the SO2 to SO3 conversion. There are several factors determining this phenomenon. The most important is sulfur content in coal which is mostly converted to SO2 in the combustion chamber. Another is oxygen concentration—the more oxygen present in the combustion process, the higher the SO3 concentrations will be [60]. Most of the SO3 formation in the combustion chamber takes place at temperatures above 530 °C. A high content of alkaline/earth-alkaline compounds (CaO in particular) would result in lower SO3 concentration, due to formation of sulfates. These reactions are most likely to happen in the temperature range 300–800 °C. The SO2/SO3 conversion is, on the other hand, catalyzed by the presence of iron oxide (Fe2O3) in flue gas. This catalytic conversion is most intense at about 600–700 °C (it does vary depending on the Fe2O3 particle size) and it has the most significant effect on the final SO3 concentration in flue gas [60]. Also the presence of alumina, CuO and V2O5 results in a higher SO2/SO3 conversion rate. Typical conversion upstream of the SCR reactor in a pulverized coal boiler is within the range of 1–2% [55,61,62,63,64]. In a coal-fired boiler, all the aforementioned parameters are subjected to dynamic changes. Therefore, SO3 concentration at the outlet of the combustion chamber may vary significantly and cannot be robustly predicted without a dedicated measurement system [65].
Downstream of the combustion chamber, the concentration of SO3 in a flue gas further increases when passing the SCR reactor. Next to the ammonia slip, high SO2/SO3 conversion is another undesired parameter of the SCR installations. Depending on the catalyst wear and operating conditions, SO2/SO3 conversion rate can reach up to 2% for the V2O5-WO3/TiO2 catalyst. The conversion rate increases with temperature—at the temperature of 400 °C it can be more than twice the value obtained at 300 °C. Also, the SO2/SO3 conversion increases proportionally to NH3 concentration. According to Lu et al. [66], conversion can increase from 1.1% to 2.1% when NH3 concentration increases from 40 ppm to 140 ppm. Furthermore, the content of TiO2 and V2O5 can significantly affect the conversion. On the other hand, the SO2/SO3 conversion in a SCR reactor increases by roughly 30% in the absence of ammonia injection. This occurs because the lack of NH3 makes the catalyst pores more accessible for SO2 adsorption and, therefore, contact with TiO2 and V2O5 is increased [55,63,66,67,68]. According to the results published by Electric Power Research Institute [69], SO2/SO3 conversion rate increases also with catalyst age as well as for the regenerated catalysts. The effect of H2O on the SO2/SO3 conversion is not consistent. According to Qing et al. [70], a rise in conversion can be visible for H2O concentration in a flue gas between 7.5% and 15%, while for higher concentrations a visible drop can be noticed.
Sulfur trioxide is responsible for the formation of sulfuric acid in a flue gas from coal-fired power plants and it participates in ammonium bisulfate formation. The pure form of SO3 gas exists at the temperature higher than 427 °C. As the temperature drops along the boiler path, SO3 combines with water vapor forming H2SO4. This process ends when temperature reaches about 200 °C, when almost all SO3—that has not reacted with ammonia in the meantime—is converted to sulfuric acid [67,71]. The latter may condense at temperature below ca. 200 °C [72], which can lead to serious corrosion issues [71].

3.3. Ammonia Bisulfate Formation

Both ammonia slip and sulfur trioxide influence the formation of ammonium bisulfates (ABS), sticky and corrosive substances that block active sites on the catalyst [17,42,73,74]. According to Wang et al. [75], the deposition of 30% wt. NH4HSO4 on the catalyst results in a pore volume drop of more than 50% and a significant reduction in BET surface area. When ABS starts depositing, firstly the larger pores are affected and smaller pores remain untouched. If the temperature of flue gas is above the ABS dew point, condensation can still occur inside the catalyst pores. Crucial chemical reactions that describe ABS formation in a coal-fired boiler are [76]:
NH3 + SO3 +H2O → NH4HSO4
NH3 + H2SO4 → NH4HSO4
SO3 + H2O → H2SO4
The obvious way to abate issues with ABS is to avoid or decrease its formation. The latter is proportional to the concentration of SO3, moisture and H2SO4 in the flue gas, as well as ammonia slip from SCR reactors. The latter increases when the catalyst activity drops, which indicates the relation between ammonia slip and other mechanisms of deactivation, such as a decrease in Brønsted acid sites for instance. The higher the concentration of NH3, SO3 and moisture in flue gas, the higher the temperature of ABS condensation [76,77]. It is important to determine the accurate temperature of ABS condensation, because the temperature of flue gas varies significantly with the boiler load [61]. Knowing the edge temperature allows to operate the boiler in a manner that could prevent ABS condensation. However, there is no consistency in the literature about the temperature of ABS condensation. Furthermore, the temperature of NH4HSO4 decomposition on the catalyst surface can be higher due to the specific pore structure. If catalyst pore size decreases below 10 nm, the decomposition temperature increases due to the capillary forces [55,65]. According to Menasha et. al. [77], ABS can precipitate in the temperature range of 262–327 °C, with peak intensity in the range of 284–300 °C. Shi et al. [78] pointed out that the temperature of ABS condensation is in the range of 190–240 °C. It can be concluded that at the temperature below ca. 330 °C, there is a risk of ammonia bisulfate precipitation. Producers, knowing the actual pore structure, can define the minimum safe operating temperature for a catalyst, such as presented in Figure 1 [65]. Power plant operators may consider closing NH3 injection if the actual temperature of the flue gas is lower than the minimum safe temperature, as operating the SCR reactor below the ABS dew point can be seriously damaging for catalyst activity. Results of laboratory tests show that after 20 h of operation in such conditions, coefficient K/K0 is equal to ca. 0.6 for the first layer and 0.8 for the second layer of the SCR reactor [76,77]. Ammonium bisulfates is also dangerous for devices placed downstream the SCR reactor. In a coal-fired power plant, the rotating air preheater is typically affected due to increased fouling, which is a highly undesirable phenomenon [79].

4. Deactivation Mechanisms

Catalyst deactivation phenomena can be divided into three major types—mechanical, chemical and thermal deactivation [6]. As far as chemical deactivation is concerned, it is vital to discuss issues related with catalyst poisoning which, besides already discussed ammonium bisulfates formation, are crucial for the V2O5−WO3/TiO2 catalyst operation.

4.1. Poisoning

Poisoning of the SCR catalyst is mainly associated with alkali metals and sulfur. Also, water present in the flue gas can affect activity of the vanadium-based catalyst as it occupies acidic sites [80]. The concentration of water vapor equal to 8% in a flue gas can decrease activity by roughly 15% after 3 h of operation. If water is removed from the flue gas, however, activity is restored to the initial level [41]. At temperature above 450 °C, however, moisture can enhance NOx conversion [81]. The V2O5-WO3/TiO2 catalyst is well known for its resistance to SO2 poisoning. Despite that, such catalysts are still affected by sulfur dioxide content in the flue gas, particularly when the flue gas temperature drops below 300 °C [82]. Its presence allows compounds to form that are directly responsible for catalyst deactivation or plugging. The role of pure SO2 is also significant. According to Zang et al. [83], SO2 adsorption on the catalyst occurs and SO42− ions are formed as the result, particularly in the presence of oxygen, and they can occupy active sites of the catalyst. Xiao et al. [84] indicated, that when adding roughly 1000 ppm of pure SO2 to the flue gas stream, the activity of the catalyst can drop by ca. 20%. The major reason for V2O5−WO3/TiO2 catalyst deactivation is, however, related to interactions with alkali metals present in flue gas. As the temperature required by the V2O5−WO3/TiO2 catalyst is relatively high, the most common position of the SCR installation is upstream of the dust collector (called a high-dust SCR) [2,6,85]. Because of that, the catalyst surface interacts with alkali and earth alkali metals carried out by the flue gas stream. Along with the SO2, alkali metals can poison the catalyst and, in consequence, reduce its activity. One of the major reasons for that is the neutralization of acid sites. Sodium, calcium and potassium are known to have the most damaging effect on the vanadium-based catalysts [86,87,88,89,90,91]. Tang et al. [92] confirmed that cations Na+ and Ca2+ can affect the activity of the catalyst, due to neutralization of Brønsted acid sites as well as a decrease of vanadium reducibility. The CaSO4, responsible for plugging the catalyst as well as increased emission of N2O, is formed when Ca combines with SO2 and O2 [89,93,94,95]. Also the CaO and CaCO3 reduce catalyst activity [95]. Deng et al. studied the effect of NaCl, KCl, Na2CO3, K2CO3, Na2SO4 and K2SO4 on the catalyst’s performance. In the worst case scenario, the V2O5-WO3/TiO2 catalyst exposed to the stream of pure KCl and NaCl aerosol for 5 h lost, respectively, about 26% and 16% of its initial efficiency. At low K/V molar ratios (range 0.02–0.1), K2S2O7 is formed and it interacts with V2O5. The product of this interaction—eutectic V2O5-K2S2O7—decreases NH3 adsorption on Brønsted acid sites [25]. Zheng et al. [95] tested deactivation of a catalyst exposed to a pure KCl and K2SO4 stream and concluded that the main reason for deactivation was reaction between poison and Brønsted acid sites. The compounds, such as Na2CO3, K2CO3, Na2SO4 or K2SO4, as well as the presence of KCl and NaCl in coal, results in lower, but still clearly noticeable deactivation [6,87]. According to Xiao et al. [84], Na2O can decrease the number of V5+ = 0 bonds as well as the surface oxygen, which affect NOx conversion. Li et al. [41] compared the loss of V2O5−WO3/TiO2 catalyst activity due to poisoning by different alkali metals. Tests revealed that the most damaging effect on the catalyst was from NaOH, followed by Na2O, Al(OH)3 and Al2O3. Already 0.5% concentration of NaOH in flue gas is sufficient to decrease catalyst activity by 20% at 400 °C temperature and almost 40% at 350 °C. The presence of SO2, despite its negative effect in other situations, can abate deactivation caused by K, Na and Ca [2,96]. On top of that, Guo et al. [32] suggested, despite the negative effect of CaSO4, that sulfation can be actually beneficial for the catalyst activity as it increases the number of active Brønsted acid sites. According to Liu et al., the presence of Cl in flue gas can inhibit adsorption of NH3 and NOx on the catalyst, which would affect its activity as well [90,97]. Another element that is capable of decreasing catalyst activity is phosphorus [30,98]. Also, arsenic has been reported as a poisonous element for vanadium-based catalyst which, along with the toxicity of vanadium, is the reason why spent catalysts are treated as a hazardous waste [5,99,100].

4.2. Mechanical Deactivation

In coal-fired boilers, flue gas entering the SCR reactors contains a significant amount of particulate matter, such as fly ash particles, cenospheres or unburnt coal particles [57]. They hit the surface of catalysts at some point with a given momentum and can damage the surface, cause surface attrition or result in dust build-up in some areas of the SCR reactors [101]. Mechanical damage of the catalyst’s surface, dust accumulation as well as attrition will result in the decrease of activity, affecting the number of available active sites. In an ideal situation, flow through the catalyst’s plates should be optimized in such a way that the interaction between flue gas and SCR reactor surface has minimal impact on the catalyst’s performance. Besides that, the amount of ammonia injected per unit of flue gas volume should be uniform. In industrial reality, flue gas distribution and velocity are usually uneven, which can affect a catalyst’s performance [101,102,103]. In order to ensure a uniform distribution of flue gas entering the SCR reactor, installations are typically equipped with guide vanes directing the flow [104]. This solution might not be sufficient, however, as the flue gas distribution can pivot between left and right side of the boiler during its operation, so theoretically vanes should be able to dynamically adjust to the inlet flue gas distribution.

4.3. Thermal Deactivation

Thermal deactivation of the catalyst is related to TiO2 phase transition (or sintering). The TiO2 most commonly occurs as one of two polymorphs—rutile or anatase. The rutile form is stable, while the anatase is stable only below certain temperature and can undergo irreversible phase transition to rutile. This process is influenced by the temperature at first place, but also particle size and shape, surface area, excess of oxygen or presence of promoters and inhibitors. Temperature of the transition can vary from 400 °C to 1200 °C, although it is mostly expected at roughly 600–750 °C. Elements, such as K, Na, Ni, Li, Fe, Zn, Cu, Cd, Mn and Cr, may act as promoters while elements, such as Ca, Si, P, S, Zr, Al, Ba, B and Ce may act as inhibitors [105,106,107,108,109]. Du et al. [110] pointed out that if the structure of TiO2 is hollow, it could be possible to avoid phase transition at temperatures below 800 °C. Also, an amorphous carbon coating can prevent transition, by hindering the access of oxygen. One of the roles of tungsten in the V2O5−WO3/TiO2 catalyst is to reduce phase transition of TiO2 at high temperature [111,112]. However, according to Khutan et al. [113], temperature of the transition can be decreased by vanadium doping, which might be unfavorable for SCR catalysts containing both anatase TiO2 and V2O5.
Thermal deactivation can result in overall surface area decrease and changes in surface structure resulting in a larger share of NH3 adsorbed on Lewis-type sites. High temperature can also provoke vanadium volatilization [14,114,115,116]. As a result, the activity of the catalyst is affected as well as the temperature window for DeNOx reaction becoming narrower. Besides that, the SO2/SO3 conversion increases, which can make the catalyst eventually unusable [117].

5. Industrial Boilers Characteristics

SCR reactors equipped with a V2O5-WO3/TiO2 catalyst are widely used in the power production sector. A pulverized coal boiler is a good example of SCR application unit with its rather difficult and rapidly changing operation conditions. According to Ma et al. [118], NOx specific emission from coal-fired units is lower for high capacity units, as well as for units burning bituminous coal. In typical configuration, the standard SCR reactor is placed at the end of the boiler, upstream of the air preheater. It must be designed specifically for each unit in order to locate reactors in the best temperature conditions. There are plenty of issues related to such configuration.
In the coal-fired power plant, temperature variation is a common problem. In the combustion chamber temperature can be as high as 1700–1900 °C and it decreases along the flue gas path [57]. The SCR reactors are typically installed close to the boiler outlet or the economizer (water heater). The temperature of flue gas entering the reactor depends on a variety of factors. The first is the combustion process itself, where the temperature depends mainly on the amount of air (oxygen) in the process as well as volatile matter content in coal [119]. The temperature of flue gas leaving the combustion chamber will depend on the firing system configuration and whether there is air staging or auxiliary air supply. Afterwards, flue gases are cooled down by means of heat exchangers (superheaters, economizers) along their pathway. If heating surfaces in the boiler are contaminated, temperature of the flue gas downstream of a given heat exchanger will be higher due to reduced heat transfer. Temperature in the combustion chamber may also fluctuate rapidly due to the difference in fuel properties, as coal fed to the boiler does not have a consistent composition most of the time. Any temperature fluctuation in the combustion chamber is going to propagate and probably affect the temperature of flue gas before the SCR reactors. Another factor that may influence this parameter is steam—its flow and temperature in the boiler, as well as cooling water injection level. Finally, the load changes. Depending on the boiler construction, the load level can decrease to about 40%. In older constructions equipped with a drum, this is usually 60%. The decreased load also results in lower flue gas temperature entering the SCR reactor [61]. Figure 2 presents a correlation between flue gas temperature, measured in the second pass of the boiler, and the load level for a medium-size, 120 kg/s nominal steam capacity, pulverized coal-fired boiler (hard coal). As can be seen, maximum flue gas temperature variation is equal to about 41 °C. Taking this into account, it can be concluded that the operation of the SCR reactor within optimal temperature range in coal-fired boilers can be difficult to achieve.
Moreover, in coal-fired boilers it is not easy to maintain an even flue gas distribution in the cross‑section of the duct. It is a common knowledge that most of the time, the parameters of the flue gas in the boiler are offset to the right or left side, and hardly ever are they distributed evenly. This depends on the burners’ configuration (tangential or front wall), as well as appropriate pulverized coal and air distribution in each of the burner and auxiliary air nozzles. The flue gas distribution can also depend on the cleanliness of the boiler heating surfaces. For instance, an intense dust built up on the one side of superheater causes additional pressure drop and thus disturbs the flue gas distribution downstream of this superheater [120]. It is also difficult to control the flue gas distribution during boiler operation and it is often impossible to obtain real-time information about this distribution. Some estimations can be made, however, by taking into account the temperature measurement along the boiler or ID fans’ power consumption. CFD (Computational Fluid Dynamics) modelling is often employed to get knowledge about the real flue gas distribution, but it would be rather difficult to have such model working on-line in the power plant (although technically possible).

6. Conclusions

In the article, deactivation factors of vanadium-based catalyst for the DeNOx process have been comprehensively reviewed. They can be divided into three major categories: chemical, mechanical and thermal deactivation. As suggested by literature, the main problems influencing the operation of SCR reactors are the formation of SO3 and ammonia slip. Both have a crucial effect on the formation of ammonium bisulfates, which is one of the most undesired side-products of DeNOx processes in coal-fired power plants. Ammonia bisulfate is a sticky substance, capable of clogging the surface of the catalyst as well as other equipment downstream of the SCR reactor. It is also responsible for a vast share of chemical deactivation of the catalyst, next to the poisoning arising from the presence of alkali and earth alkali metals in flue gas. The role of sulfur is ambiguous. On the one hand, it contributes to the ammonia bisulfate formation, but on the other hand, sulfation can increase the number of active Brønsted acid sites which is beneficial for catalyst activity. The latter can be also decreased by operating the SCR reactor at too high or low temperatures. Above 400 °C there is a risk of TiO2 phase transition, while below 300 °C activity drops due to reaction kinetics as well as the effect of sulfur, alkali metals and ammonium bisulfates. A commercial SCR reactor installed on a pulverized coal boiler, is subjected to rapid and vast flue gas temperature fluctuations. Also, uneven flue gas distribution can result in a significant temperature difference along the reactor cross-section. Besides that, uneven flue gas distribution is responsible for faster deactivation of some parts of the catalysts due to attrition or dust build-up.

Author Contributions

Conceptualization, M.M. and B.S.; methodology, M.Z.; formal analysis, M.Z.; investigation, M.Z.; resources, M.Z.; data curation, M.Z.; writing—original draft preparation, M.Z. and B.S.; writing—review and editing, B.S and K.S.; visualization, M.Z.; supervision, M.M. and K.S.; project administration, M.M.; funding acquisition, M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was founded by the grant of the Ministry of Science and Higher Education no. 0039/DW/2018/02.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. UE. “Directive Implementation UE 2017/1442 from 31st of July 2017,” [Online]. Available online: http://data.europa.eu/eli/dec_impl/2017/1442/oj. (accessed on 23 September 2020).
  2. Kiełtyka, M.; Dias, A.P.S.; Kubiczek, H.; Sarapata, B.; Grzybek, T. The influence of poisoning on the deactivation of DeNOx catalysts. C. R. Chim. 2015, 18, 1036–1048. [Google Scholar] [CrossRef]
  3. Zheng, Y.; Jensen, A.D.; Johnsson, J.E. Deactivation of V2O5-WO3-TiO2 SCR catalyst at a biomass-fired combined heat and power plant. Appl. Catal. B Environ. 2005, 60, 253–264. [Google Scholar] [CrossRef] [Green Version]
  4. Yu, Y.; Chi, H.E.; Chen, J.; Meng, X. Deactivation mechanism of de-NOx catalyst (V2O5-WO3/TiO2) used in coal fired power plant. J. Fuel Chem. Technol. 2012, 40, 1359–1365. [Google Scholar] [CrossRef]
  5. Li, M.; Liu, B.; Wang, X.; Yu, X.; Zheng, S.; Du, H.; Dreisinger, D.; Zhang, Y. A promising approach to recover a spent SCR catalyst: Deactivation by arsenic and alkaline metals and catalyst regeneration. Chem. Eng. J. 2018, 342, 1–8. [Google Scholar] [CrossRef]
  6. Deng, L.; Liu, X.; Cao, P.; Zhao, Y.; Du, Y.; Wang, C.; Che, D. A study on deactivation of V2O5eWO3eTiO2 SCR catalyst by alkali metals during entrained-flow combustion. J. Energy Inst. 2017, 90, 743–751. [Google Scholar] [CrossRef]
  7. Tomasic, V. Application of the monoliths in DeNOx catalysis. Catal. Today 2007, 119, 106–113. [Google Scholar] [CrossRef]
  8. Soleimanzadeh, H.; Niaei, A.; Salari, D.; Tarjomannejad, A.; Penner, S.; Grunbacher, M.; Hosseini, S.A.; Mousavi, S.M. Modeling and optimization of V2O5/TiO2 nanocatalysts for NH3 -Selective catalytic reduction (SCR) of NOx by RSM and ANN techniques. J. Environ. Manag. 2019, 238, 360–367. [Google Scholar] [CrossRef]
  9. Rauch, D.; Albrecht, G.; Kubinski, D.; Moos, R. A microvave-based method to monitor the ammonia loading of a vanadia-based SCR catalyst. Appl. Catal. B Environ. 2015, 165, 36–42. [Google Scholar] [CrossRef]
  10. Song, I.; Youn, S.; Lee, H.; Lee, S.G.; Cho, S.J.; Kim, D.H. Effects of microporous TiO2 support on the catalytic and structural properties of V2O5/microporous TiO2 for the selevtive catalytic reduction of NO by NH3. Appl. Catal. B Environ. 2017, 210, 421–431. [Google Scholar] [CrossRef]
  11. He, Y.; Ford, M.E.; Zhu, M.; Liu, Q.; Tumuluri, U.; Wu, Z.; Wachs, I.E. Influence of catalyst synthesis method on selective catalytic reduction (SCR) of NO by NH3 with V2O5-WO3/TiO2 catalysts. Appl. Catal. B Environ. 2016, 193, 141–150. [Google Scholar] [CrossRef] [Green Version]
  12. Yates, M.; Martin, J.A.; Martin-Luengo, M.A.; Suarez, S.; Blanco, J. N2O formation in the ammonia oxidation and in the SCR process with V2O5-WO3 catalysts. Catal. Today 2005, 107, 120–125. [Google Scholar] [CrossRef]
  13. Huang, Y.; Tong, Z.; Wu, B.; Zhang, J. Low temperature selective catalytic reduction of NO by ammonia over VxO5-CeO2/TiO2. J. Fuel Chem. Technol. 2008, 36, 616–620. [Google Scholar] [CrossRef]
  14. Zhou, X.; Huang, X.; Xie, A.; Lou, S.; Yao, C.; Li, X.; Zuo, S. V2O5-decorated Mn-Fe/attapulgite catalyst with high SO2 tolerance for SCR of NOx with NH3 at low temperature. Chem. Eng. J. 2017, 326, 1074–1085. [Google Scholar] [CrossRef]
  15. Nicosia, D.; Czekaj, I.; Krocher, O. Chemical deactivation of V2O5/WO3–TiO2 SCR catalysts by additives and impurities from fuels, lubrication oils and urea solution Part II. Characterization study of the effect of alkali and alkaline earth metals. Appl. Catal. B Environ. 2008, 77, 228–236. [Google Scholar] [CrossRef]
  16. Wan, Q.; Duan, L.; Li, J.; Chen, L.; He, K.; Hao, J. Deactivation performance and mechanism of alkali (earth) metals on V2O5–WO3/TiO2 catalyst for oxidation of gaseous elemental mercury in simulated coal-fired flue gas. Catal. Today 2011, 175, 189–195. [Google Scholar] [CrossRef]
  17. Youn, S.; Song, I.; Lee, H.; Cho, S.J.; Kim, D.H. Effect of pore structure of TiO2 on the SO2 poisoning over V2O5/TiO2 catalysts for selective catalytic reduction of NOx with NH3. Catal. Today 2018, 303, 19–24. [Google Scholar] [CrossRef]
  18. Hammershoi, P.S.; Jangjou, Y.; Epling, W.S.; Jensen, A.D.; Janssens, T.V. Reversible and irreversible deactivation of Cu-CHA NH3-SCRcatalysts by SO2 and SO3. Appl. Catal. B Environ. 2018, 226, 38–45. [Google Scholar] [CrossRef] [Green Version]
  19. Bai, S.; Wang, Z.; Li, H.; Xu, X.; Liu, M. SO2 promotion in NH3-SCR reaction over V2O5/SiC catalyst at low temperature. Fuel 2017, 194, 36–41. [Google Scholar] [CrossRef]
  20. Samojeden, B.; Grzybek, T. The influence of the promotion of N-midified activated carbon with iron on NO removal by NH3-SCR (Selective catalytic reduction). Energy 2016, 116, 1484–1491. [Google Scholar] [CrossRef]
  21. Zhang, P.; Chen, T.; Zou, X.; Zhu, C.; Chen, C.; Liu, H. V2O5/hematite catalyst for low temperature selective catalytic reduction of NOx with NH3. Chin. J. Catal. 2014, 35, 99–107. [Google Scholar] [CrossRef]
  22. Kowalczyk, A.; Święs, A.; Gil, B.; Rutkowska, M.; Piwowarska, Z.; Borcuch, A.; Michalik, M.; Chmielarz, L. Effective catalysts for the low-temperature NH3-SCR process based on MCM-41 modified with copper by template ion-exchange (TIE) method. Appl. Catal. B Environ. 2018, 237, 927–937. [Google Scholar] [CrossRef]
  23. EPRI. Laboratory Testing Protocol for Coal-Fired SCR Catalyst, 3rd ed.; Report no. 3002013048; Electric Power Research Institute: Palo Alto, CA, USA, 2018. [Google Scholar]
  24. Odenbrand, C. CaSO4 deactivated V2O5-WO3/TiO2 SCR catalyst for a diesel power plant. Characterization and simulation of the kinetics of the SCR reaction. Appl. Catal. B Environ. 2018, 234, 365–377. [Google Scholar] [CrossRef]
  25. Li, Q.; Chen, S.; Liu, Z.; Liu, Q. Combined effect of KCl and SO2 on the selective catalytic reduction of NO by NH3 over V2O5/TiO2 catalyst. Appl. Catal. B Environ. 2015, 164, 475–482. [Google Scholar] [CrossRef]
  26. Wang, J.; Miao, J.; Yu, W.; Chen, Y.; Chen, J. Study on the local difference of monolithic honeycomb V2O5-WO3/TiO2 denitration catalyst. Mater. Chem. Phys. 2017, 198, 193–199. [Google Scholar] [CrossRef]
  27. Zhang, S.; Zhong, Q. Surface characterization studies on the interaction of V2O5-WO3/TiO2 catalyst for low temperature SCR of NO with NH3. J. Solid State Chem. 2015, 221, 49–56. [Google Scholar] [CrossRef]
  28. Zhang, S.; Zhong, Q. Promotional effect of WO3 on O2 over V2O5/TiO2 catalyst for selective catalytic reduction of NO with NH3. J. Mol. Catal. A Chem. 2013, 373, 108–113. [Google Scholar] [CrossRef]
  29. Madia, G.; Koebel, M.; Elsener, M.; Wokaun, A. Side Reactions in the Selective Catalytic Reduction of NOx with Various NO2 Fractions. Ind. Eng. Chem. Res. 2002, 41, 4008–4015. [Google Scholar] [CrossRef]
  30. Castellino, F.; Rasmussen, S.B.; Jansen, A.D.; Johnsson, J.E.; Fehrmann, R. Deactivation of vanadia-based commercial SCR catalysts by polyphosphoric acids. Appl. Catal. B Environ. 2008, 83, 110–122. [Google Scholar] [CrossRef]
  31. Gao, Y.; Liu, Q.; Bian, L. Numerical Simulation and Optimization of Flow Field in the SCR Denitrification System on a 600 MW Capacity Units. Energy Procedia 2012, 14, 370–375. [Google Scholar] [CrossRef] [Green Version]
  32. Guo, X.; Bartholomew, C.; Hecker, W.; Baxter, L. Effects of sulfate species on V2O5/TiO2 SCR catalysts in coal and biomass-fired systems. Appl. Catal. B Environ. 2009, 92, 30–40. [Google Scholar] [CrossRef]
  33. Soyer, S.; Uzun, A.; Senkan, S.; Onal, I. A quantum chemical study of nitric oxide reduction by ammonia (SCR reaction) on V2O5 catalyst surface. Catal. Today 2006, 118, 268–278. [Google Scholar] [CrossRef]
  34. Yao, H.; Chen, Y.; Wei, Y.; Zhao, Z.; Liu, Z.; Xu, C. A periodic DFT study of ammonia adsorption on the V2O5 (001), V2O5 (010) and V2O5 (100) surfaces: Lewis versus Brönsted acid sites. Surf. Sci. 2012, 606, 1739–1748. [Google Scholar] [CrossRef]
  35. Orsenigo, C.; Beretta, A.; Forzatti, P.; Svachula, J.; Tronconi, E.; Bregani, F.; Baldacci, A. Theoretical and experimental study of the interaction between NOx reduction and SO2 oxidation over DeNOx-SCR catalysts. Catal. Today 1996, 27, 15–21. [Google Scholar] [CrossRef]
  36. Boudali, L.; Ghorbel, A.; Grange, P. Characterization and reactivity of WO3eV2O5 supported on sulfated titanium pillared clay catalysts for the SCR-NO reaction. C. R. Chim. 2009, 12, 779–786. [Google Scholar] [CrossRef]
  37. Querel, C.; Bonfils, A.; Grondin, O.; Creff, Y. Control of a SCR system using a virtual NOx sensor. In Proceedings of the 7th IFAC Symposium on Advances in Automotive Control, Tokyo, Japan, 4–7 September 2013; pp. 9–14. [Google Scholar]
  38. Bonfilis, A.; Creff, Y.; Lepreux, O.; Petit, N. Closed-loop control of a SCR system using a NOx sensor cross-sensitive to NH3. In Proceedings of the 8th IFAC Symposium on Advanced Control of Chemical Processes, Singapore, 10–13 July 2012; pp. 738–743. [Google Scholar]
  39. Kang, Z.; Qixin, Y.; Lizheng, Z.; YuKun, D.; Baomin, S.; Wang, T. Study of the performance, simplification and characteristics of SNCR de-NOx in large-scale cyclone separator. Appl. Therm. Eng. 2017, 123, 635–645. [Google Scholar] [CrossRef]
  40. Daood, S.S.; Javed, M.T.; Gibbs, B.M.; Nimmo, W. NOx control in coal combustion by combining biomass co-firing, oxygen enrichment and SNCR. Fuel 2013, 105, 283–292. [Google Scholar] [CrossRef]
  41. Li, S.; Huang, W.; Xu, H.; Chen, T.; Ke, Y.; Qu, Z.; Yan, N. Alkali-induced deactivation mechanism of V2O5-WO3/TiO2 catalyst during selective catalytic reduction of NO by NH3 in aluminum hydrate calcining flue gas. Appl. Catal. B Environ. 2020, 270, 118872. [Google Scholar] [CrossRef]
  42. Song, L.; Chao, J.; Fang, Y.; He, H.; Li, J.; Qiu, W.; Zhang, G. Promotion of ceria for decomposition of ammonia bisulfate over V2O5-MoO3/TiO2 catalyst for selective catalytic reduction. Cham. Eng. J. 2016, 303, 275–281. [Google Scholar] [CrossRef]
  43. Putluru, S.; Schill, L.; Godiksen, A.; Poreddy, R.; Mossin, S.; Jensen, A.; Fehrmann, R. Promoted V2O5/TiO2 catalysts for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2016, 183, 282–290. [Google Scholar] [CrossRef] [Green Version]
  44. Koebel, M.; Madia, G.; Raimondi, F.; Wokaun, A. Enhanced Reoxidation of Vanadia by NO2 in the Fast SCR Reaction. J. Catal. 2002, 209, 159–165. [Google Scholar] [CrossRef]
  45. U.S. Environmental Protection Agency, “Nitrogen Oxides (NOx), Why and How They Are Controlled,” 1999. [Online]. Available online: http://www.epa.gov/ttn/catc. (accessed on 5 September 2020).
  46. Salazar, M.; Hoffmann, S.; Singer, V.; Becker, R.; Grunert, W. Hybrid catalysts for the selective catalytic reduction (SCR) of NO by NH3. On the role of fast SCR in the reaction network. Appl. Catal. B Environ. 2016, 199, 433–438. [Google Scholar] [CrossRef]
  47. Li, Q.; Hou, X.; Yang, H.; Ma, Z.; Zheng, J.; Liu, F.; Zhang, X.; Yuan, Z. Promotional effect of CeOX for NO reduction over V2O5/TiO2-carbon nanotube composites. J. Mol. Catal. A: Chem. 2012, 356, 121–127. [Google Scholar] [CrossRef]
  48. Li, F.; Shen, B.; Tian, L.; Li, G.; He, C. Enhancement of SCR activity and mechanical stability on cordierite supported V2O5-WO3/TiO2 catalyst by substrate acid pretreatment and addition of silica. Powder Technol. 2016, 297, 384–391. [Google Scholar] [CrossRef]
  49. Chen, M.; Zhao, M.; Tang, F.; Ruan, L.; Yang, H.; Li, N. Effect of Ce doping into V2O5-WO3/TiO2 catalysts on the selective catalytic reduction of NOx by NH3. J. Rare Earths 2017, 35, 1206–1215. [Google Scholar] [CrossRef]
  50. Shen, M.; Xu, L.; Wang, J.; Li, C.; Wang, W.; Wang, J.; Zhai, Y. Effect of synthesis methods on activity of V2O5/CeO2/WO3-TiO2 catalyst for selective catalytic reduction of NOx with NH3. J. Rare Earths 2016, 34, 259–267. [Google Scholar] [CrossRef]
  51. Shwan, S.; Partridge, W.; Choi, J.; Olsson, L. Kinetic modeling of NOx storage and reduction using spatially resolved MS measurements. Appl. Catal. B Environ. 2014, 147, 1028–1041. [Google Scholar] [CrossRef]
  52. Dolanc, G.; Strmcnik, S.; Petrovcic, J. NOx selective catalytic reduction control based on simple models. J. Process Control 2001, 11, 35–51. [Google Scholar] [CrossRef]
  53. Chen, C.; Cao, Y.; Liu, S.; Chen, J.; Jia, W. SCR catalyst doped with copper for synergistic removal of slip ammonia and elemental mercury. Fuel Process. Technol. 2018, 181, 268–278. [Google Scholar] [CrossRef]
  54. Krishnan, A.T.; Boehman, A.L. Selective catalytic reduction of nitric oxide with ammonia at low temperatures. Appl. Catal. B Environ. 1998, 18, 189–198. [Google Scholar] [CrossRef]
  55. Zheng, C.; Wang, Y.; Liu, Y.; Yang, Z.; Qu, R.; Ye, D.; Liang, C.; Liu, S.; Gao, X. Formation, transformation, measurement, and control of SO3 in coal-fired power plants. Fuel 2019, 241, 327–346. [Google Scholar] [CrossRef]
  56. Zhao, Y.; Yang, W.; Zhou, J.; Wang, Z.; Liu, J.; Cen, K. Experimental study on ammonia adsorption by coal ashes. J. Fuel Chem. Technol. 2015, 43, 266–272. [Google Scholar] [CrossRef]
  57. Zyrkowski, M.; Neto, R.C.; Santos, L.; Witkowski, K. Characterization of fly-ash cenospheres from coal-fired power plant unit. Fuel 2016, 174, 49–53. [Google Scholar] [CrossRef]
  58. EPRI. State-of-Knowledge of Fate, Removal, and Impact of Ammonia in Ash; Report no. 1004700; Electric Power Research Institute: Palo Alto, CA, USA, 2002. [Google Scholar]
  59. EPRI. Investigation of Ammonia Adsorption on Fly Ash and Potential Impacts of Ammoniated Ash; Report no. TR-113777; Electric Power Research Institute: Palo Alto, CA, USA, 1999. [Google Scholar]
  60. Wang, X.; Zhang, J.; Wang, Z.; Wang, Y.; Vujanovic, M.; Li, P.; Tan, H. Experimental and kinetics study on SO3 catalytic formation by Fe2O3 in oxy-combustion. J. Environ. Manag. 2019, 236, 420–427. [Google Scholar] [CrossRef]
  61. Zyrkowski, M.; Zymelka, P. Modelling of flexible boiler operation in coal fired power plant. IOP Conf. Ser. Earth Environ. Sci. 2019, 214, 012074. [Google Scholar] [CrossRef]
  62. Sporl, R.; Maier, J.; Scheffknecht, G. Sulphur Oxide Emissions from Dust-Fired Oxy-Fuel Combustion of Coal. Energy Procedia 2013, 37, 1435–1447. [Google Scholar] [CrossRef] [Green Version]
  63. Dunn, J.P.; Koppula, P.R.; Stenger, H.G.; Wachs, I.E. Oxidation of sulfur dioxide to sulfur trioxide over supported vanadia vatalysts. Appl. Catal. B Environ. 1998, 19, 103–117. [Google Scholar] [CrossRef]
  64. EPRI. Modeling of SO3 Formation Process in Coal-Fired Boilers; Report no. 1012689; Electric Power Research Institute: Palo Alto, CA, USA, 2007. [Google Scholar]
  65. Charles, A. Lockert, Breen Energy Solution., “Dynamic Control of SCR Minimum Operating Temperature,” [Online]. Available online: https://breenes.com/. (accessed on 6 September 2020).
  66. Lu, J.; Zhou, Z.; Zhang, H.; Yang, Z. Influenced factors study and evaluation for SO2/SO3 conversion rate in SCR process. Fuel 2019, 245, 528–533. [Google Scholar] [CrossRef]
  67. Breen Energy Solutions, “Comprehensive acid gas management,” [Online]. Available online: http://breenes.com/wp-content/uploads/2017/04/Comprehensive_Condensable_Management_Program-RevA.pdf. (accessed on 29 July 2019).
  68. Forzatti, P.; Nova, I.; Beretta, A. Catalytic properties in deNOx and SO2–SO3 reactions. Catal. Today 2000, 56, 431–441. [Google Scholar] [CrossRef]
  69. EPRI. Formation of N2O and NO2 Across Conventional DeNOx SCR Catalysts; Report no. 1019614; Electric Power Research Institute: Palo Alto, CA, USA, 2009. [Google Scholar]
  70. Qing, M.; Su, S.; Wang, L.; Liu, L.; Sun, Z.; Mostafa, M.; Xu, K.; Hu, S.; Wang, Y.; Xiang, J. Effects of H2O and CO2 on the catalytic oxidation property of V/W/Ti catalysts for SO3 generation. Fuel 2019, 237, 545–554. [Google Scholar] [CrossRef]
  71. Li, Y.; Zhu, Q.; Yi, Q.; Zuo, W.; Feng, Y.; Chen, S.; Dong, Y. Experimental method for observing the fate of SO3/H2SO4 in a temperature decreasing flue gas flow: Creation of state diagram. Fuel 2019, 249, 449–456. [Google Scholar] [CrossRef]
  72. Li, Z.; Sun, F.; Shi, Y.; Li, F.; Ma, L. Experimental study and mechanism analysis on low temperature corrosion of coal fired boiler heating surface. Appl. Therm. Eng. 2015, 80, 355–361. [Google Scholar] [CrossRef]
  73. Qi, G.; Yang, R.T. Performance and kinetics study for low-temperature SCR of NO with NH3 over MnOx-CeO2 catalyst. J. Catal. 2003, 217, 434–441. [Google Scholar] [CrossRef]
  74. Hammershoi, P.S.; Vennestrom, P.N.; Falsig, H.; Jansen, A.D.; Janssens, T.V. Importance of the Cu oxidation state for the SO2 poisoning of a Cu-SAPO-34 catalyst in the NH3-SCR reaction. Appl. Catal. B Environ. 2018, 236, 377–383. [Google Scholar] [CrossRef]
  75. Wang, X.; Du, X.; Zhang, L.; Chen, Y.; Yang, G.; Ran, J. Promotion of NH4HSO4 decomposition in NO/NO2 contained atmosphere at low temperature over V2O5-WO3/TiO2 catalyst for NO reduction. Appl. Catal. A Gen. 2018, 559, 112–121. [Google Scholar] [CrossRef]
  76. Muzio, L.; Bogseth, S.; Himes, R.; Chien, Y.; Dunn-Rankin, D. Ammonium bisulfate formation and reduced load SCR operation. Fuel 2017, 206, 180–189. [Google Scholar] [CrossRef]
  77. Menasha, J.; Dunn-Rankin, D.; Muzio, L.; Stallings, J. Ammonium bisulfate formation temperature in a bench-scale single-channel air preheater. Fuel 2011, 90, 2445–2453. [Google Scholar] [CrossRef]
  78. Shi, Y.; Shu, H.; Zhang, Y.; Fan, H.; Zhang, Y.; Yang, L. Formation and decomposition of NH4HSO4 during selective catalytic reduction of NO with NH3 over V2O5-WO3/TiO2 catalysts. Fuel Process. Technol. 2016, 150, 141–147. [Google Scholar] [CrossRef]
  79. Si, F.; Romero, C.; Yao, Z.; Schuster, E.; Xu, Z.; Morey, R.; Liebowitz, B. Optimization of coal-fired boiler SCRs based on modified support vector machine models and genetic algorithms. Fuel 2009, 88, 806–816. [Google Scholar] [CrossRef]
  80. Ning, R.; Liu, X.; Zhu, T. Research progress of low-temperature SCR danitration catalysts. Chin. J. Process Eng. 2019, 19, 223–234. [Google Scholar]
  81. Lee, S.M.; Kim, S.S.; Hong, S.C. Systematic mechanism study of the high temperature SCR of NOX by NH3 over a W/TiO2 catalyst. Chem. Eng. Sci. 2012, 79, 177–185. [Google Scholar]
  82. Wang, F.; Shen, B.; Zhu, S.; Wang, Z. Promotion of Fe and Co doped Mn-Ce/TiO2 catalysts for low temperature NH3-SCR with SO2 tolerance. Fuel 2019, 249, 54–60. [Google Scholar] [CrossRef]
  83. Zang, S.; Zhang, G.; Qiu, W.; Song, L.; Zhang, R.; He, H. Resistance to SO2 poisoning of V2O5/TiO2-PILC catalyst for the selective catalytic reduction of NO by NH3. Chin. J. Catal. 2016, 37, 888–897. [Google Scholar] [CrossRef]
  84. Xiao, H.; Chen, Y.; Qi, C.; Ru, Y. Effect of Na poisoning catalyst (V2O5-WO3/TiO2) on denitration process and SO3 formation. Appl. Surf. Sci. 2018, 433, 341–348. [Google Scholar] [CrossRef]
  85. Yan, Z.; Shi, X.; Yu, Y.; He, H. Alkali resistance promotion of Ce-doped vanadium-titanic-based NH3-SCR catalysts. J. Environ. Sci. 2018, 73, 155–161. [Google Scholar] [CrossRef]
  86. Tian, Y.; Yang, J.; Yang, C.; Lin, F.; Hu, G.; Kong, M.; Liu, Q. Comparative study of the poisoning effect of NaCl and Na2O on selective catalytic reduction of NO with NH3 over V2O5-WO3/TiO2 catalyst. J. Energy Inst. 2019, 92, 1045–1052. [Google Scholar] [CrossRef]
  87. Kong, M.; Liu, Q.; Zhu, B.; Yang, J.; Li, L.; Zhou, Q.; Ren, S. Synergy of KCl and Hgel on selective catalytic reduction of NO with NH3 over V2O5–WO3/TiO2 catalysts. Chem. Eng. J. 2015, 264, 815–823. [Google Scholar] [CrossRef]
  88. Yu, W.; Wu, X.; Si, Z.; Weng, D. Influences of impregnation procedure on the SCR activity and alkali resistance of V2O5-WO3/TiO2 catalyst. Appl. Surf. Sci. 2013, 283, 209–214. [Google Scholar] [CrossRef]
  89. Li, X.; Liu, C.; Li, X.; Peng, Y.; Li, J. A neutral and coordination regeneration method of Ca-poisoned V2O5-WO3/TiO2 SCR catalyst. Catal. Commun. 2017, 100, 112–116. [Google Scholar] [CrossRef]
  90. Lisi, L.; Lasorella, G.; Malloggi, S.; Russo, G. Single and combined deactivating effect of alkali metals and HCl on commercial SCR catalysts. Appl. Catal. B Environ. 2004, 50, 251–258. [Google Scholar] [CrossRef]
  91. Klimczak, M.; Kern, P.; Heinzelmann, T.; Lucas, M.; Claus, P. High-throughput study of the effects of inorganic additives and poisons on NH3-SCR catalysts—Part I: V2O5–WO3/TiO2 catalysts. Appl. Catal. B Environ. 2010, 95, 39–47. [Google Scholar] [CrossRef]
  92. Tang, F.; Xu, B.; Shi, H.; Qiu, J.; Fan, Y. The poisoning effect of Na+ and Ca2+ ions doped on the V2O5/TiO2 catalysts for selective catalytic reduction of NO by NH3. Appl. Catal. B Environ. 2010, 94, 71–76. [Google Scholar] [CrossRef]
  93. Chen, L.; Li, J.; Ge, M. The poisoning effect of alkali metals doping over nano V2O5–WO3/TiO2 catalysts on selective catalytic reduction of NOx by NH3. Chem. Eng. J. 2011, 170, 531–537. [Google Scholar] [CrossRef]
  94. Li, X.; Li, X.; Yang, R.T.; Mo, J.; Li, J.; Hao, J. The poisoning effect of calcium on V2O5-WO3/TiO2 catalyst for the SCR reaction: Comparison of different forms of calcium. Mol. Catal. 2017, 434, 16–24. [Google Scholar] [CrossRef]
  95. Zheng, Y.; Jensen, A.D.; Johnsson, J.E.; Thogersen, J.R. Deactivation of V2O5-WO3-TiO2 SCR catalyst at biomass fired power plants: Elucidation of mechanisms by lab- and pilot-scale experiments. Appl. Catal. B Environ. 2008, 83, 186–194. [Google Scholar] [CrossRef]
  96. Dahlin, S.; Nilsson, M.; Backstrom, D.; Bergman, S.L.; Bengtsson, E.; Bernasek, S.L.; Pettersson, L.J. Multivariate analysis of the effect of biodiesel-derived contaminants on V2O5-WO3/TiO2 SCR catalysts. Appl. Catal. B Environ. 2016, 183, 377–385. [Google Scholar] [CrossRef]
  97. Liu, S.; Guo, R.; Sun, X.; Liu, J.; Pan, W.; Xin, Z.; Shi, X.; Wang, Z.; Liu, X.; Qin, H. The deactivation effect of Cl on V/TiO2 catalyst for NH3-SCR process: A DRIFT study. J. Energy Inst. 2019, 92, 1610–1617. [Google Scholar] [CrossRef]
  98. Andonova, S.; Vovk, E.; Sjoblom, J.; Ozensoy, E.; Olsson, L. Chemical deactivation by phosphorous under lean hydrothermal conditions over Cu/BEA NH3-SCR catalysts. Appl. Catal. B Environ. 2014, 147, 251–263. [Google Scholar] [CrossRef] [Green Version]
  99. Wang, L.; Su, S.; Qing, M.; Dai, Z.; Sun, Z.; Liu, L.; Wang, Y.; Hu, S.; Xu, K.; Xiang, J. MeltingsolidificationandleachingbehaviorsofV/Asduringco-combustion of the spent SCR catalyst with coal. Fuel 2019, 252, 164–171. [Google Scholar] [CrossRef]
  100. Peng, Y.; Si, W.; Li, X.; Luo, J.; Li, J.; Crittenden, J.; Hao, J. Comparison of MoO3 and WO3 on arsenic poisoning V2O5/TiO2 catalyst: DRIFTS and DFT study. Appl. Catal. B Environ. 2016, 181, 692–698. [Google Scholar] [CrossRef]
  101. Xu, Y.; Zhang, Y.; Liu, F.; Shi, W.; Yuan, J. CFD analysis on the catalyst layer breakage failure of an SCR-DeNOx system for a 350 MW coal-fired power plant. Comput. Chem. Eng. 2014, 69, 119–127. [Google Scholar] [CrossRef]
  102. Nova, I.; Ciardelli, C.; Tronconi, E.; Chatterjee, D.; Bandl-Konrad, B. NH3–NO/NO2 chemistry over V-based catalysts and its role in the mechanism of the Fast SCR reaction. Catal. Today 2006, 114, 3–12. [Google Scholar] [CrossRef]
  103. Lang, E.; Drtina, P. Numerical simulation of the fluid flow and the mixing process in a static mixer. Int. J. Heat Nass Transf. 1995, 38, 2239–2250. [Google Scholar] [CrossRef]
  104. Xu, Y.; Zhang, Y.; Wang, J.; Yuan, J. Application of CFD in the optimal design of a SCR-DeNOx system for a 3000 MW coal-fired power plant. Comput. Chem. Eng. 2013, 49, 50–60. [Google Scholar] [CrossRef]
  105. Khatun, N.; Tiwari, S.; Lal, J.; Tseng, C.; Liu, S.W.; Biring, S.; Sen, S. Stabilization of anatase phase by uncompensated Ga-V co-doping in TiO2: A structural phase transition, grain growth and optical property study. Ceram. Int. 2018, 44, 22445–22455. [Google Scholar] [CrossRef]
  106. Byrne, C.; Moran, L.; Hermosilla, D.; Merayo, N.; Blanco, A.; Rhatigan, S.; Hinder, S.; Ganguly, P.; Nolan, M.; Pillai, S. Effect of Cu doping on the anatase-to-rutile phase transition in TiO2 photocatalysts: Theory and experiments. Appl. Catal. B Environ. 2019, 246, 266–276. [Google Scholar] [CrossRef]
  107. Hanaor, D.; Sorrell, C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef] [Green Version]
  108. Liu, R.; Li, H.; Duan, L.; Shen, H.; Zhang, Q.; Zhao, X. Influences of annealing atmosphere on phase transition temperature, optical properties and photocatalytic activities of TiO2 phase-junction microspheres. J. Alloys Compd. 2019, 789, 1015–1021. [Google Scholar] [CrossRef]
  109. Dauksta, E.; Medvids, A.; Onufrijevs, P.; Shimomura, M.; Fakuda, Y.; Murakami, K. Laser-induced crystalline phase transition from rutile to anatase of niobium doped TiO2. Curr. Appl. Phys. 2019, 19, 351–355. [Google Scholar] [CrossRef]
  110. Du, X.; Wu, Y.; Kou, Y.; Mu, J.; Yang, Z.; Hu, X.; Teng, F. Amorphous carbon inhibited TiO2 phase transition in aqueous solution and its application in photocatalytic degradation of organic dye. J. Alloys Compd. 2019, 810, 151917. [Google Scholar] [CrossRef]
  111. Chen, H.; Xia, Y.; Fang, R.; Huang, H.; Gan, Y.; Liang, C.; Zhang, J.; Zhang, W.; Liu, X. The effects of tungsten and hydrothermal aging in promoting NH3-SCR activity on V2O5/WO3-TiO2 catalysts. Appl. Surf. Sci. 2018, 459, 639–646. [Google Scholar] [CrossRef]
  112. Zhao, K.; Han, W.; Tang, Z.; Zhang, G.; Lu, J.; Lu, G.; Zhen, X. Investigation of coating technology and catalytic performance overmonolithic V2O5–WO3/TiO2 catalyst for selective catalytic reductionof NOx with NH3. Colloids Surf. A Physicochem. Eng. Asp. 2016, 503, 53–60. [Google Scholar] [CrossRef]
  113. Khatun, N.; Anita; Rajput, P.; Bhattacharya, D.; Jha, S.; Biring, S.; Sen, S. Anatase to rutile phase transition promoted by vanadium substitution in TiO2: A structural, vibrational and optoelectronic study. Ceram. Int. 2017, 43, 14128–14134. [Google Scholar]
  114. Liu, X.; Wu, X.; Xu, T.; Weng, D.; Si, Z.; Ran, R. Effects of silica additive on the NH3-SCR activity and thermal stability of a V2O5/WO3-TiO2 catalyst. Chin. J. Catal. 2016, 37, 1340–1346. [Google Scholar] [CrossRef]
  115. Sun, C.; Dong, L.; Yu, W.; Liu, L.; Li, H.; Gao, F.; Dong, L.; Chen, Y. Promotion effect of tungsten oxide on SCR of NO with NH3 for the V2O5–WO3/Ti0.5Sn0.5O2 catalyst: Experiments combined with DFT calculations. J. Mol. Catal. A Chem. 2011, 346, 29–38. [Google Scholar] [CrossRef]
  116. Marberger, A.; Elsener, M.; Nuguid, M.; Ferri, D.; Krocher, O. Thermal activation and aging of a V2O5/WO3-TiO2 catalyst for the selective catalytic reduction of NO with NH3. Appl. Catal. A General 2019, 573, 64–72. [Google Scholar] [CrossRef]
  117. Nova, I.; dall’Acqua, L.; Lietti, L.; Giamello, E.; Forzatti, P. Study of thermal deactivation of a de-NOx commercial catalyst. Appl. Catal. B Environ. 2001, 35, 31–42. [Google Scholar] [CrossRef]
  118. Ma, Z.; Deng, J.; Li, Z.; Li, Q.; Zhao, P.; Wang, L.; Sun, Y.; Zheng, H.; Pan, L.; Zhao, S.; et al. Characteristics of NOx emission from Chinese coal-fired power plants equipped with new technologies. Atmos. Environ. 2016, 131, 164–170. [Google Scholar] [CrossRef]
  119. Tomeczek, J. Coal Combustion; Kriege Publishing Company: Malabar, FL, USA, 1994. [Google Scholar]
  120. Bujalski, M.; Zyrkowski, M.; Nabaglo, D.; Szczepanek, K. The algorithm of steam soot blowers operation based on the monitoring of fouling factors of heating surfaces of a coal-fired boiler under operating conditions. E3S Web Conf. 2019, 82, 01001. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Minimum safe operating temperature for selective catalytic reduction (SCR) reactors as the function of SO3 concentration, according to Charles [65].
Figure 1. Minimum safe operating temperature for selective catalytic reduction (SCR) reactors as the function of SO3 concentration, according to Charles [65].
Energies 13 06200 g001
Figure 2. Correlation between flue gas temperature and the load level of pulverized-coal boiler with 120 kg/s nominal steam capacity (own source).
Figure 2. Correlation between flue gas temperature and the load level of pulverized-coal boiler with 120 kg/s nominal steam capacity (own source).
Energies 13 06200 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zyrkowski, M.; Motak, M.; Samojeden, B.; Szczepanek, K. Deactivation of V2O5−WO3/TiO2 DeNOx Catalyst under Commercial Conditions in Power Production Plant. Energies 2020, 13, 6200. https://doi.org/10.3390/en13236200

AMA Style

Zyrkowski M, Motak M, Samojeden B, Szczepanek K. Deactivation of V2O5−WO3/TiO2 DeNOx Catalyst under Commercial Conditions in Power Production Plant. Energies. 2020; 13(23):6200. https://doi.org/10.3390/en13236200

Chicago/Turabian Style

Zyrkowski, Maciej, Monika Motak, Bogdan Samojeden, and Krzysztof Szczepanek. 2020. "Deactivation of V2O5−WO3/TiO2 DeNOx Catalyst under Commercial Conditions in Power Production Plant" Energies 13, no. 23: 6200. https://doi.org/10.3390/en13236200

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

Article Metrics

Back to TopTop