Next Article in Journal
Carbazolyl-Substituted [OSSO]-Type Zirconium(IV) Complex as a Precatalyst for the Oligomerization and Polymerization of α-Olefins
Next Article in Special Issue
Characteristics of Water and Urea–Water Solution Sprays
Previous Article in Journal
Highly Selective Oxidation of 5-Hydroxymethylfurfural to 5-Hydroxymethyl-2-Furancarboxylic Acid by a Robust Whole-Cell Biocatalyst
Previous Article in Special Issue
Sulfur Poisoning Effects on Modern Lean NOx Trap Catalysts Components
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

WO3–V2O5 Active Oxides for NOx SCR by NH3: Preparation Methods, Catalysts’ Composition, and Deactivation Mechanism—A Review

by
Weidong Zhang
1,2,
Shuhua Qi
2,
Giuseppe Pantaleo
1,* and
Leonarda Francesca Liotta
1,*
1
Institute for the Study of Nanostructured Materials-Council National Research (ISMN-CNR), 90146 Palermo, Italy
2
Department of Applied Chemistry, School of Natural and Applied Sciences, Northwestern Polytechnical University, Xi’an 710072, China
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(6), 527; https://doi.org/10.3390/catal9060527
Submission received: 19 April 2019 / Revised: 31 May 2019 / Accepted: 4 June 2019 / Published: 13 June 2019

Abstract

:
Researchers in the field of the selective catalytic reduction (SCR) of nitrogen oxides (NOx: NO, NO2, or N2O) by NH3 are still greatly challenging the optimization of low-temperature activity and selectivity, high-temperature stability, resistance to alkali metals and other poisoning agents, such as Hg, As, etc. The present study reviews the research progress, related to the latest 20 years, on WO3–V2O5-based catalysts that are expected to overcome the catalytic performances of the current SCR catalytic devices. In details, the effects of the synthesis methods, chemical composition, type of supports (metal oxides, molecular sieves, and filters), doping elements, or metal oxides added as promoters of WO3–V2O5-based catalysts and, finally, the influence of SO2 and H2O in the reaction mixture are addressed. The importance of understanding the deactivation mechanism in the presence of several poisoning agents is also emphasized, which should be taken into consideration for the design of new catalysts.

1. Introduction

1.1. NOx Problem, Control Regulations, and Removal Technologies

Agriculture, industries, power generation plants, and transportation are the major sources responsible of air pollution [1]: Ozone depletion, eutrophication, smog, global warming, and acid rain have attracted wide attention from human beings [2,3]. In order to reduce air pollution, new regulations have been introduced in the latest years, such as the Directive 2010/75/EU of the European Parliament and of the Council concerning industrial activities, European standards for vehicles’ exhaust gas emissions in Europe, as well as EPA (Environmental Protection Agency) tier regulations in the USA and analogous emission standards developed by the Japanese government.
Naval diesel engines use heavy fuel oil with a high sulfur content, thus most of the pollutants are nitrogen oxides (NOx), particulate matter (PM), and sulfur oxides (SOx) [4]. To limit NOx emissions, strict regulations and legislations have been set up by the international maritime organization (IMO) that came into force on 1 January 2016 (MARPOL Annex VI) and will be tighter in the future [5].
Many techniques have been applied for the abatement of NOx emissions, such as pre-combustion and combustion modifications [6,7], as well tail-end control equipment [2,8]. Tail-end control devices are the most efficient in the removal of NOx, including selective noncatalytic reduction (SNCR) [9,10], selective catalytic reduction (SCR) [11], NOx storage/reduction (NSR) [12], and electron beam radiation [2]. Among the above-mentioned methods, NOx-SCR, which was firstly developed in Japan in the 1970s, is applied worldwide today, being regarded as a promising technology [13].
One of the last key reviews addressing the chemical and mechanistic aspects of the selective catalytic reduction of NOx by ammonia over oxide catalysts was published in 1998 by Busca et al. [14]. The authors reviewed the performances of V2O5–WO3/TiO2 and V2O5–MoO3 catalysts and compared with catalysts containing Fe2O3, CuO, MnOx, and CrOx with special attention given to spectroscopic studies, adsorption–desorption measurements, and kinetic experiments. However, the reaction mechanism was not unanimously recognized and some disagreement points were underlined. Since then, the number of articles dealing with NO SCR by NH3 over WO3–V2O5 basic catalysts has undergone dramatic growth. Moreover, in the last 20 years, new restrictive limits to NOx emissions have been introduced by world regulations. Therefore, we consider that is worthwhile to summarize here the most recent results on WO3-V2O5 active oxides, paying attention to the chemical effects (preparation methods, catalysts’ composition, promoting effects, and deactivation mechanism) that are still a matter of debate in the literature.

1.2. SCR Application for After-Treatment of Fossil Fueled Engines in Transportation

For transportation, NOx emissions originate from vehicles’ and ships’ engines. The automotive industry, especially for heavy-duty truck and bus engines, has adopted urea-SCR (Selective Catalytic Reduction) to reach the NOx limits required by regulations to meet the Euro V (2008) and the JP 2005 NOx limits. The first commercial diesel truck applications were launched in November 2004 by Nissan Diesel in Japan and in early 2005 by Daimler in Europe [15]. In light-duty vehicles, the SCR was introduced in some US EPA tier 2 vehicles. By 2012 to 2015, most of the tier 2 vehicles with NOx adsorbers had been converted to urea-SCR. In Europe, the SCR was introduced on certain Euro 5 models, with a much wider application of the technology in Euro 6 vehicles to meet the US Tier 4i/EU Stage IIIB emission standards.
For marine SCR applications, the main challenges are sulfur resistance and activation at low temperatures [16]. The latter is also a challenge for cars, trucks, and non-road applications and may imply that the catalyst remains inactive during start-up and maneuvering [17]. Nowadays, one of the most investigated catalysts for NH3–SCR is V2O5–WO3 (MoO3)/TiO2, which can meet the requirement at temperatures in range ~300–400 °C [18]. At temperatures as low as 100–150 °C, no effective catalytic after-treatment of exhaust gases occurs. For that reason, the SCR catalyst should be placed upstream of the turbocharger, as close to the engine as possible to allow high temperatures for high NOx conversion and to avoid the formation of ammonium sulfates. In fact, on the one hand, the SCR unit is located downstream of the electrostatic precipitator and desulfurizer to avoid the catalyst deactivation caused by ash and SO2 poisoning [19,20]. On the other hand, the high pressure increases the catalytic reaction but also enhances the undesired side reactions. Based on the above considerations, an exploitation and investigation of V2O5–WO3-based catalysts with low-temperature activity and high-temperature stability as well as resistance to Hg, alkali metals, is still a great challenge for practical applications.
Over the last 20 years, to overcome the deficiency, many kinds of V2O5–WO3-based catalysts have been prepared with different methods, supports, chemical compositions, and doping agents. Therefore, in this article, we overview the recent progresses in catalytic NOx SCR by NH3, especially focusing on V2O5–WO3-based catalysts.

2. WO3–V2O5-Based SCR Catalysts

2.1. Reasons for the Use of WO3–V2O5-Based Catalysts

To date, WO3–V2O5/TiO2 catalysts are the most effective systems for practical applications and are widely used for NO SCR with NH3. Therefore, the effect of WO3 and V2O5 on the catalytic performance must be understood for further research in this area.
According to the investigation of Zhang et al. [21], WO3 play an important role in catalyzing NO SCR with NH3, due to the following promotional effects of WO3: (i) Increases the amounts of Lewis acidic sites; (ii) inhibits the sintering of anatase TiO2; and (iii) WO3 can be easily reduced to W5+ and the standard reduction potential of W6+ to W5+ is only −0.03 eV. Thus, the reduced W5+ species can be oxidized by oxygen, meanwhile, oxygen may be reduced to superoxide ions. In the case of V2O5, (i) the surface acidity (Brønsted and Lewis) can be improved by adding V2O5; and (ii) there is a direct correlation between high surface V3+ + V4+/Vn+ and the V4+/V5+ ratio and the SCR activity. Furthermore, the behavior of the oxygen species bonded to the vanadium and the dispersion of the active species are influenced by the support. Well-dispersed and isolated vanadium oxide species show low activity for the SCR reaction but present a high selectivity towards N2.
More importantly, it has been reported that the interaction between WO3 and V2O5 enhances the catalytic performance [22]: (i) V2O5 increases Brønsted and Lewis acid sites, while WO3 promotes Lewis acid sites. Both types of acid sites promote the NO SCR—the Brønsted acidity favors the conversion at low temperatures, up to 300 °C, while Lewis acid sites are the active sites above 300 °C; (ii) by increasing the amount of reduced V4+ species, the NO conversion also increases. The effects of the supporting WO3 and V2O5, active oxides, over titania have been also addressed by the authors [21], who investigated V2O5/TiO2, V2O5-WO3/TiO2, and WO3/TiO2-nanotube model catalysts for NO SCR with NH3. The addition of WO3 between 3 and 6 wt % in the V2O5/TiO2 catalysts increased the reduced V4+ species on the V2O5–WO3/TiO2 catalysts. However, it has been reported that above 6 wt % of WO3, the reduced V4+ species remain constant on the V2O5–WO3/TiO2 catalysts; and (iii) tetrahedral monovanadate and polymeric surface vanadate also coexist in V2O5-WO3/TiO2 catalysts.
Finally, the effects of SO2 and H2O on the catalytic activity of V2O5–WO3/TiO2 catalysts were also evaluated by the same authors [22]. They found that the NO conversion is not affected by SO2 but water inhibits the catalytic activity. The influence of SO2 and H2O in the reaction stream, strongly depending on the catalyst composition and reaction conditions, will be addressed in more detail later in the manuscript.
The changes occurring on WO3 and V2O5 after supporting over TiO2 must be highly evaluated. The relation between tungsten and vanadium oxide species in V2O5–WO3/TiO2 catalysts was widely investigated for two series of catalysts containing 0.5 to 5 wt % V2O5 and 0 or 10 wt % WO3 impregnated onto a high surface-area titania hydrate with an anatase structure [23]. The above-mentioned loadings were chosen in order to keep the theoretical coverages of the transition metal oxides on the surface equal or less than 1. It has been reported that the interaction of TiO2 with supported V(V) or W(VI) oxide species can strongly affect the structure of the surface oxide phases formed which strongly differ from the corresponding bulk oxides. For instance, isolated surface oxide species having strongly distorted tetrahedral geometry were formed in the case of vanadium at low V oxide coverage, in dehydrated samples. At higher loadings, oligomeric or polymeric metavanadate species forming two-dimensional islands along with V2O5 nanoparticles were observed. It is not clear if aggregates or oxide crystals are formed for V2O5 contents below the theoretical monolayer, however, it was clearly demonstrated that well-prepared catalysts form two-dimensional surface oxides until the support is completely covered, then the surface oxide species grow into the third dimension. In the case of dehydrated WO3/TiO2 catalysts, formation of strongly distorted octahedral surface species ((-O)5–W = O) occurs up to high tungsten oxide coverage [23].

2.2. Effect of Preparation Method on Activity and Selectivity

According to the literature, impregnation (wet, dry, and incipient wetness) [24,25,26,27], co-precipitation [26], deposition–precipitation [28], sol–gel [29], and grafting [30] are the main synthesis methods to prepare V2O5–WO3-based SCR catalysts and all of them are summarized in Table 1.
As shown in Figure 1, Yu et al. [24] investigated the NH3-SCR activities of V2O5–WO3/TiO2 catalysts as fresh and poisoned by potassium chloride, prepared by wet and dry impregnation methods. They found that the wet-impregnated sample, WWT(W), shows a slightly higher activity than the impregnated one (D) at around 200 °C and this was ascribed to the formation of agglomerated V species which increases the intensity of Brønsted acid sites and the oxidation of adsorbed NH3, while isolated V species were predominant on the dry-impregnated catalyst. After the addition of KCl, both catalysts deactivated due to the decreasing amount of ammonium ions coordinated on the Brønsted acid sites. Isolated vanadia species (DK) appeared to be more reactive with potassium and therefore more severely deactivated.
Dong et al. [25] optimized the preparation conditions (wet impregnation) by focusing on the effect of the pH value of the vanadium precursor solution on the catalytic performance of V2O5-WO3/TiO2 in low temperature NH3-SCR of NOx. According to XPS (X-Ray Photoelectron Spectroscopy), Raman, H2-TPR (Temperature Programmed Reduction), NH3-TPD (Temperature Programmed Desorption), and NH3-DRIFT (Diffuse Reflectance Infrared Fourier Transform Spectroscopy) characterizations, it resulted that the outstanding activity of the catalyst with the decreased pH value could be responsible for the following: More polymeric vanadium species were formed on the catalyst surface; meanwhile, the ratio of V4+(3+)/V5+, surface acidity, and quantity of active sites were increased. Thus, it should be considered that an enhancement of the precursor solution acidity is an effective way to improve SCR catalysts’ activity.
The influence of the catalyst synthesis method on the SCR of NO by NH3 over V2O5–WO3/TiO2 catalysts was also investigated by He et al. [26]. In this article, they prepared the catalysts by two different synthesis methods: Co-precipitation of aqueous vanadium and tungsten oxide precursors with TiO(OH)2 and incipient wetness impregnation of the aqueous precursors on a crystalline TiO2 support. The composition difference of the two types of catalysts was confirmed by XRD (X-Ray Diffraction) and HS-LEIS (High Sensitivity-Low Energy Ion Scattering Spectroscopy) and a distinction of their activity was also revealed by in situ Raman and IR (Infrared) spectroscopy. Co-precipitated catalysts contain oligomerized surface mono-oxo O = VO3 and O = WO4 sites on the TiO2 support. The enhanced performance of co-precipitated catalysts compared to the impregnated ones was associated with the formation of new O = WO4 species, able to adsorb NH3, and to the presence of redox surface mono-oxo O = VO3 sites.
Qi et al. [27] reported the synthesis of V2O5–WO3/TiO2 catalysts from waste commercial SCR catalysts through oxalic acid leaching and impregnation with different V2O5 loadings (0.5%, 1.0%, and 1.5 wt %). The NO conversion of the newly synthesized catalyst with 1.0% V2O5 almost recovered the activity of the fresh catalyst (91% of the NO conversion at 300 °C) and showed good resistance to SO2 and H2O.
Meiqing et al. [28] systematically investigated Ce-modified V2O5–WO3/TiO2 (V/WTi) catalysts by the deposition–precipitation method (V/Ce/WTi-DP and Ce/WTi-DP) and the conventional impregnation method (V/Ce/WTi-IMP and Ce/WTi-IMP). As shown in Figure 2, the results revealed that the activity was mainly dependent on the status of Ce, which not only possessed more surface Ce species but also presented a higher reducibility of the Ce species. The higher NOx conversion can be assigned to more active Lewis acid sites, the weakly adsorbed NO2, and the monodentate nitrate.
Djerad et al. [29] prepared V2O5–WO3/TiO2 catalysts by the sol–gel method with different W and V loadings and calcined at different temperatures. The results are shown in Figure 3, in which it is shown that the V2O5 content strongly influences the catalytic behavior: Weakly active for the SCR reaction but with a high selectivity to N2. Moreover, the NOx conversion and N2 selectivity achieved up to 100% in the temperature range of 250 to 325 °C. Meanwhile, 3 wt % V2O5–9 wt % WO3/TiO2 was the best catalyst not only for the NOx conversion, but also for the N2 selectivity.
Reiche et al. [30] explored the influence of the grafting sequence and simultaneous grafting of V2O5 and WO3 on the properties of V2O5–WO3/TiO2 catalysts. DRIFT and XPS indicated that V and W species do not exist independently on the TiO2 surface. More importantly, V2O5–WO3/TiO2 catalysts synthesized by sequential grafting do not result in a hierarchical structure. However, changing the grafting mode and metal oxides loadings significantly affected the strength of interaction of V2O5 and WO3.
In Table 1 and from the above-mentioned research, we cannot directly affirm which was the best method to synthesize catalysts with high activity and selectivity in a broad temperature window. New methods able to improve the specific surface area and to increase the active sites of catalysts should be considered in the design of new catalysts. In addition, the activity and selectivity are significantly affected by the active metal oxides loading.

2.3. Effect of the Support on Catalytic Activity

To enhance the stability and the activity, V2O5–WO3-based multiple metal oxides and supported catalysts have been widely studied. They can be classified into two types: Metal oxides and molecular sieves or filters as carriers.

2.3.1. Metal Oxides as Carriers

V2O5–WO3/TiO2 is a classical commercial catalyst that has been widely investigated for NO SCR. It has been reported that only the TiO2 anatase phase has no activity for NO SCR with NH3 [31]. However, it is well known that TiO2 anatase improves the catalytic activity by dispersing the V2O5-WO3 active phases [32]. At high temperatures, >700 °C, transition from the anatase to rutile phase occurs and such process is strongly influenced by the oxygen vacancies’ concentration [21]. As will be discussed in the following Section 2.4, the presence of oxygen vacancies in the support helps the formation of W5+ species able to active superoxide ions.
Reiche et al. [30,33] prepared V2O5–WO3/TiO2 by simultaneous grafting of TiO2 with V2O5 and WO3. According to the DRIFT, XPS, and TOF-SIMS (Time-of-Flight secondary ion mass spectrometry) investigations, indicating V and W species do not exist independently on the titania surface (V-O-W connectivity were present on the V2O5–WO3/TiO2 catalysts), which results in a higher activity compared to the corresponding titania-supported single oxides. The studies demonstrate that the activity could be enhanced by grafting V2O5–WO3 on TiO2.
Zhang et al. [34] elucidated the interaction of V, W, and Ti species on the surface of V2O5-WO3/TiO2 catalysts for the improvement of the catalytic activity. XRD, UV-vis, PL spectra, DFT theoretical calculations, and in situ DRIFT showed that WO3 could interact with TiO2 to improve the electrons transfer due to the V-O-W species, in agreement with Reiche et al. [30,33]. Moreover, the XPS and EPR observations confirmed the presence V4+ species, which are the active sites for the superoxide ion formation. Such species promoted NO oxidation to NO3 and improved the decomposition of NO3 to NO2, facilitating the reaction rate of “fast SCR”.
Kompio et al. [23] highlighted a new point on the relations between tungsten and vanadium, studying two series of V2O5–WO3/TiO2 catalysts with different V2O5 contents (0.5–5 wt %) and 0 or 10 wt % WO3 impregnated over titania hydrate with an anatase structure. The catalysts were investigated by XRD, specific surface area measurements, Raman and EPR (Electron Paramagnetic Resonance) spectroscopy, temperature-programmed reduction (TPR), and SCR. It was concluded that tungsten plays a promotional effect on the neighboring vanadium oxide species, favoring their dispersion.
The positive role of W was confirmed also by Zhang et al. [35] and Wang et al. [36], who found that WO3 species play an important role in the formation of superoxide ions (O2) over V2O5-WO3/TiO2 catalyst for NH3-SCR of NO. This was due to an improved electron transfer between the V-O-W species, which facilitates the formation of reduced V2O5 (V4+ and V3+). The best activity was registered with a WO3 loading of 6 wt %. However, it should be noticed that WO3 was an electron withdrawal, so higher WO3 loadings might trap electrons to inhibit the formation of superoxide ions.
Zhang et al. [37] investigated the promotional roles of ZrO2 and WO3 over V2O5-WO3/TiO2-ZrO2 SCR catalysts. They found that ZrO2 improved the thermally stability, while WO3 increased the Brønsted acid sites of the catalysts. Furthermore, a combination of WO3 and ZrO2 enhanced the dispersion of all metal oxides. As shown in Figure 4, the highest NOx conversion was achieved over V2O5-9wt.% WO3/TiO2-ZrO2 and this finding partially results from the ZrO2 doping that considerably increased the BET surface area and favored the dispersion of the active components on the surface. However, the mainly promotional role was to change the pathway of NOx reduction and the surface acidity of V2O5-WO3/TiO2. There are only Lewis acid sites on the V2O5–WO3 supported over TiO2–ZrO2, while Brønsted and Lewis acid sites coexisted on V2O5–WO3/TiO2.
Herrmann et al. [38] prepared a V2O5–WO3/TiO2 catalyst by supporting the V2O5-WO3 over TiO2 EUROCAT and investigated the electrical properties under different atmospheres, such as pure oxygen, NO, and NH3 with and without O2. In the absence of oxygen, NO has an electron acceptor character (as NO) that is able to fill the anionic vacancies, which are mostly present on V2O5. On the other hand, NH3 had a strong reduction behavior with the formation of ionized vacancies. In the presence of oxygen, the anionic vacancies of the catalysts were filled by oxygen, therefore, the NO did not exhibit its electron acceptor character anymore, reacting directly with NH3, bounded on the acid sites. Under the NO + NH3 mixture, without oxygen, an increased electrical conductivity was found, ascribed to the formation of anionic vacancies, especially over vanadia, due to dihydroxylation and dehydration of the surface.
Jung et al. [39] supported V2O5–WO3 on TiO2–SO42− and characterized the catalysts by BET, XRD, XPS, TPD, and H2-TPR analyses. Compared with the V2O5–WO3/TiO2 catalyst, in the V2O5-WO3/TiO2-SO42−, the sulfate deposited over titania did not affect the physical properties (in terms of the BET surface area and phase transformation). Conversely, in TPD and H2-TPR, it was found that the sulfate stability strongly depended on the loaded metal oxides. In addition, the reduction property of V2O5 also changed due to the interaction between the V-O-W species. In Figure 5, the SCR activity in terms of the NO conversion over TiO2-SO42−, V2O5/TiO2-SO42−, WO3/TiO2-SO42−, and V2O5-WO3/TiO2-SO42−, is compared.
The similar activity between V2O5/TiO2-SO42− and V2O5-WO3/TiO2-SO42− at low temperatures was explained by focusing on the dispersion between acid and redox sites. At high temperatures, the redox property modified by the V-O-W species contributed to the high activity.
Sun et al. [40] prepared a V2O5–WO3/Ti0.5Sn0.5O2 (TS) SCR catalyst by supporting V2O5–WO3 on Ti0.5Sn0.5O2. The results revealed that the activity of the NH3-SCR reaction was closely correlated to the Brønsted acid sites, and they were in direct proportion with the amount of WO3 up to the highly-dispersed particles formed. Indeed, when the WO3 exist in a dispersed state, the tungsten atoms are adjacent to the vanadium atoms forming the V-O-W bonds that are responsible for the increased Brønsted acidity. However, further increases of WO3 (loadings ≥ 1.0 mmol/100 m2 TS) results in the formation of crystalline WO3 that covers the active sites, leading to a decrease of the activities.
Camposeco et al. [41] synthesized a series of titanic acid nanotubes (TANs) by the conventional hydrothermal method. Then, V2O5/WO3 was added by wet impregnation. The catalytic activity of the V2O5 (5 wt %)–WO3 (3 wt %)/TAN catalysts was strongly improved for the SCR-NH3 process in a wide range of temperatures in comparison with both samples, V2O5 (10 wt %)/TAN and V2O5-WO3/TiO2 (see Figure 6a). Interestingly, the above mentioned V2O5 (5 wt %)-WO3 (3 wt %)/TAN catalyst showed better NO conversion values than V2O5 (10 wt %)/TAN in the presence of water (10 vol %) and SO2 (50 ppm) (Figure 6b). The reason for the higher activity can be explained by the following factors: (1) The excellent physical properties that TANs possess, such as tubular structure, good thermal stability (460 °C), and high specific surface area (314 m2/g); and (2) the presence of Brønsted and Lewis acid sites that promote NO conversion in a wide range of temperatures.
Analogous effects of TANs were reported by Mejía-Centeno et al. [42] (Figure 6c,d). Aguilar-Romero et al. [22] also prepared TANs and employed these as support for V2O5-WO3. They investigated: The effect of WO3 and V2O5 loading on the TiO2-NT surface acidity, the effect of WO3 on the V4+/V5+ ratio, and its correlation with the catalytic SCR activity. V2O5-WO3/TiO2-NT displayed a higher NO conversion than WO3/TiO2-NT and V2O5/TiO2-NT. In fact, the addition of both V2O5 and WO3 is responsible for the increased surface acidity (Brønsted and Lewis) of the TiO2–nanotubes. Moreover, they concluded that the morphology of the nanotubes and the metal loading play important roles for the catalyst activity.
Boudali et al. [43] prepared a WO3-V2O5/STi-PILC (Sulfated Titanium Pillared Clay) catalyst for NH3-SCR. The results showed that the highest NO conversion can reach up to 100% at 300 °C. Vanadia enhanced the NO conversion over STi-PILC, while no significant effect of tungsten was observed for WO3-V2O5/STi-PILC catalyst. When both tungsten and sulfate were simultaneously present on the surface of vanadia supported STi-PILC, the sulfate species seemed to play a more important role for NO abatement than tungsten.
Najbar et al. [44] synthesized Ti–Sn-rutile for deposition of WO3-V2O5. The SCR activity and surface species structure were determined for the freshly prepared catalyst, for the catalyst previously used in NO reduction by ammonia (320 ppm NO, 335 ppm NH3, and 2.35 vol % O2) at 300 °C as well as for the catalyst previously annealed at 300 °C under 2.35 vol % O2/He. A significant evolution of the active species occurred during the use of the catalyst at 300 °C, especially the increase of the oxidation state of vanadium and the decrease of the tungsten content in the surface layers of the catalyst. Changes in the SCR activity occurring upon the different treatments were discussed in terms of dissociative adsorption of water and the OH group’s Brønsted acidity. In detail, the dissociative adsorption of water on V5+OW6+ sites appeared to be mainly responsible for the activity at 200 °C, while water adsorption on both the V5+OW6+ and V4+OW6+ sites determined the activity at 250 °C.

2.3.2. Molecular Sieves or Filters as the Carrier

Li et al. [45] reported an enhanced activity of NO SCR and mechanical stability of V2O5-WO3/TiO2 catalyst supported on cordierite after acid pre-treatment of the cordierite substrate and the addition of silica. The acid pre-treatment of cordierite significantly increased the BET surface area and pore volume of the catalysts.
Silica increased the surface acidity of the V2O5–WO3/TiO2 catalyst, and the strongest surface acidity was registered at the Si/Ti molar ratio of 5/5. Moreover, by XPS analysis, it was shown that the incorporation of Si shifted the V valence from V5+ to V4+, increasing the amount of chemisorbed oxygen, with a positive effect on SCR activity. The optimal catalyst showed a high NO conversion above 90% at 300 to 500 °C, high mechanical stability, and resistance to sulfur and hydrothermal aging.
Kim et al. [46] fabricated three kinds of V2O5–WO3/TiO2 supported-SiC catalytic filters by using the rotational coating method and investigated their NO conversion performances. Three different fabrication routes were used in order to find the effective method for the preparation of the best V2O5–WO3/TiO2 supported-SiC catalytic filters. Method-III among the three different methods used appeared to be the most effective one. This method deposited the active component, V2O5-WO3, by the wet impregnation method over titania, then the powder was dried and calcined at 450 °C for 5 h. The resulting preformed catalyst was ground in wet solution conditions with a ball mill and then used for coating the SiC filter. It was demonstrated that by grinding the catalyst with ball milling, the catalyst particle size was lowered and the particles were dispersed in the coating solution (to produce the high catalytic surface area), which diffused into the pores of the SiC filters, favoring the catalyst dispersion and increasing the exposed active sites. The catalytic filter prepared by method-III (with a 50 nm catalyst particle confirmed with SEM) showed the 99.9% NO conversion with the Nx-slip concentration below 15 ppm in the optimal temperature range from 230 to 350 °C. Figure 7a,b display the NO conversions and N2O concentration, respectively, over V2O5–WO3/TiO2 supported-SiC catalysts prepared with the three different methods.
Choi et al. [47] investigated the effect of Pt on the V2O5–WO3/TiO2 supported-SiC catalytic filter (Pt-V2O5-WO3-TiO2/SiC). The Pt-added catalytic filter lowered the optimum working temperature from 280 to 330 °C for the non-Pt impregnated catalyst (V2O5-WO3-TiO2/SiC) to 180 to 230 °C for the Pt-V2O5-WO3-TiO2/SiC. Pt promoted catalytic activity at low temperatures, but increased ammonia oxidation. The Pt promotional effect was believed to result from a high electron transfer achieved on the Pt–catalytic system, as shown in Figure 7c,d.
Najbar et al. [48] focused on the competition between NO reduction and decomposition over V2O5-WO3/Ti(Sn)O2 catalyst on a Cr-Al steel monolith. Furthermore, they also revealed that the NO conversion was higher than NH3 over the reduced catalyst. An increase of the pre-reduction temperature was in favor of improving the activity and stability in direct NO decomposition. The reason for the high selectivity of NO decomposition to N2 was attributed to tungsten nitrosyl complexes via the W-N bond.
Zhao et al. [49] obtained promising catalytic performance over monolithic V2O5-WO3/TiO2 catalysts by impregnating the active components (V2O5, WO3, and TiO2) and optimizing the relative amounts and the deposition procedure. The active coating layers were stably deposited over the monolithic substrate as revealed by ultrasonic treatment. XPS analysis confirmed the co-existence of V5+ and V4+ species, while the BET surface area was significantly improved after the coating of the monolith due to its porous structure.
According to the data summarized in Table 2, it reveals that SiC or Pt doped SiC filters are the best carriers for promotion, at low-temperatures, of the activity of V2O5-WO3-based catalysts. Indeed, SiC and Pt-doped Si-C filters play an important role in increasing the BET surface area and the dispersion of active components, as well as in the promotion of electron transfer due to Pt introduction. All these factors should be taken into account in the design of new catalysts in the future.

2.4. Effect of Chemical Composition on Catalytic Activity

Shen et al. [50] prepared V2O5-WO3/TiO2 and V2O5-WO3/TiO2-SiO2 SCR catalysts and investigated the resistance toward K and SO2 poisoning. The calcination temperature of V2O5-WO3/TiO2 catalysts was varied between 400 and 750 °C and played an important role for the activity (Figure 8a). Three different weight ratios of SiO2/TiO2 were used, 3 wt %, 6 wt %, and 9 wt %, respectively, and the loading of K poison was fixed as a K/V molar ratio of 0.5. A different catalytic behavior was observed for silica doped and non-doped catalysts: Over K-poisoned V2O5-WO3/TiO2-SiO2 catalysts, (Figure 8b), the NO conversion values reached up to 90% to 98% (280–410 °C), with a positive effect observed by increasing the silica loading. On the contrary, the K-poisoned catalyst without SiO2 showed lower activity with a maximum NO conversion of ~70% between 350 and 420 °C. The outstanding performances of V2O5–WO3/TiO2–SiO2 SCR catalysts were explained by the catalysts’ increased acidity due to the silica presence that prevents K poisoning, forming Si-OK scavenger species. As it concerns the SO2 poisoning, the authors investigated the effect on the catalyst, V2O5-WO3/TiO2, calcined at 750 °C, by adding 1500 ppm to the flue gas. They found that the NO conversion declined sharply by 25% in half an hour, then, by decreasing the SO2 concentration to 750 ppm over 4 h, the catalyst partially recovered the activity and maintained the 45% NO conversion for the following 3 h.
Liu et al. [51] also studied the effect of SiO2 additive on the NH3-SCR activity and thermal stability of V2O5-WO3/TiO2-SiO2 catalysts. Compared with V2O5-WO3/TiO2 catalyst, the SCR activity and selectivity of V2O5-WO3/TiO2-SiO2 catalyst was improved after aging at 750 °C in 10 vol % H2O/air for 24 h (in Figure 8c,d). The different behavior was explained by the addition of SiO2, which plays different roles: Inhibition of the shrinkage of the catalyst BET surface area due to the phase transition from anatase to rutile and the growth of the TiO2 size; and maintenance of the highly-dispersed W species.
Dong et al. [52] prepared Mn-Ce/TiO2 (denoted as M) and Cu-Ce/TiO2 (denoted as C) by the sol–gel method, then a series of V2O5-WO3/M/cordierite honeycomb ceramics and V2O5-WO3/C/cordierite honeycomb ceramics complex catalysts were obtained by coating V2O5-WO3, M, and C onto cordierite using the impregnation method. The results indicated that the higher activity originated from the high V4+/V5+ ratio and the large amount of surface chemisorbed oxygen, as well as the high specific surface area increased by the M and C composites.
Li et al. [53] prepared V2O5-WO3/TiO2 and CeO2–WO3/TiO2 catalysts by the impregnation method. In comparison with V2O5-WO3/TiO2 catalyst, the novel CeO2-WO3/TiO2 catalyst exhibited a much better resistance to hydrothermal aging. After aging at 670 °C in 5% H2O/air for 64 h, the NO conversion was equal to to 80% (238–573 °C) over CeO2-WO3/TiO2 catalyst. Such result was attributed to the well-dispersed CeO2 and W species, as well as to the formation of a new species, Ce2(WO4)3, due to the interaction of CeO2 and W. Ce2(WO4)3 prevented the W species’ crystallization and TiO2 phase transition (anatase to rutilo) and sintering. Meanwhile, the existence of the Ce4+-Ce3+ couple provided more oxygen vacancies and improved the surface Lewis acid of the catalyst.
Shi et al. [54] added zirconia to V2O5-WO3/TiO2 catalyst and investigated the performance before and after aging treatment (at 750 °C for 12 h). Compared with V2O5–WO3/TiO2 catalyst, ZrO2-V2O5-WO3/TiO2 catalyst without aging was slightly more active at high temperatures, but less active at low temperatures. After aging treatment, the NH3-SCR activity of the catalysts was significantly enhanced and exhibited the highest NO conversions and N2 selectivity (150–550 °C). The following key factors were invoked as being responsible for the improved catalytic behavior: (1) Addition of ZrO2 improves the thermal stability of ZrO2-V2O5-WO3/TiO2 catalysts and preserves the surface area from sintering, resulting in a better dispersion of V sites after aging; (2) ZrO2 enhances the strength of the surface acidic groups. Regarding the non-aged ZrO2-V2O5-WO3/TiO2 catalyst, the decreased activity at low temperatures was attributed to the inhibitions due to ZrO2, which decrease the amount of Brønsted acid sites and V5+ = O active sites, inhibiting ammonia absorption.
Tao et al. [55] focused on the relationship between the loadings of V2O5 and WO3 and the catalytic performance of V2O5-WO3/TiO2 catalysts for the SCR of NO with NH3. The catalytic activity was enhanced by increasing the V2O5 loading less than 2 wt %, in agreement with the presence of TiO2 as the anatase phase (by XRD) and to the high dispersion on the catalyst surface of all the active species (by XPS). Conversely, when the V2O5 loading exceeded 2 wt %, the activity began to decline because the high V2O5 loading on TiO2 favored the phase transition from anatase to rutile. On the other hand, the NO conversion was also improved by moderate WO3 loadings up to 6 wt %. Exceeding such a percentage, WO3 inhibited the formation of superoxide ions, decreasing the activity.
As shown in Figure 9, Lietti et al. [56] investigated the NO SCR over commercial and model V2O5-WO3/TiO2 and V2O5-MoO3/TiO2 catalysts. The results demonstrated that WO3 and MoO3 increased the activity as “structural” and “chemical” promoters for the catalysts, providing better mechanical, structural, and morphological properties. In Figure 9A,B, the NO conversion and N2 selectivity values are compared for V2O5-WO3/TiO2 and V2O5-MoO3/TiO2 catalysts at different V2O5 loadings and constant WO3 or MoO3 contents, respectively, 9 and 6 wt %. In both cases, the activity was increased by increasing the vanadia loading, so that the “temperature window” of the reaction was widened and shifted towards lower temperatures. However, the V2O5(x)-WO3(9)/TiO2 samples exhibited higher selectivity in a larger range of temperatures.
Xu et al. [57] investigated the low-temperature NH3-SCR activity of V2O5-Sb2O3/TiO2 and V2O5-WO3/TiO2 monolithic catalysts in the presence of SO2. V2O5-Sb2O3/TiO2 monolithic catalyst showed a higher NO conversion and broader temperature windows as well as a higher SO2 and H2O resistance. The excellent performance was suppressed by the extent of NH4HSO4 accumulated on the catalyst surface.
Table 3 summarizes the main results discussed so far on the chemical composition effects of V2O5–WO3/TiO2 based catalysts.

2.5. Positive/Negative Effects of Different Doping Agents on Catalytic Activity and Selectivity

Addition or doping with other metal oxides or elements, such as F [21,58], Nb, Sb [59], Cu, Mn, and Ce [60,61,62] is the main way to improve catalyst activity, whereas alkaline and alkaline earth metals (Na, K, Ca, and Mg) [63] deactivate catalysts.
Zhang et al. [21] found that NO conversion was improved by F doping. As shown in Figure 10, V2O5-WO3/TiO2-F1.35 (the ratio of Ti/F was 1.35) exhibited the highest NO removal efficiency (82.8% at 210 °C), owing to the improved interaction of WO3 with TiO2 forming oxygen vacancies and increasing the number of reduced W species (W5+), which play an important role for superoxide ion formation. To further investigate the mechanism, molecular modeling was employed using density functional theory (DFT) and first principles molecular dynamics simulations [58].
Ye et al. [59] investigated the enhanced activity of Nb and Sb doped V2O5-WO3/TiO2 SCR catalysts (V2O5-WO3-Nb2O5/TiO2 and V2O5-WO3-Sb2O5/TiO2) that was assigned to the increase of reactive electrophilic oxygen species on the surface of V2O5-WO3-Nb2O5/TiO2 and V2O5-WO3-Sb2O5/TiO2. Moreover, the catalysts’ deactivation by SO2 and H2O was inhibited through the addition of Nb2O5 and Sb2O5.
Tian et al. [60] prepared Cu, Mn, and Ce doped V2O5-WO3/TiO2 catalysts and studied their performances. The activity was improved because Cu, Mn, and Ce can make the surface of the catalyst moderately acidic, increasing the V4+/V5+ ratio and enhancing the redox performance of the catalyst.
Chen et al. [61] also investigated Ce doped V2O5-WO3/TiO2 catalysts with low vanadium loadings in the NH3-SCR of the NOx, and the optimized catalyst showed an above 90% NOx conversion from 200 to 450 °C. Additionally, the selectivity could be significantly increased at high temperatures, even at a high GHSV (Gas Hourly Space Velocity) of 113,000 h−1, and NOx conversion nearly reached 100% in the temperature range of 250 to 350 °C. According to XPS, analysis cerium was present as Ce3+, causing a charge imbalance and therefore oxygen vacancies’ formation. Such defects increased the amount of chemisorbed oxygen, indicating that more NO can be oxidized to NO2.
The effect of Ce doping of TiO2 support on NH3-SCR activity over V2O5-WO3/CeO2-TiO2 catalyst was investigated by Cheng et al. [62]. The catalytic activity of V2O5-WO3/TiO2 was greatly enhanced by Ce doping, the best NO conversion being achieved at the molar ratio of Ce/Ti = 1/10 (see Figure 11). The crystalline structure also played an important role on the activity; indeed, the catalysts showing predominantly anatase TiO2 showed better catalytic performance than the catalysts with predominantly a fluorite CeO2 structure. Moreover, the Ce additive enhanced the amount of oxygen adsorbed on the surface with beneficial effects on the SCR reaction.
The activity of alkaline and alkaline earth metals (Na, K, Ca, and Mg) doped V2O5–WO3/TiO2 catalysts was investigated by Chen et al. [63]. The activity downward trend is: K > Na > Ca > Mg. Several effects must be taken into account in order to explain these findings: (1) Na and K decrease the quantity and stability of the Brønsted acid sites more than Mg and Ca; (2) the surface chemisorbed oxygen decreased in the same order; (3) the reducibility of vanadium species on the Na and K doped catalysts is lower than that on the Ca and Mg doped catalysts with the same alkali metal oxides loadings; (4) Na and K ions also affect the reduction degree of tungsten species, while no effect was experienced by Ca and Mg; and (5) the poisoning by alkali metals of vanadium species appears to be the main factors for V2O5-WO3/TiO2 catalysts’ deactivation.
In Table 4, the NH3-SCR performances of V2O5-WO3/TiO2 catalysts doped by various elements (F, Nb, Sb, Na, K, Ca, Mg, Cu, Mn, Ce), as discussed above, are summarized. Based on such results, simultaneous doping with F and Ce could be the most effective way to optimize the performance of new V2O5-WO3/TiO2 based catalysts due to the positive roles played individually by F and Ce in enhancing the NOx conversion and selectivity.

3. Deactivation or Poisoning Mechanism

In practical applications of NO-SCR by NH3, the V2O5-WO3/TiO2 catalysts can be poisoned by several species, such as impurities of biodiesel (K, Na, P), urea solution (K, Na, Ca, Mg), and abrasion of the engine (Cr, Cu). As introduced in Section 2.5, alkaline and alkaline earth metals have a negative effect. Following the study of Chen et al. [63], further investigations on alkali ions as well as on other poisoning species have drawn great attention: Hg [64,65,66], alkali metals [66,67,68,69,70], P [69], Cr [69], Cu [69], Mg [69], Ca [69,71], and As [72]. The main poisoning elements and corresponding reasons for catalysts’ deactivation are listed in Table 5.
Kong et al. [64] investigated the HgCl2 deactivation mechanism over V2O5-WO3/TiO2 SCR catalysts. The results showed that HgCl2 caused an agglomeration of active components, the BET surface area decreased significantly, and the reaction of HgCl2 with Brønsted acid sites (V-OH) occurred. From NH3-TPD, NH3 chemisorption, FTIR, and XPS characterization data, it was found that V-OH and V = O bonds were affected by HgCl2. NH3 was adsorbed on new sites (Cl-V-O-H) that were not active for NO SCR, and HgCl2 existed in a stable form bonded to the bridge sites.
Yang et al. [65] evaluated the activity of mercury exposed catalysts by XPS and found that mercury oxidation led to the transformation of the V5+ to V4+ species and consumed the lattice oxygen of catalyst, thus decreasing the activity. In another article, Kong et al. [66] introduced KCl by impregnation of the V2O5-WO3/TiO2 catalysts and added Hg by the dry-gas adsorption method. The catalysts modified with KCl were deactivated due to the reaction of KCl with V-OH acidic groups forming -V-O-K and Cl-V-O-K inactive sites. Hg had a weaker negative effect, and it did not cause catalyst agglomeration and increased the amount of chemisorbed NH3, thus delaying the deactivation caused by KCl. However, Hg reacted with Cl (from KCl), forming HgCl or HgCl2. Then, inactive species, such as -V-O~Hg or -V-O-Hg-Cl, were formed that were able to adsorb NH3 in competition with the active sites.
The deactivation of V2O5-WO3/TiO2 catalyst induced by KCl was investigated systematically via three methods (wet impregnation, solid diffusion, and vapor deposition) by Lei et al. [67]. The deactivation rate occurred in the following order: Vapor deposition >> solid diffusion > wet impregnation, and was mainly attributed to the presence of different forms of V species that reduced the acidity to different extents as previously found by Yu et al. [24].
When KCl was added by wet impregnation, the activity of V2O3 species was only slightly reduced, possibly due to the intimate interaction between K and oxygen in vanadium oxides. On the contrary, by the solid diffusion method, the formation of V2O5-K2S2O7 eutectic occurred, leading to a very small surface area and a too strong oxidation ability, which were the main reasons for the catalyst deactivation.
Xie et al. [68] assessed the deactivation mechanism by investigating a commercial V2O5-WO3/TiO2 catalyst through exposure to the flue gas of a coal-fired power plant by a transient kinetic analysis that focused on the distinction between the deactivation behaviors of adsorption sites and redox sites. The results showed that the alkali element preferentially poisons the active sites associated with vanadium (V5+-OH and/or V5+ = O) rather than the sites associated with Ti and W. Thus, it was concluded that the poisoning of vanadium-linked species caused decreased NH3 adsorption and increased NH3 desorption, negatively affecting the SCR reaction.
The deactivation mechanism of V2O5-WO3/TiO2 ceramic honeycomb catalysts induced by P, K, Na, Mg, Ca, Cr, and Cu was assayed by Klimczak et al. [69]. The catalysts were poisoned by two different methods: (i) wet impregnation of dilute aqueous solution of the corresponding nitrate or ammonium salts; (ii) deposition of the inorganic aerosol particles. The second approach is more realistic as it uses reaction conditions close to the mobile application and it seems to be more harmful. Both impregnated and aerosol deactivated catalysts showed a strong poisoning effect of alkaline metals due to a reduced capacity of absorbed NH3. P, Cr, and Cu moderately poisoned the catalysts because there was competition between the increased N2O production and reduced acid sites. In Figure 12, the NOx conversion curves over V2O5-WO3/TiO2 (400 cpsi monoliths) are compared [69]. Figure 12a shows fresh, hydrothermal aged and K, Na and Ca poisoned samples; Figure 12b displays fresh, hydrothermal aged and Cr and Cu poisoned catalysts. In both cases, a molar throughput of 3 mmol of poisoning was used and added by aerosol.
Deng et al. [70] studied alkali metals’ deactivation process of V2O5-WO3/TiO2 SCR catalyst by using an entrained-flow combustion system. The conclusions can be described as follows, and the activity of the catalyst sample was affected by the fly ash. The order of the poisoning effects of different alkali salts on the activity was Na2CO3 < K2CO3 < Na2SO4 < K2SO4 < NaCl < KCl. Both the decrease of the specific surface area and the masking of V2O5 caused the activity reduction, which can be attributed to mechanical deactivation. More importantly, the deposited particles of alkali salts can both coordinate to the -OH on the surface and interact with the lattice oxygen in the V species, which can be assigned to chemical deactivation. All of them can be ascribed to the reduction of the capacity of ab-NH3.
Li et al. [71] studied the activity of Ca poisoning V2O5-WO3/ TiO2 catalysts based on the modification of activity, structure, reducibility, acidity, and reaction properties. The results revealed that the bulk W species and surface acid sites could be passivated with high contents of CaO, leading to CaWO4 formation. Additionally, the reducibility and surface oxygen species were also inhibited by CaO species.
Kong et al. [72] studied the poisoning of V2O5-WO3/TiO2 catalysts by arsenic (As). In Figure 13a,b, the NOx conversion curves over fresh and poisoned catalysts and the deactivation mechanism are displayed. The main reason for catalytic deactivation originates from the following: As2O3 would be adsorbed on the surface of catalyst, then oxidized into As2O5 by surface chemisorption oxygen. The As2O5 dense layers deposited on the catalyst surface prevent NH3 adsorption and the active sites’ recovery.

4. SO2 and H2O Effects

A discussion about deactivation and poisoning effects on V2O5-WO3-based catalysts cannot exclude the influence of SO2 and/or H2O for realist applications. Indeed, flue gas emissions from power plants always contain SO2 (200−1500 ppm) that can be further oxidized to SO3 by the supported V2O5-WO3/TiO2 catalysts. Moreover, SO3 is corrosive to stainless steel SCR reactor components and readily reacts with ammonia to form ammonium sulfate/bisulfate that deposit on the catalysts, causing deactivation, and on the walls of heat exchangers to reduce their efficiency. Regarding the effects of water, it must be pointed out that significant amounts of moisture, 10% to 30%, are present in the flue gas exiting from industrial combustion processes. Moreover, moisture interacts with the surface of the catalysts and modifies the surface’s active sites and the distribution of Lewis and Brønsted acid sites. However, the role of water has not been extensively investigated, so it deserves further attention.
In the 1990s, preliminary investigations about SO2 effects were reported by Chen et al. [73], who found that SO2 promotes the activity by increasing the Brønsted acidity, in accordance with the assumption that Brønsted acid sites are the active sites for the reaction. A few years later, the same authors [74] investigated SO42−/TiO2 superacid for NO SCR with NH3. Since SO42−/TiO2 is formed on TiO2 under SCR reaction conditions when SO2 is present, the results elucidated the role of TiO2 support (and SO2) in the SCR reaction. XPS and chemisorption measurements along with SCR kinetic results indicated that the reaction takes place by an Eley–Rideal mechanism, on Brønsted acid sites on SO42−/TiO2 surface. The catalytic results showed NO conversion is in the range of 96% to 100% at temperatures between 400 and 425 °C and the good performances were explained considering that at 400 °C, about 70% of the SO42− ions are occupied by ammonia that is active for reactions with NO.
The promoting effect of the sulphate content on the catalytic activity of titania was also reported by Ciambelli et al. [75], who found an enhanced NO conversion without decreased selectivity to N2, which was attributed to the enhanced chemisorption of ammonia.
The coupled effect of SO2 and H2O was investigated by Amiridis et al [76] at 350 °C over V2O5/TiO2 catalysts with variable vanadia loadings. Independently on the vanadia surface coverage, the addition of H2O to the reaction mixture resulted in a decrease of the SCR turnover frequency of approximately 40% to 50% due to a competitive adsorption of H2O on the active vanadia sites. The presence of SO2 in the gas phase resulted in a significant increase of the turnover frequency at low vanadia surface coverage due to the formation of surface sulfate species that in the presence of H2O, act as strong Brønsted acid sites. However, an excess of water had a negative effect in the SCR by slowing the reoxidation of vanadia.
The reported results on the effect of SO2 and water thus far were confirmed by Magnusson et al. [16], who investigated how SO2 affects the NOx activity of a V2O5–WO3/TiO2 urea-SCR catalyst for marine applications. Surprisingly, the authors found that the addition of SO2 in the absence of H2O promoted NOx reduction at 350 °C, while only H2O gave rise to a decrease in the NOx abatement and also an inhibition of the N2O formation. No promotional effect by SO2 was observed at temperatures below 230 °C. Further, long term effects of SO2 and H2O were investigated and the NOx reduction remained stable, also after long term exposure of SO2. Finally, it was concluded that in the presence of both H2O and SO2, the catalyst did not show any deactivation at temperatures above 300 °C and fairly low space velocities (below 12,200 h−1). However, at lower temperatures (250 °C) and/or higher space velocities, the catalytic performance for NOx reduction decreased with time. Therefore, for V2O5-WO3-based catalysts used at low temperatures, the effects of water and sulfur dioxide must be carefully addressed.
A recent detailed investigation about H2O and SO2 effects has been carried out by Lai and Wachs [77]. This is a perspective study addressing the current fundamental understanding and advances on V2O5-WO3/TiO2 catalysts taking into consideration the main parameters affecting the activity, such as the catalyst synthesis, molecular structures of titania supported vanadium and tungsten oxide species, surface acidity, catalytic active sites, surface reaction intermediates, reaction mechanism, rate determining-step, and reaction kinetics.
As remarked in [77] and as well in the references therein, water inhibits the SCR reaction rate at low (∼1%) but does not affect the SCR reaction at high concentrations (>5%). Such an effect arises from (1) competitive adsorption with the reactants (NO or NH3) and/or (2) inhibition of the reaction between NO and adsorbed NH3. H2O increased the concentration of surface NH4+* species on Brønsted acid sites and such surface NH4+* intermediates arise from the reaction of NH3* species with moisture.
Regarding the effect of SOx, the effects of SO2 and SO3 on TiO2 and supported V2O5/TiO2 has received much attention, while less investigation has focused on tungsten catalysts. For V2O5-WO3/TiO2, both surface vanadia and tungsten oxide sites contribute equally to the overall SO2 oxidation reaction to SO3.

5. Conclusions and Perspectives

In summary, a fundamental understanding and recent advances on the respective role of V2O5-WO3/TiO2 catalysts were reviewed in detail, with special concern to the literature published in the last 20 years. It is generally accepted that WO3 increases the Lewis acidity and inhibits the initial sintering of TiO2 as well as helping to form superoxide ions. Concerning V2O5, it increases the surface Brønsted and Lewis acidity and enhances N2 selectivity. The importance of interaction between V2O5 and WO3 was also revealed. In addition, the relationships between the catalytic activity and synthesis methods, supports, composition, and doping agents were discussed as well. It can be pointed out that alkaline and alkaline earth metals (K, Na, Ca, Mg) as well other elements, such as Cr, As, and Hg, deactivate the V2O5-WO3-based catalysts. Especially, alkali metals decrease the ability of NH3 adsorption, which is a disadvantage for low-temperature activity and selectivity.
The specific surface area of titania plays an important role for the dispersion and interaction of V2O5 and WO3 active oxides. In addition, the support composition influences the activity. For example, the addition of SiO2 inhibits the shrinkage of the catalyst’s specific surface area, avoiding the phase transition from anatase to rutile and growth of the TiO2 size.
A Pt doped SiC filter is a promising carrier that not only improves the surface acid sites, but also significantly enlarges the BET surface area, the pore volume of the catalysts, and favors high electron transfer, enhancing the absorption of NH3 and lighting-off NO conversion at low temperatures.
The presence of SO2 and/or H2O on the flue gases must be carefully addressed for practical applications, paying special attention to the experimental conditions that can affect positively or negatively the catalytic performances of V2O5–WO3 based catalysts.
Considering that V2O5-WO3 based catalysts are inactive during the engine cold-start, the enhancement of NO SCR at low temperatures (<200 °C), maintaining high selectivity to N2, is still challenging for a successful application of such systems for automotive and naval usage. Such improvement could be achieved by doping V2O5–WO3 based catalysts with new oxides. According with recent data, MnEuOX, MnSmOX, MnGdOx, MnCeOx represent promising co-catalysts.

Author Contributions

W.Z. and S.Q. prepared the original draft of the manuscript, G.P. and L.F.L. equally contributed to the writing, review and editing.

Acknowledgments

Weidong Zhang is very grateful to China Scholarship Council for supporting two years of study at the Institute for the Study of Nanostructured Materials-Council National Research (ISMN-CNR), 90146, Palermo (Italy). The project TECBIA “Tecnologie a Basso Impatto Ambientale per la produzione di energia sui mezzi navali” (progetto n. F.090041/01/X36 –CUP B98I17000680008)” is acknowledged for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shaddick, G.; Thomas, M.L.; Amini, H.; Broday, D.; Cohen, A.; Frostad, J.; Green, A.; Gumy, S.; Liu, Y.; Martin, R.V. Data integration for the assessment of population exposure to ambient air pollution for global burden of disease assessment. Env. Sci. Technol. 2018, 52, 9069–9078. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, Y.; Zhao, J.; Lee, J.M. Conventional and new materials for selective catalytic reduction (SCR) of NOx. Chem. Cat. Chem. 2018, 10, 1499–1511. [Google Scholar]
  3. Liotta, L.F.; Wu, H.; Pantaleo, G.; Venezia, A.M. Co3O4 nanocrystals and Co3O–MOx binary oxides for CO, CH4 and VOC oxidation at low temperatures: A Review. Catal. Sci. Technol. 2013, 3, 3085–3102. [Google Scholar] [CrossRef]
  4. Roy, S.; Hegde, M.; Madras, G. Catalysis for NOx abatement. Appl. Energy 2009, 86, 2283–2297. [Google Scholar] [CrossRef]
  5. Bergmeijer, P. The International Convention for the Prevention of Pollution from Ships. In Ports as Nodal Points in a Global Transport System; Elsevier: Amsterdam, The Netherlands, 1992; pp. 259–270. [Google Scholar]
  6. Skalska, K.; Miller, J.S.; Ledakowicz, S. Trends in NOx abatement: A review. Sci. Total. Env. 2010, 408, 3976–3989. [Google Scholar] [CrossRef]
  7. Fu, M.; Li, C.; Lu, P.; Qu, L.; Zhang, M.; Zhou, Y.; Yu, M.; Fang, Y. A review on selective catalytic reduction of NOx by supported catalysts at 100–300 °C-catalysts, mechanism, kinetics. Catal. Sci. Technol. 2014, 4, 14–25. [Google Scholar] [CrossRef]
  8. Nakatsuji, T.; Miyamoto, A. Removal technology for nitrogen oxides and sulfur oxides from exhaust gases. Catal. Today 1991, 10, 21–31. [Google Scholar] [CrossRef]
  9. Sorrels, J.L.; Randall, D.D.; Fry, C.R.; Schaffner, K.S. Selective Noncatalytic Reduction; U.S. Environmental Protection Agency: Washington, DC, USA, 2015.
  10. Radojevic, M. Reduction of Nitrogen Oxides in Flue Gases. In Nitrogen, the Confer-Ns; Elsevier: Amsterdam, The Netherlands, 1998; pp. 685–689. [Google Scholar]
  11. Liu, J.; Li, X.; Li, R.; Zhao, Q.; Ke, J.; Xiao, H.; Wang, L.; Liu, S.; Tadé, M.; Wang, S. Facile synthesis of tube-shaped Mn-Ni-Ti solid solution and preferable Langmuir-Hinshelwood mechanism for selective catalytic reduction of NOx by NH3. Appl. Catal. A 2018, 549, 289–301. [Google Scholar] [CrossRef]
  12. Crocker, M.; Shi, C. Non-Thermal Plasma Nox Storage-reduction. NOx Trap Catal. Technol. 2018, 384–406. [Google Scholar]
  13. Zhang, R.; Yang, W.; Luo, N.; Li, P.; Lei, Z.; Chen, B. Low-Temperature NH3-SCR of NO by Lanthanum Manganite Perovskites: Effect of A-/B-Site Substitution and TiO2/CeO2 Support. Appl. Catal. B 2014, 146, 94–104. [Google Scholar] [CrossRef]
  14. Busca, G.; Lietti, L.; Ramis, G.; Berti, F. Chemical and Mechanistic Aspects of the Selective Catalytic Reduction of NOx by Ammonia over Oxide Catalysts: A Review. Appl. Catal. B 1998, 18, 1–36. [Google Scholar] [CrossRef]
  15. Hirata, K.; Masaki, N.; Ueno, H.; Akagawa, H. Development of Urea-SCR System for Heavy-Duty Commercial Vehicles. SAE Trans. 2005, 678–685. [Google Scholar]
  16. Magnusson, M.; Fridell, E.; Ingelsten, H.H. The Influence of Sulfur Dioxide and Water on the Performance of a Marine Scr Catalyst. Appl. Catal. B 2012, 111, 20–26. [Google Scholar] [CrossRef]
  17. Guo, M.; Fu, Z.; Ma, D.; Ji, N.; Song, C.; Liu, Q. A Short Review of Treatment Methods of Marine Diesel Engine Exhaust Gases. Procedia Eng. 2015, 121, 938–943. [Google Scholar] [CrossRef] [Green Version]
  18. Chen, C.; Cao, Y.; Liu, S.; Chen, J.; Jia, W. Review on the Latest Developments in Modified Vanadium-Titanium-Based SCR Catalysts. Chin. J. Catal. 2018, 39, 1347–1365. [Google Scholar] [CrossRef]
  19. Tsuji, K.; Shiraishi, I. Combined desulfurization, denitrification and reduction of air toxics using activated coke. 2. Process applications and performance of activated coke. Fuel 1997, 76, 555–560. [Google Scholar] [CrossRef]
  20. Wu, Z.; Jin, R.; Liu, Y.; Wang, H. Ceria modified MnOx/TiO2 as a superior catalyst for NO reduction with NH3 at low-temperature. Catal. Commun. 2008, 9, 2217–2220. [Google Scholar] [CrossRef]
  21. Zhang, S.; Li, H.; Zhong, Q. Promotional Effect of F-Doped V2O5-WO3/TiO2 Catalyst for NH3-SCR of NO at Low-Temperature. Appl. Catal. A 2012, 435, 156–162. [Google Scholar] [CrossRef]
  22. Aguilar-Romero, M.; Camposeco, R.; Castillo, S.; Marín, J.; Rodríguez-González, V.; García-Serrano, L.A.; Mejía-Centeno, I. Acidity, Surface Species, and Catalytic Activity Study on V2O5-WO3/TiO2 Nanotube Catalysts for Selective NO Reduction by NH3. Fuel 2017, 198, 123–133. [Google Scholar] [CrossRef]
  23. Kompio, P.G.; Brückner, A.; Hipler, F.; Auer, G.; Löffler, E.; Grünert, W. A New View on the Relations between Tungsten and Vanadium in V2O5-WO3/TiO2 Catalysts for the Selective Reduction of NO with NH3. J. Catal. 2012, 286, 237–247. [Google Scholar] [CrossRef]
  24. 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]
  25. Dong, G.-J.; Zhang, Y.-F.; Yuan, Z.; Yang, B. Effect of the pH value of precursor solution on the catalytic performance of V2O5-WO3/TiO2 in the low temperature NH3-SCR of NOx. J. Fuel Chem. Technol. 2014, 42, 1455–1463. [Google Scholar] [CrossRef]
  26. 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 2016, 193, 141–150. [Google Scholar] [CrossRef]
  27. Qi, C.; Bao, W.; Wang, L.; Li, H.; Wu, W. Study of the V2O5-WO3/TiO2 catalyst synthesized from waste catalyst on selective catalytic reduction of NOx by NH3. Catalysts 2017, 7, 110. [Google Scholar] [CrossRef]
  28. Meiqing, S.; Lili, X.; Jianqiang, W.; Chenxu, L.; Wulin, W.; Jun, W.; Yanping, Z. 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]
  29. Djerad, S.; Tifouti, L.; Crocoll, M.; Weisweiler, W. Effect of vanadia and tungsten loadings on the physical and chemical characteristics of V2O5-WO3/TiO2 catalysts. J. Mol. Catal. A Chem. 2004, 208, 257–265. [Google Scholar] [CrossRef]
  30. Reiche, M.; Buergi, T.; Baiker, A.; Scholz, A.; Schnyder, B.; Wokaun, A. Vanadia and Tungsta Grafted on TiO2: Influence of the Grafting Sequence on Structural and Chemical Properties. Appl. Catal. A 2000, 198, 155–169. [Google Scholar] [CrossRef]
  31. Thirupathi, B.; Smirniotis, P.G. Nickel-Doped Mn/TiO2 as an Efficient Catalyst for the Low-Temperature SCR of NO with NH3: Catalytic Evaluation and Characterizations. J. Catal. 2012, 288, 74–83. [Google Scholar] [CrossRef]
  32. Ettireddy, P.R.; Ettireddy, N.; Mamedov, S.; Boolchand, P.; Smirniotis, P.G. Surface Characterization Studies of TiO2 Supported Manganese Oxide Catalysts for Low Temperature SCR of NO with NH3. Appl. Catal. B 2007, 76, 123–134. [Google Scholar] [CrossRef]
  33. Reiche, M.; Hug, P.; Baiker, A. Effect of grafting sequence on the behavior of titania-supported V2O5-WO3 catalysts in the selective reduction of NO by NH3. J. Catal. 2000, 192, 400–411. [Google Scholar] [CrossRef]
  34. 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]
  35. 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]
  36. Wang, C.; Yang, S.; Chang, H.; Peng, Y.; Li, J. Dispersion of tungsten oxide on SCR performance of V2O5-WO3/TiO2: Acidity, surface species and catalytic activity. Chem. Eng. J. 2013, 225, 520–527. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Wang, L.; Li, J.; Zhang, H.; Xu, H.; Xiao, R.; Yang, L. Promotional roles of ZrO2 and WO3 in V2O5-WO3/TiO2-ZrO2 catalysts for NOx reduction by NH3: Catalytic performance, morphology, and reaction mechanism. Chin. J. Catal. 2016, 37, 1918–1930. [Google Scholar] [CrossRef]
  38. Herrmann, J.-M.; Disdier, J. Electrical properties of V2O5-WO3/TiO2 EUROCAT catalysts evidence for redox process in selective catalytic reduction (SCR) deNOx reaction. Catal. Today 2000, 56, 389–401. [Google Scholar] [CrossRef]
  39. Jung, S.M.; Grange, P. Characterization and reactivity of V2O5-WO3 supported on TiO2-SO42− catalyst for the SCR reaction. Appl. Catal. B 2001, 32, 123–131. [Google Scholar] [CrossRef]
  40. 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]
  41. Camposeco, R.; Castillo, S.; Mugica, V.; Mejía-Centeno, I.; Marín, J. Role of V2O5-WO3/H2Ti3O7-nanotube-model catalysts in the enhancement of the catalytic activity for the SCR-NH3 process. Chem. Eng. J. 2014, 242, 313–320. [Google Scholar] [CrossRef]
  42. Mejía-Centeno, I.; Castillo, S.; Camposeco, R.; Marín, J.; García, L.A.; Fuentes, G.A. Activity and selectivity of V2O5/H2Ti3O7, V2O5-WO3/H2Ti3O7 and Al2O3/H2Ti3O7 model catalysts during the SCR-NO with NH3. Chem. Eng. J. 2015, 264, 873–885. [Google Scholar] [CrossRef]
  43. Boudali, L.K.; Ghorbel, A.; Grange, P. Characterization and reactivity of V2O5-WO3 supported on sulfated titanium pillared clay catalysts for the SCR-NO reaction. C.R. Chim. 2009, 12, 779–786. [Google Scholar] [CrossRef]
  44. Najbar, M.; Mizukami, F.; Białas, A.; Camra, J.; Wesełucha-Birczyńska, A.; Izutsu, H.; Góra, A. Evolution of Ti–Sn-Rutile-supported V2O5-WO3 catalyst during its use in nitric oxide reduction by ammonia. Top. Catal. 2000, 11, 131–138. [Google Scholar] [CrossRef]
  45. 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]
  46. Kim, J.H.; Choi, J.H.; Phule, A.D. Development of high performance catalytic filter of V2O5-WO3/TiO2 supported SiC for NOx reduction. Powder Technol. 2018, 327, 282–290. [Google Scholar] [CrossRef]
  47. Choi, J.-H.; Kim, J.-H.; Bak, Y.-C.; Amal, R.; Scott, J. Pt-V2O5-WO3/TiO2 catalysts supported on SiC filter for NO reduction at low temperature. Korean J. Chem. Eng. 2005, 22, 844–851. [Google Scholar] [CrossRef]
  48. Najbar, M.; Banaś, J.; Korchowiec, J.; Białas, A. Competition between NO reduction and NO decomposition over reduced V–W–O catalysts. Catal. Today 2002, 73, 249–254. [Google Scholar] [CrossRef]
  49. Zhao, K.; Han, W.; Tang, Z.; Zhang, G.; Lu, J.; Lu, G.; Zhen, X. Investigation of coating technology and catalytic performance over monolithic V2O5-WO3/TiO2 catalyst for selective catalytic reduction of NOx with NH3. Colloids Surf. A 2016, 503, 53–60. [Google Scholar] [CrossRef]
  50. Shen, B.; Wang, F.; Zhao, B.; Li, Y.; Wang, Y. The behaviors of V2O5-WO3/TiO2 loaded on ceramic surfaces for NH3-SCR. J. Ind. Eng. Chem. 2016, 33, 262–269. [Google Scholar] [CrossRef]
  51. 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]
  52. Dong, G.-J.; Yuan, Z.; Zhang, Y.-F. Preparation and performance of V-W reparation and performance of VW/x(Mn-Ce-Ti)/y(Cu-Ce-Ti)/cordierite catalyst by impregnation method in sequence for SCR reaction with urea. J. Fuel Chem. Technol. 2014, 42, 1093–1101. [Google Scholar] [CrossRef]
  53. Li, Z.; Li, J.; Liu, S.; Ren, X.; Ma, J.; Su, W.; Peng, Y. Ultra-hydrothermal stability of CeO2-WO3/TiO2 for NH3-SCR of NO compared to traditional V2O5-WO3/TiO2 catalyst. Catal. Today 2015, 258, 11–16. [Google Scholar] [CrossRef]
  54. Shi, A.; Wang, X.; Yu, T.; Shen, M. The effect of zirconia additive on the activity and structure stability of V2O5-WO3/TiO2 ammonia SCR catalysts. Appl. Catal. B 2011, 106, 359–369. [Google Scholar] [CrossRef]
  55. Tao, P.; Sun, M.; Qu, S.; Song, C.; Li, C.; Yin, Y.; Cheng, M. Effects of V2O5 and WO3 loadings on the catalytic performance of V2O5-WO3/TiO2 catalyst for SCR of NO with NH3. Glob. Nest J. 2017, 19, 160–166. [Google Scholar]
  56. Lietti, L.; Nova, I.; Forzatti, P. Selective catalytic reduction (SCR) of NO by NH3 over TiO2-supported V2O5-WO3 and V2O5-MoO3 catalysts. Top. Catal. 2000, 11, 111–122. [Google Scholar] [CrossRef]
  57. Xu, T.; Wu, X.; Gao, Y.; Lin, Q.; Hu, J.; Weng, D. Comparative study on sulfur poisoning of V2O5-Sb2O3/TiO2 and V2O5-WO3/TiO2 monolithic catalysts for low-temperature NH3-SCR. Catal. Commun. 2017, 93, 33–36. [Google Scholar] [CrossRef]
  58. Zhang, S.; Yang, X.; Zhong, Q. Cluster molecular modeling of strong interaction for F-doped V2O5-WO3/TiO2 supported catalyst. J. Fluor. Chem. 2013, 153, 26–32. [Google Scholar] [CrossRef]
  59. Ye, D.; Qu, R.; Zheng, C.; Cen, K.; Gao, X. Mechanistic investigation of enhanced reactivity of NH4HSO4 and NO on Nb-and Sb-doped VW/Ti SCR catalysts. Appl. Catal. A 2018, 549, 310–319. [Google Scholar] [CrossRef]
  60. Tian, X.; Xiao, Y.; Zhou, P.; Zhang, W.; Luo, X. Investigation on performance of V2O5-WO3/TiO2-cordierite catalyst modified with Cu, Mn and Ce for urea-SCR of NO. Mater. Res. Innov. 2014, 18, S2-202–S2-206. [Google Scholar] [CrossRef]
  61. Chen, L.; Li, J.; Ge, M. Promotional effect of Ce-doped V2O5-WO3/TiO2 with low vanadium loadings for selective catalytic reduction of NOx by NH3. J. Phys. Chem. C 2009, 113, 21177–21184. [Google Scholar] [CrossRef]
  62. Cheng, K.; Liu, J.; Zhang, T.; Li, J.; Zhao, Z.; Wei, Y.; Jiang, G.; Duan, A. Effect of Ce doping of TiO2 support on NH3-SCR activity over V2O5-WO3/CeO2-TiO2 catalyst. J. Env. Sci. 2014, 26, 2106–2113. [Google Scholar] [CrossRef]
  63. 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]
  64. Kong, M.; Liu, Q.; Jiang, L.; Guo, F.; Ren, S.; Yao, L.; Yang, J. Property influence and poisoning mechanism of HgCl2 on V2O5-WO3/TiO2 SCR-deNOx catalysts. Catal. Commun. 2016, 85, 34–38. [Google Scholar] [CrossRef]
  65. Yang, J.; Yang, Q.; Sun, J.; Liu, Q.; Zhao, D.; Gao, W.; Liu, L. Effects of mercury oxidation on V2O5-WO3/TiO2 catalyst properties in NH3-SCR process. Catal. Commun. 2015, 59, 78–82. [Google Scholar] [CrossRef]
  66. 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]
  67. Lei, T.; Li, Q.; Chen, S.; Liu, Z.; Liu, Q. KCl-induced deactivation of V2O5-WO3/TiO2 catalyst during selective catalytic reduction of NO by NH3: Comparison of poisoning methods. Chem. Eng. J. 2016, 296, 1–10. [Google Scholar] [CrossRef]
  68. Xie, X.; Lu, J.; Hums, E.; Huang, Q.; Lu, Z. Study on the deactivation of V2O5-WO3/TiO2 selective catalytic reduction catalysts through transient kinetics. Energy Fuels 2015, 29, 3890–3896. [Google Scholar] [CrossRef]
  69. 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 2010, 95, 39–47. [Google Scholar] [CrossRef]
  70. Deng, L.; Liu, X.; Cao, P.; Zhao, Y.; Du, Y.; Wang, C.; Che, D. A study on deactivation of V2O5-WO3/TiO2 SCR catalyst by alkali metals during entrained-flow combustion. J. Energy Inst. 2017, 90, 743–751. [Google Scholar] [CrossRef]
  71. Li, X.; Li, X.-S.; Chen, J.; Li, J.; Hao, J. An efficient novel regeneration method for Ca-poisoning V2O5-WO3/TiO2 catalyst. Catal. Commun. 2016, 87, 45–48. [Google Scholar] [CrossRef]
  72. Kong, M.; Liu, Q.; Wang, X.; Ren, S.; Yang, J.; Zhao, D.; Xi, W.; Yao, L. Performance impact and poisoning mechanism of arsenic over commercial V2O5-WO3/TiO2 SCR catalyst. Catal. Commun. 2015, 72, 121–126. [Google Scholar] [CrossRef]
  73. Chen, J.; Yang, R. Mechanism of poisoning of the V2O5/TiO2 Catalyst for the reduction of NO by NH3. J. Catal. 1990, 125, 411–420. [Google Scholar] [CrossRef]
  74. Chen, J.; Yang, R. Selective Catalytic Reduction of NO with NH3 on SO4−2/TiO2 Superacid Catalyst. J. Catal. 1993, 139, 277–288. [Google Scholar] [CrossRef]
  75. Ciambelli, P.; Fortuna, M.; Sannino, D.; Baldacci, A. The Influence of Sulphate on the Catalytic Properties of V2O5-TiO2 and WO3-TiO2 in the Reduction of Nitric Oxide with Ammonia. Catal. Today 1996, 29, 161–164. [Google Scholar] [CrossRef]
  76. Amiridis, M.D.; Wachs, I.E.; Deo, G.; Jehng, J.-M. Reactivity of V2O5 catalysts for the Selective Catalytic Reduction of NO by NH3: Influence of Vanadia Loading, H2O, and SO2. J. Catal. 1996, 161, 247–253. [Google Scholar] [CrossRef]
  77. Lai, J.-K.; Wachs, I.E. A Perspective on the Selective Catalytic Reduction (SCR) of NO with NH3 by Supported V2O5-WO3/TiO2 Catalysts. ACS Catal. 2018, 8, 6537–6551. [Google Scholar] [CrossRef]
Figure 1. NH3-SCR (Selective Catalytic Reduction) activities of the fresh and poisoned (by KCl) V2O5-WO3/TiO2 catalysts prepared by wet and dry impregnation methods and denoted as VWT (V2O5-WO3/TiO2) (W), VWT (D), VWT (WK), and VWT (DK), respectively. (Reprinted with permission from [24]. Copyright (2013) Elsevier).
Figure 1. NH3-SCR (Selective Catalytic Reduction) activities of the fresh and poisoned (by KCl) V2O5-WO3/TiO2 catalysts prepared by wet and dry impregnation methods and denoted as VWT (V2O5-WO3/TiO2) (W), VWT (D), VWT (WK), and VWT (DK), respectively. (Reprinted with permission from [24]. Copyright (2013) Elsevier).
Catalysts 09 00527 g001
Figure 2. NH3-SCR performance over various catalysts as a function of temperature (a), (c), (d) NOx conversion; (b) N2O concentration. (Reprinted with permission from [28]. Copyright (2016) Elsevier).
Figure 2. NH3-SCR performance over various catalysts as a function of temperature (a), (c), (d) NOx conversion; (b) N2O concentration. (Reprinted with permission from [28]. Copyright (2016) Elsevier).
Catalysts 09 00527 g002
Figure 3. NOx conversion (a) and N2 selectivity (b) vs. temperature over 3 wt.%V2O5–9 wt.% WO3/TiO2 and 8 wt % V2O5–9 wt.% WO3/TiO2 catalysts. (Reprinted with permission from [29]. Copyright (2004) Elsevier).
Figure 3. NOx conversion (a) and N2 selectivity (b) vs. temperature over 3 wt.%V2O5–9 wt.% WO3/TiO2 and 8 wt % V2O5–9 wt.% WO3/TiO2 catalysts. (Reprinted with permission from [29]. Copyright (2004) Elsevier).
Catalysts 09 00527 g003
Figure 4. NOx conversions over different catalysts with (a) and without (b) O2. (Reprinted with permission from [37]. Copyright (2016) Elsevier).
Figure 4. NOx conversions over different catalysts with (a) and without (b) O2. (Reprinted with permission from [37]. Copyright (2016) Elsevier).
Catalysts 09 00527 g004
Figure 5. SCR activity: (a) TiO2-SO42−, (b) V2O5/TiO2-SO42−, (c) WO3/TiO2-SO42−, and (d) V2O5-WO3/TiO2–SO42− (Reprinted with permission from [39]. Copyright (2001) Elsevier).
Figure 5. SCR activity: (a) TiO2-SO42−, (b) V2O5/TiO2-SO42−, (c) WO3/TiO2-SO42−, and (d) V2O5-WO3/TiO2–SO42− (Reprinted with permission from [39]. Copyright (2001) Elsevier).
Catalysts 09 00527 g005
Figure 6. (a) NO conversions over different catalysts; (c) NO and NH3 conversions over 5 wt %V2O5-5 wt %WO3/TAN catalyst; NO conversions in the presents of H2O and SO2 (b,d) over the above mentioned catalysts. (Figure 6a,b reprinted with permission from [41]. Copyright (2014) Elsevier. Figure 6c,d reprinted with permission from [42]. Copyright (2015) Elsevier).
Figure 6. (a) NO conversions over different catalysts; (c) NO and NH3 conversions over 5 wt %V2O5-5 wt %WO3/TAN catalyst; NO conversions in the presents of H2O and SO2 (b,d) over the above mentioned catalysts. (Figure 6a,b reprinted with permission from [41]. Copyright (2014) Elsevier. Figure 6c,d reprinted with permission from [42]. Copyright (2015) Elsevier).
Catalysts 09 00527 g006
Figure 7. NO conversions (a) and N2O concentration (b) over V2O5–WO3/TiO2 supported-SiC catalysts, NO conversion over Pt–V2O5–WO3/TiO2 supported-SiC catalyst containing different Pt loadings (c) and containing different TiO2 loadings (d). (Figure 7a,b Reprinted with permission from [46]. Copyright (2018) Elsevier. Figure 7c,d Reprinted with permission from [47]. Copyright (2005) Springer Link).
Figure 7. NO conversions (a) and N2O concentration (b) over V2O5–WO3/TiO2 supported-SiC catalysts, NO conversion over Pt–V2O5–WO3/TiO2 supported-SiC catalyst containing different Pt loadings (c) and containing different TiO2 loadings (d). (Figure 7a,b Reprinted with permission from [46]. Copyright (2018) Elsevier. Figure 7c,d Reprinted with permission from [47]. Copyright (2005) Springer Link).
Catalysts 09 00527 g007
Figure 8. Effects of the TiO2 gel calcination temperatures on NO conversions over V2O5-WO3/TiO2 catalysts (a) and NO conversions of K poisoned catalysts (b). NO conversion over V2O5-WO3/TiO2 and V2O5-WO3/TiO2-SiO2 catalyst (c) and corresponding selectivity (d). (Figure 8a,b Reprinted with permission from [50]. Copyright (2018) Elsevier. Figure 8c,d Reprinted with permission from [51]. Copyright (2005) Springer Link).
Figure 8. Effects of the TiO2 gel calcination temperatures on NO conversions over V2O5-WO3/TiO2 catalysts (a) and NO conversions of K poisoned catalysts (b). NO conversion over V2O5-WO3/TiO2 and V2O5-WO3/TiO2-SiO2 catalyst (c) and corresponding selectivity (d). (Figure 8a,b Reprinted with permission from [50]. Copyright (2018) Elsevier. Figure 8c,d Reprinted with permission from [51]. Copyright (2005) Springer Link).
Catalysts 09 00527 g008
Figure 9. Results of catalytic activity runs in the SCR reaction (NO conversion and N2 selectivity vs. temperature) performed over: (A) V2O5(x)-WO3(9)/TiO2 samples and (B) V2O5(x)-MoO3(6)/TiO2. (a) V2O5n = 0% w/w; (b) V2O5 = 0.8% w/w; and (c) V2O5 = 1.5% w/w. (Reprinted with permission from [56]. Copyright (2018) Elsevier).
Figure 9. Results of catalytic activity runs in the SCR reaction (NO conversion and N2 selectivity vs. temperature) performed over: (A) V2O5(x)-WO3(9)/TiO2 samples and (B) V2O5(x)-MoO3(6)/TiO2. (a) V2O5n = 0% w/w; (b) V2O5 = 0.8% w/w; and (c) V2O5 = 1.5% w/w. (Reprinted with permission from [56]. Copyright (2018) Elsevier).
Catalysts 09 00527 g009
Figure 10. NO conversion vs. temperature performed over V2O5-WO3/TiO2-F1.35 and a series of catalysts. (Reprinted with permission from [21]. Copyright (2012) Elsevier).
Figure 10. NO conversion vs. temperature performed over V2O5-WO3/TiO2-F1.35 and a series of catalysts. (Reprinted with permission from [21]. Copyright (2012) Elsevier).
Catalysts 09 00527 g010
Figure 11. NOx conversion over Ce doped V2O5-WO3/TiO2 catalysts with various Ce loadings (Reprinted with permission from [62]. Copyright (2014) Elsevier).
Figure 11. NOx conversion over Ce doped V2O5-WO3/TiO2 catalysts with various Ce loadings (Reprinted with permission from [62]. Copyright (2014) Elsevier).
Catalysts 09 00527 g011
Figure 12. NOx conversion over V2O5-WO3/TiO2 (400 cpsi monoliths) as (a) fresh, hydrothermal aged and K, Na and Ca poisoned samples; (b) fresh, hydrothermal aged and Cr and Cu poisoned catalysts. Conditions: 1000 vppm NO, 1000 vppm CO, 5 vol.% CO2, 8 vol.% O2, 5 vol.% H2O and balance N2 at NH3 slip of 25 vppm and 50,000 h−1. (Reprinted with permission from [69]. Copyright (2010) Elsevier).
Figure 12. NOx conversion over V2O5-WO3/TiO2 (400 cpsi monoliths) as (a) fresh, hydrothermal aged and K, Na and Ca poisoned samples; (b) fresh, hydrothermal aged and Cr and Cu poisoned catalysts. Conditions: 1000 vppm NO, 1000 vppm CO, 5 vol.% CO2, 8 vol.% O2, 5 vol.% H2O and balance N2 at NH3 slip of 25 vppm and 50,000 h−1. (Reprinted with permission from [69]. Copyright (2010) Elsevier).
Catalysts 09 00527 g012
Figure 13. NOx conversion over fresh and poisoned V2O5–WO3/TiO2 catalysts (a); picture for deactivation mechanism (b). (Reprinted with permission from [72]. Copyright (2015) Elsevier).
Figure 13. NOx conversion over fresh and poisoned V2O5–WO3/TiO2 catalysts (a); picture for deactivation mechanism (b). (Reprinted with permission from [72]. Copyright (2015) Elsevier).
Catalysts 09 00527 g013
Table 1. Summary of the synthesis methods of V2O5-WO3-based SCR catalysts.
Table 1. Summary of the synthesis methods of V2O5-WO3-based SCR catalysts.
CatalystsSyntheses Method (Cal. Tem./°C)The Key Facts to the ActivityRef.
V2O5-WO3/TiO2Wet impregnation (450 °C)Compared with dry-impregnated catalyst, wet-impregnated catalyst increases the intensity of Brønsted acid sites and the oxidation of the ab-NH3.[24]
V2O5-WO3/TiO2Dry impregnation (450 °C)
V2O5-WO3/TiO2Wet impregnation (500 °C)By increasing the precursor solution acidity, more polymeric vanadium species are formed on the catalyst surface; the ratio of V4+(3+)/V5+, surface acidity, and quantity of active sites are increased, whereas the activity of the catalyst is largely improved.[25]
V2O5-WO3/TiO2Co-precipitation (500 °C)The enhanced performance of co-precipitated compared to the impregnated catalysts was associated with the co-precipitated V2O5-WO3/TiO2 catalysts possess new surface: O = WO4 (enhanced the adsorption of NH3) and redox surface mono-oxo O = VO3 sites, because of the surface defects of the TiO2.[26]
V2O5-WO3/TiO2Incipient wetness impregnation (500 °C)
V2O5-WO3/TiO2Oxalic acid leaching and impregnation (600 °C)Oxalic acid leaching is an effective way to reduce the impurities and increase reducibility of the recovered catalyst.[27]
Ce-V2O5-WO3/TiO2Deposition-precipitation (500 °C)The higher NOx conversion at low temperature has origin from surface Ce species and can be assigned to more active Lewis acid sites, the weakly adsorbed NO2 and monodentate nitrate.[28]
Ce-V2O5-WO3/TiO2Impregnation method (500 °C)
V2O5-WO3/TiO2Sol–gel method (400–800 °C)Well dispersed and isolated vanadium oxide species were found to be weakly active for the SCR reaction but with a high selectivity to N2.[29]
V2O5-WO3/TiO2Grafting (300 °C)Grafting mode and metal oxides loadings were significantly affecting the strength of interaction of V2O5 and WO3.[30]
Table 2. Summary of the supports used for the V2O5-WO3-based SCR catalysts.
Table 2. Summary of the supports used for the V2O5-WO3-based SCR catalysts.
CatalystsCarriesNOx Conversion (Tem./°C)Surface Active Componentsthe Key Facts to the ActivityRef.
V2O5-WO3/TiO2TiO290–100% (225–375 °C)V2O5 and WO3The well dispersed and isolated vanadium oxide species were found to be weakly active for the SCR reaction but with a high selectivity to N2.[29]
V2O5-WO3/TiO2TiO2~V and W species do not exist independently on the titania surface (V-O-W connectivity were present on the V2O5-WO3/TiO2 catalysts), which results in a higher activity.[30]
V2O5-WO3/TiO2TiO2~[33]
V2O5-WO3/TiO2TiO290% (>245 °C)[34]
V2O5-WO3/TiO2TiO280% (320 °C)The promotional effect of tungsten may originate from the follows: the influence of the V species or the dispersion of V species ensembles[23]
V2O5-WO3/TiO2TiO290–100% (250–360 °C)WO3 species play an important role to form superoxide ions (O2), which improved the activity.[35]
V2O5-WO3/TiO2-ZrO2TiO2-ZrO290–100% (290–370 °C)V2O5, WO3 and ZrO2ZrO2 mainly promotional role was to change the path-way of NOx reduction and surface acidity of catalyst.[37]
V2O5-WO3/TiO2-SO42-TiO2-SO42-90–98% (260–480 °C)V2O5, WO3 and SO42-Sulfate stability strongly depends on the loaded metal oxides in addition the reduction property of V2O5 also changed due to the V-O-W species.[39]
V2O5-WO3/Ti0.5Sn0.5O2Ti0.5Sn0.5O290% (>300 °C)V2O5, WO3 and TiSnO2Activity was closely correlated to the amounts of the Brønsted acid sites: they were in direct proportion to highly dispersed WO3. However, too much WO3 will cover the active sites and lead to the decrease of the activities; About the reduction temperature, which becomes higher due to the V-O-W species. [40]
V2O5-WO3/ H2Ti3O7H2Ti3O790–98% (325–450 °C)V2O5, WO3 and H2Ti3O7The higher activity can be explained by the following factors: (1) The excellent physics properties that titanic acid nanotubes possess tubular structure, good thermal stability (460 °C) and high specific surface area (314 m2/g). (2) TAN as carriers not only significantly increase Brønsted acid sites that light-off conversion at low temperatures, but also increase Lewis acid sites that promote the NO reduction at high temperature.[41,42]
WO3-V2O5/STi-PILCSTi-PILC90–100% (250–400 °C)~Sulfate species seems to play a more important role for NO removal activity than WO3.[43]
WO3-V2O5/Ti–Sn-rutileTi–Sn-rutile90–100% (>230 °C)~W species mainly responsible for the oxidation-induced outward vanadium segregation in the vanadia-like species, which was play an important role to improve activity.[44]
WO3-V2O5/TiO2-NTTiO2-NT90–92% (320–380°C)V2O5, WO3 and H2Ti3O7Morphology of the nanotubes and the metal loading (increased surface Brønsted and Lewis acidity sites) play an important role for the catalyst’s activity[22]
V2O5-WO3-TiO2/cordieriteCordierite90–98%(300–500 °C)V2O5, WO3 and SiO2For acid pre-treatment of cordierite, which enlarged the BET surface area and pore volume of the catalysts significantly; SiO2 shifted the V valence from V5+ to V4+ and increased chemisorbed oxygen. All of them enhanced the activity.[45]
V2O5-WO3-TiO2/SiCSiC90–99%(210–360 °C)V2O5, WO3 and TiO2Smaller particle size and better dispersion of catalyst coating solution that is helpful to the catalyst go into the inside of pores of the SiC filters to improve the catalytic activity by increasing the BET surface area and exposing more active sites.[46]
Pt-V2O5-WO3-TiO2/SiCSiC95–100% (170–250 °C)V2O5, WO3 and TiO2Pt promoted catalytic activity at low temperatures but increased ammonia oxidation properties. The promotional effect was believed to result from a high electron transfer of Pt-V2O5-WO3-TiO2/SiC.[47]
V2O5-WO3-Ti(Sn)O2/monolithCr–Al steel monolith95–100% (150–200 °C)V2O5, WO3 and TiO2Increase the pre-reduction temperature was in favor of improving the activity and stability in direct NO decomposition due to the tungsten cations were substituted by vanadium.[48]
V2O5-WO3-TiO2/monolithMonolith90–92% (420–450 °C)V2O5, WO3 and TiO2It was important to facilitate the NH3-SCR reaction that XPS confirmed the co-existence of V5+ and V4+ and the BET surface area was significantly improved due to the pore structure.[49]
Table 3. The relationship between composition and the activity of the V2O5-WO3-based catalysts.
Table 3. The relationship between composition and the activity of the V2O5-WO3-based catalysts.
CatalystsNOx Conversion (Tem./°C)the Key Factors for the ActivityRef.
V2O5-WO3/TiO2-SiO290–98% (280–410 °C)The introduction of SiO2 significantly increased the catalysts acidity and leaded to more V and W species were exposure on the surface of catalyst by forming Si–OK that was the result of the reaction between enriched K and Si.[50]
V2O5-WO3/TiO242–70% (280–410 °C)
V2O5-WO3/TiO2-SiO290–98% (500–725 °C)The addition of SiO2 on the one hand, inhibited the shrinkage of catalyst BET surface area due to the phase transition from anatase to rutile and the growth of TiO2 size. On the other hands, the W species remained highly dispersion.[51]
V2O5-WO3/TiO2<58% (500–725 °C)
V2O5-WO3/TiO2<58% (500–725 °C)The higher activity origin from high V4+/V5+ ratio and large amount of surface chemisorbed oxygen as well as BET surface area caused by the M and C composites.[52]
V2O5-WO3/C/ ceramics52–99% (200–50 °C)
V2O5-WO3/C/M/ ceramics78–99% (200–550 °C)
V2O5-WO3/ ceramics10–65% (200–550 °C)
CeO2-WO3/TiO290–99% (250–550 °C)The higher activity of CeO2-WO3/TiO2 is responsible for the well dispersion of CeO2 and W species; Ce2(WO4)3 prevent the W species crystallization and TiO2 sintering and phase transition; The existence of Ce4+-Ce3+ couple that provide more oxygen vacancies and improve the surface Lewis acid of catalyst.[53]
V2O5-WO3/TiO290–95% (275–350 °C)
1.5%V2O5-1%WO3/TiO296–100% (100–550 °C)Catalytic activities could improve by increasing V2O5 loading less than 2 wt%. However, when V2O5 loading exceeds 2 wt%, the activity begins to decline because high V2O5 loading on TiO2 speeds up the phase transition from anatase to rutile. On other hands, NO conversion also can be significantly improved with the increase of moderate WO3 loadings; however, too much WO3 will inhibit the formation of superoxide ions.[55]
1.5%V2O5-6%WO3/TiO270–100% (100–550 °C)
0.5%V2O5-1%WO3/TiO270–100% (100–550 °C)
0.5%V2O5-6%WO3/TiO270–100% (100–550 °C)
V2O5-WO3/TiO290–98% (550–725 °C)WO3 and MoO3 could promote the activity not only due to the special structural, but also to the unique chemical property; Compared with the commercial V2O5-WO3/TiO2 catalyst, V2O5- MoO3/TiO2 catalyst was more active, but less selective.[56]
V2O5-MoO3/TiO290–100% (500–725 °C)
V2O5-Sb2O3/TiO290–100% (225–380 °C)Sb2O3-based catalyst showed a higher NO conversion and broader temperature windows as well as higher SO2 and H2O resistance. The excellent performance was suppressed by the extent of NH4HSO4 accumulated on the catalyst surface[57]
V2O5-WO3/TiO290–94% (275–350 °C)
Table 4. NH3-SCR performances of V2O5-WO3/TiO2 catalysts doped by different agents: positive or negative effects.
Table 4. NH3-SCR performances of V2O5-WO3/TiO2 catalysts doped by different agents: positive or negative effects.
Doping-AgentsCo/Cd (Cc)Reaction ConditionsDeactivated/Recovered Activity (concentration of H2O and SO2)Key Factors for Enhancing SCR EfficiencyRef.
F52%/82.8% (+57%)[NO] = [NH3] = 500 ppm, [O2] = 5 Vol % and N2 in balance, GHSV = 43,000 h−1 at 240 °C62%/99% ([H2O] = [SO2] = 300)Improve the interaction of WO3 with TiO2 by oxygen vacancies;
Increase the number of the reduced W species (W5+).
[21,58]
Na100%/20% (−80%)[NO] = [NH3] = 500 ppm, [O2] = 3 Vol % and N2 in balance, GHSV = 70,000 h−1 at 450 °C~Decrease the quantity and stability of the Brønsted acid sites;
Reduced catalysts surface chemisorbed oxygen;
Decrease the reducibility of vanadium and tungsten species.
[63]
K100%/30% (−70%)
Ca100%/53% (−47%)
Mg100%/95% (−5%)
Cu72%/98% (+36.1%)exhaust gas and [NO] = 1200 ppm
[O2] = 2 Vol % and N2 in balance, GHSV = 10,800 h−1 at 550 °C
~Form moderate acidity and improve the V4+/V5+ ratio on the catalyst surface;
Increase the redox performance of the catalyst.
[60]
Mn72%/97% (+34.7%)
Ce52%/95% (+82.6%)[NO] = [NH3] = 500 ppm, [O2] = 3 Vol % and N2 in balance, GHSV = 28,000 h−1 at 200 °C95%/100% (100 ppm SO2 and 10% H2O)Increase chemisorbed oxygen;
Provide stronger and more active Brønsted acid centers
[60,61,62]
*Note: 1. Co and Cd represent the original and doped-catalyst’s maximum NO conversion. 2. Cc reflect the gradient of NO conversion and it was calculated by the formula: C c = C d C o C o × 100 % .
Table 5. Mainly deactivating species and corresponding key factors over V2O5-WO3/TiO2 catalysts.
Table 5. Mainly deactivating species and corresponding key factors over V2O5-WO3/TiO2 catalysts.
Deactivating SpeciesKey FactorsRef.
HgCl2Reduce the Brønsted acid sites (V-OH) but produce new NH3 adsorption sites (Cl-V-O-H).[64]
HgTransform V5+ species to V4+ species and consumed the lattice oxygen.[65]
KClKCl could react with V-OH, leading the active sites for NH3 absorption inactive.[66]
HgThe gaseous Hg adsorbed on the vanadia sites reduce the active sites.
KCl and HgThere is the competition for active sites that are partially reduced by gaseous Hg, while others can be increased by delaying the deactivation caused by KCl.
KClLowering the acidity at different extent by different forms of V species; NH3 adsorption temperature allows to evaluate the acid sites.[67]
Alkali metalDecrease the ability of NH3 adsorption but increase in the NH3 desorption rate.[68]
P, Cr and CuModerate poisoning of the catalysts due to competition between increase the N2O production and lowering the number of acid sites.[69]
K, Na, Mg and CaReduced capacity of ab-NH3.[69,70]
CaCan passivate surface acid sites of fresh and bulk W species.[71]
AsAs2O5 dense layers derived from the ab-As2O3 oxidized of by chemisorption oxygen on catalyst surface prevent NH3 adsorption and active sites recovery.[72]

Share and Cite

MDPI and ACS Style

Zhang, W.; Qi, S.; Pantaleo, G.; Liotta, L.F. WO3–V2O5 Active Oxides for NOx SCR by NH3: Preparation Methods, Catalysts’ Composition, and Deactivation Mechanism—A Review. Catalysts 2019, 9, 527. https://doi.org/10.3390/catal9060527

AMA Style

Zhang W, Qi S, Pantaleo G, Liotta LF. WO3–V2O5 Active Oxides for NOx SCR by NH3: Preparation Methods, Catalysts’ Composition, and Deactivation Mechanism—A Review. Catalysts. 2019; 9(6):527. https://doi.org/10.3390/catal9060527

Chicago/Turabian Style

Zhang, Weidong, Shuhua Qi, Giuseppe Pantaleo, and Leonarda Francesca Liotta. 2019. "WO3–V2O5 Active Oxides for NOx SCR by NH3: Preparation Methods, Catalysts’ Composition, and Deactivation Mechanism—A Review" Catalysts 9, no. 6: 527. https://doi.org/10.3390/catal9060527

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