Next Article in Journal
Exclusive Papers of the Editorial Board Members and Topical Advisory Panel Members of Catalysts in Section “Catalysis in Organic and Polymer Chemistry”
Previous Article in Journal
Impact of Co-Fed Hydrogen on High Conversion Propylene Aromatization on H-ZSM-5 and Ga/H-ZSM-5
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on Novel SCR Catalysts for Denitration of High Concentrated Nitrogen Oxides and Their Reaction Mechanisms

by
Bo Yu
1,
Xingyu Liu
1,
Shufeng Wu
2,
Heng Yang
1,
Shuran Zhou
1,
Li Yang
1 and
Fang Liu
1,*
1
School of Low-Carbon Energy and Power Engineering, China University of Mining and Technology, Xuzhou 221116, China
2
Lanzhou Petrochemical Research Center, Petrochemical Research Institute, PetroChina, Lanzhou 730060, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(7), 406; https://doi.org/10.3390/catal14070406
Submission received: 17 May 2024 / Revised: 22 June 2024 / Accepted: 25 June 2024 / Published: 27 June 2024
(This article belongs to the Section Environmental Catalysis)

Abstract

:
With the rapid development of industrialization, the emission of nitrogen oxides (NOx) has become a global environmental issue. Uranium is the primary fuel used in nuclear power generation. However, the production of uranium, typically based on the uranyl nitrate method, usually generates large amounts of nitrogen oxides, particularly NO2, with concentrations in the exhaust gas exceeding 10,000 ppm. High concentrations of nitrogen dioxide are also produced during silver electrolysis processing and the treatment of waste electrolyte solutions. Traditional V-W/TiO2 NH3-SCR catalysts typically exhibit high catalytic activity at temperatures ranging from 300 to 400 °C, under conditions of low NOx concentrations and high gas hourly space velocity. However, their performance is not satisfying when reducing high concentrations of NO2. This study aims to optimize the traditional V-W/TiO2 catalysts to enhance their catalytic activity under conditions of high NO2 concentrations (10,000 ppm) and a wide temperature range (200–400 °C). On the basis of 3 wt% Mo/TiO2, various loadings of V2O5 were selected, and their catalytic activities were tested. Subsequently, the optimal ratios of active component vanadium and additive molybdenum were explored. Simultaneously, doping with WO3 for modification was selected in the V-Mo/TiO2 catalyst, followed by activity testing under the same conditions. The results show that: the NOx conversion rates of all five catalysts increase with temperature at range of 200–400 °C. Excessive loading of MoO3 decreased the catalytic performance, with 5 wt% being the optimal loading. The addition of WO3 significantly enhanced the low-temperature activity of the catalysts. When the loadings of WO3 and MoO3 were both 3 wt%, the catalyst exhibited the best denitrification performance, achieving a NOx conversion rate of 98.8% at 250 °C. This catalyst demonstrates excellent catalytic activity in reducing very high concentration (10,000 ppm) NO2, at a wider temperature range, expanding the temperature range by 50% compared to conventional SCR catalysts. Characterization techniques including BET, XRD, XPS, H2-TPR, and NH3-TPD were employed to further study the evolution of the catalyst, and the promotional mechanisms are explored. The results revealed that the proportion of chemisorbed oxygen (Oα) increased in the WO3-modified catalyst, exhibiting lower V reduction temperatures, which are favorable for low-temperature denitrification activity. NH3-TPD experiments showed that compared to MoOx species, surface WOx species could provide more acidic sites, resulting in stronger surface acidity of the catalyst.

Graphical Abstract

1. Introduction

Nitrogen oxides (NOx) are primarily produced through industrial processes and combustion, including high concentrations of NOx generated during metal processing. Uranium oxide plays a key role in the production of fuel [1]. The production of uranium trioxide from uranyl nitrate is illustrated in Equation (1) [2], a process with low flue gas temperatures and high NO2 concentrations. Furthermore, during the manufacture of silver electrolytes, a vigorous reaction occurs between silver powder and nitric acid, resulting in the generation of a large amount of nitrogen oxide gases. Additionally, when treating waste electrolytes from silver electrolysis and silver powder washing solution, significant amounts of nitrogen oxides are also produced during the heating process of mixing these two solutions, with concentrations reaching up to 40,000 mg/m3 [3]. Nitrogen oxides contribute to acid rain, photochemical smog, ozone depletion, and greenhouse effects. When these gases react with other compounds in the atmosphere, they form harmful substances such as ozone and fine particulate matter, which directly impact air quality and have significant implications for human health and climate change [4,5,6].
UO2(NO3)2 → UO3 + xN2O + yNO2 + O2
The conventional methods for treating nitrogen oxides (NOx) in flue gas include adsorption, wet scrubbing, selective catalytic reduction (SCR), selective non-catalytic reduction (SNCR), electron beam methods, plasma methods, etc. [7]. Due to their high efficiency, selectivity, and economic viability, selective catalytic reduction (SCR) and selective non-catalytic reduction (SNCR) are considered two effective approaches for reducing NOx emissions [8,9]. In order to reduce the emissions of nitrogen oxides resulting from industrial production, the selective catalytic reduction (SCR) technique is commonly employed, where NH3 is used as a reducing agent [10,11,12]. This technique utilizes the catalytic effect of a catalyst to selectively reduce nitrogen oxides into N2 and H2O, thereby meeting emission standards. The reaction equation is as follows:
4NH3 + 4NO + O2 → 4N2 + 6H2O
2NH3 + NO + NO2 → 2N2 + 3H2O
4NH3 + 2NO2 + O2 → 3N2 + 6H2O
Reaction (2) is referred to as the “conventional SCR reaction”, while reaction (3) is known as the “fast SCR reaction”. The fast SCR reaction, compared to the conventional SCR reaction, can occur at lower temperatures and does not require the involvement of oxygen. Due to the rapid oxidation process, the reaction can be completed in a shorter time [13,14,15].
The V2O5/TiO2 catalyst possesses excellent resistance to SO2 and high selectivity towards N2, making it widely applied in industrial processes [5,16]. However, the V2O5/TiO2 catalyst exhibits a narrow effective temperature range (300–400 °C). When the flue gas temperature falls below 300 °C, the catalytic activity of conventional denitrification catalysts decreases, leading to reduced denitrification efficiency and increased ammonia escape rate. The key to SCR technology lies in the development and utilization of catalysts [17,18]: Catalysts with high specific surface area and unique crystal structures signify increased contact area and a greater number of vacancies and defects [19,20,21]. The increased number of acidic sites on the catalyst enhances the adsorption of NH3 [22,23,24]. The reduction in the reduction temperature of metal ions leads to a higher redox capability [25,26,27].
The activity of the catalyst can be increased by the addition of active components. Liu et al. [28] investigated the influence of manganese on the performance of the V2O5/TiO2 catalyst, the results demonstrated that, below 400 °C, the addition of manganese significantly enhanced the activity of the V2O5/TiO2 catalyst in NH3-SCR, the existence of Mn can promote the formation of reactive intermediates. Jihene Arfaoui et al. [29] reported on cerium and sulfate co-modified V2O5/TiO2 catalysts, and characterization revealed that the catalyst possessed excellent features, including high specific surface area, large pore volume, and good thermal stability, the cerium was introduced with the aim of enhancing the NO-SCR activity at low temperature, while, sulfate groups were added in order to improve the N2 selectivity at high temperature. Min et.al. [30] investigated the effects of incorporating Sb, La, Ce, and Mo additives into V2O5-WO3/TiO2 catalysts with high vanadium content for NH3-SCR, the addition of Sb, La, or Ce to the V-site of the V4/W5/G5 catalyst led to improved reaction activity, which was attributed to an increase in the nonstoichiometric V4+(3+) ratio and enhanced surface adsorbed oxygen species. Su et.al. [31] investigated the Sb/Si-doped TiO2 prepared using the sol–gel method, which was used as support of V2O5/TiO2 SCR catalysts, the result indicates that the introduction of Sb or Si in TiO2 support indeed affected the physicochemical properties of the prepared V2O5/TiO2 catalysts, simultaneous enhanced surface acidity and redox capacity led to the better SCR activity for V2O5/SbTiO2.
Some scholars have improved the low-temperature activity of catalysts through catalyst modification. Dong et al. [32] employed antimony-doped V-based catalysts, and when V was loaded onto the synthesized Sb/TiO2, 2V/2Sb/TiO2 prepared after antimony was added to TiO2 showed the best denitrification efficiency at temperatures below 250 °C it exhibited excellent performance at low temperatures, the formation of VSbO4 structure increased redox property, which was important in the SCR reaction, making the valence state of vanadium unstable. Wang et.al. [33] found that the doping of Cu on the V/TiO2 catalyst can enhance the low-temperature SCR activity, the enhancement of redox ability and surface acidity, and the increase in reduced vanadium species and active oxygen species all contribute to the good performance of V-Cu/TiO2. Wang et al. [34] regulated the pH value and doping cobalt based on a V2O5-MoO3/TiO2 catalyst to develop novel low-temperature SCR catalysts, when the pH was 10 and the molar fraction of Co doping was 1 wt%, the catalyst exhibited optimal denitrification activity, the regulation of pH can increase the number of acid sites. Jeongtak Kim et al. [35] reported that the effects of Fe doping on the performance of low-temperature NH3-SCR using V/W/TiO2 catalysts and V/W/TiO2 exhibited an excellent De-NOx efficiency at 180 °C when the loading of Fe is 0.5%, the denitrification rate is decreasing with increasing Fe content, because of the formation of Fe3+, nonstoichiometric V4+, and W5+ species, the redox properties of the V/W-0.5Fe/TiO2 catalyst were significantly improved by the addition of Fe.
There are few studies on catalyst modification under high concentration, wide temperature range, and low gas hourly space velocity conditions, and the denitrification temperature range of catalysts should be further relaxed. In this study, on the basis of the previous laboratory research on industrial catalysts, the industrial catalysts have the most excellent low-temperature denitrification performance when the Mo loading concentration is 3 wt%, and the effect of V concentration on the activity of MoO3-V2O5/TiO2 catalysts in the wide-temperature domain (200–400 °C) was investigated. On the basis of this, by doping WO3 to modify the V2O5-MoO3/TiO2 catalyst, optimizing the catalysts and analyzing mechanisms by characterization, so that it has a very good denitrification performance under the conditions of high concentration of NO2, wide temperature range, and low gas hourly space velocity.

2. Results and Discussion

2.1. Catalyst Activity Testing

Figure 1 illustrates the performance trend of V1Mo3/TiO2, V2Mo3/TiO2, V3Mo3/TiO2, V4Mo3/TiO2, and V5Mo3/TiO2 catalysts over the temperature range of 200–400 °C. The following conclusions can be drawn:
(1)
Within the temperature range of 200–400 °C, with increasing vanadium loading, there is little variation observed in the catalytic performance of the catalysts, the NOx conversion rates of the five catalysts are similar.
(2)
As the temperature rises, the catalytic denitrification performance improves, indicating that higher temperatures lead to enhanced catalyst activity.
(3)
The catalytic effect of all five catalysts was above 80% at 200 °C, showing good low-temperature catalytic activity and the catalytic effect was close to 100% at 400 °C, and the high temperature did not affect the catalytic activity.
Figure 1. NO conversions of NH3-SCR over catalysts with different V2O5 doping.
Figure 1. NO conversions of NH3-SCR over catalysts with different V2O5 doping.
Catalysts 14 00406 g001

2.2. Catalyst Modification

Since varying the loading of V2O5 did not enhance the low-temperature denitrification activity of the catalyst, considering industrial applications, the loading of V2O5 was set to 3 wt%. Additionally, catalysts were prepared by introducing two additional variables based on a 3 wt% MoO3 loading: 5 wt% MoO3 and 10 wt% MoO3. These catalysts were subjected to denitrification reaction activity tests. Furthermore, to investigate the denitrification performance of catalysts co-loaded with MoO3 and WO3 as promoters, as well as the interaction of V, W, and Mo species on the catalyst surface, V3W3Mo3/TiO2 and V3W5Mo5/TiO2 catalysts were synthesized and subjected to denitrification reaction activity testing. Figure 2 illustrates the investigation of different concentrations of MoO3 and WO3 loading on the basis of 3% V. The following conclusions can be drawn:
(1)
The increase in MoO3 loading significantly enhances the denitrification performance of vanadium-based catalysts. Within the temperature range of 200–350 °C, the NOx conversion rates of the V3Mo5/TiO2 catalyst and the V3Mo10/TiO2 catalyst are higher overall compared to the V3Mo3/TiO2 catalyst. The V3Mo5/TiO2 catalyst exhibits the optimal low-temperature denitrification performance, achieving a NOx conversion rate of 93.4% at 200 °C, while the V3Mo10/TiO2 catalyst and the V3Mo3/TiO2 catalyst achieve rates of 91.7% and 86.5%, respectively. The order of catalytic activity among different catalysts is as follows: V3Mo5/TiO2 > V3Mo10/TiO2 > V3Mo3/TiO2. It can be observed that when the MoO3 loading is further increased to 10 wt%, the catalytic activity of the catalyst decreases instead, indicating an optimal loading ratio for the promoter MoO3.
(2)
When V2O5 loading is 3 wt%, MoO3 loading is 3 wt%, and WO3 loading is 3 wt%, the V3W3Mo3/TiO2 catalyst exhibits superior denitrification performance. It achieves slightly higher NOx conversion rates at 200 °C and 225 °C compared to the optimal V-Mo ratio V3Mo5/TiO2 catalyst. Moreover, at 250 °C, the denitrification conversion rate reaches 98.8%, whereas the V3Mo5/TiO2 catalyst without WO3 only achieves 94.5%.
(3)
When both MoO3 and WO3 loading are increased to 5 wt%, the denitrification performance of the V3W5Mo5/TiO2 catalyst decreases within the temperature range of 200–300 °C, indicating an optimal loading amount for MoO3 and WO3 when co-doped. Considering the NOx conversion rates of catalysts with different compositions.
The order of catalytic activity is as follows: V3W3Mo3/TiO2 > V3Mo5/TiO2 > V3W5Mo5/TiO2 > V3Mo10/TiO2. Whether adding only the promoter MoO3 or co-loading MoO3 and WO3, the denitrification performance of the catalysts exhibits a trend of initially increasing and then decreasing with increasing loading amounts. To investigate the effects of different MoO3 loading amounts and the addition of WO3 on catalyst activity, characterization analysis was performed on the four catalysts: V3W3Mo3/TiO2, V3Mo5/TiO2, V3W5Mo5/TiO2, and V3Mo10/TiO2.

2.3. Catalyst Reaction Mechanism Investigation

2.3.1. Catalyst Physical Characterization

BET tests were conducted on the four catalysts, and the calculated specific surface areas obtained using the BET theory are presented in Table 1. It is observed from Table 1 that the specific surface area of V3Mo10/TiO2 is the largest, approximately 66.7 m2/g, while that of V3Mo5/TiO2 is the smallest, approximately 54.4 m2/g. However, based on the results of the catalyst activity tests, it is found that V3Mo10/TiO2 exhibits relatively low activity despite having the largest specific surface area, whereas V3Mo3W3/TiO2 demonstrates the best activity despite having a relatively smaller specific surface area. This suggests that the influence of specific surface area on catalyst activity is minimal, and changes in physical structural parameters have a weaker effect on catalytic performance. The alteration in activity is predominantly attributed to chemical interactions. According to the N2 adsorption-desorption isotherms shown in Figure 3a, both catalysts exhibit type IV isotherms with H2 hysteresis loops, indicating the presence of significant mesoporous characteristics, consistent with the N2 adsorption-desorption isotherms. Figure 3b presents the pore size distribution of V3Mo3W3/TiO2 and V3Mo5/TiO2. It can be observed that both catalysts primarily consist of mesopores. The peak of the pore size distribution curve of the V3Mo3W3/TiO2 catalyst is lower than that of the V3Mo5/TiO2 catalyst, indicating that the most probable pore size of the V3Mo3W3/TiO2 catalyst is smaller than that of the V3Mo5/TiO2 catalyst.

2.3.2. Catalyst Crystal Structure Analysis

Further analysis of the catalysts’ crystal structure was conducted using X-ray diffraction (XRD). As shown in Figure 4, the XRD spectra of the four catalysts are nearly identical, with diffraction peaks appearing at similar positions such as 25.24°,37.8°,48.06°, and 62.72°. The predominant substance identified is anatase TiO2. No other crystalline phases were detected by XRD, indicating that V2O5, MoO3, and WO3 are highly dispersed on the catalyst surface or may exist in an amorphous state [36]. The intensity of XRD peaks among the four catalysts follows the order: V3Mo5/TiO2 > V3W3Mo3/TiO2 > V3Mo10/TiO2 > V3W5Mo5/TiO2. It is observed that with increasing loading amounts of MoO3 and WO3, the intensity of XRD diffraction peaks decreases. This phenomenon may be attributed to the smaller grain size of the TiO2 support, leading to a more pronounced surface effect, a decrease in internal crystallinity, and a deterioration in crystallinity, along with the presence of a certain number of amorphous components. The XRD diffraction peaks of the V3Mo5/TiO2 and V3W3Mo3/TiO2 catalyst samples are stronger, indicating better crystallinity, and larger grain size of the TiO2 support.

2.3.3. Catalyst Surface Acidity Analysis

Figure 5 shows the NH3-TPD profiles of the V3Mo5/TiO2 and V3W3Mo3/TiO2 catalyst samples. The first NH3 desorption peaks of V3Mo5/TiO2 and V3W3Mo3/TiO2 catalysts occur at 120 °C and 130 °C, respectively, with similar NH3 desorption amounts. This suggests that the number of weak acidic sites on both catalysts is comparable. The desorption peaks observed in the low-temperature range are primarily influenced by VOx species, corresponding to the desorption of ammonium ions NH4+ adsorbed on weak Brønsted acidic sites generated by the addition of V2O5. The second NH3 desorption peak of the V3Mo5/TiO2 catalyst appears at 239 °C, while for the V3W3Mo3/TiO2 catalyst, the second NH3 desorption peak is observed at 273 °C, with a larger NH3 desorption amount. This indicates that the WO3-modified catalyst possesses a higher quantity of medium-strong acidic sites, mainly associated with WOx and MoOx species. These sites primarily consist of weakly stable Lewis acidic sites adsorbing ammonia NH3, with WOx species providing a greater number of acidic sites, resulting in a stronger surface acidity of the catalyst. The third NH3 desorption peak is observed at 598 °C for the V3W3Mo3/TiO2 catalyst, indicating that the addition of WO3 introduces new strong acidic sites to the catalyst, corresponding to the desorption of ammonia NH3 adsorbed on strong Lewis acidic sites.

2.3.4. Catalyst Reductivity Analysis

Figure 6 depicts the H2-TPR of the V3Mo5/TiO2 and V3W3Mo3/TiO2 catalyst samples. In the case of the V3Mo5/TiO2 catalyst, the peak observed at 421 °C corresponds to the first reduction step of polymeric octahedral Mo species (weakly bound to the support) from Mo6+ to Mo4+ and the reduction in V species. For the V3W3Mo3/TiO2 catalyst, the first peak appears at 410 °C, corresponding to the first reduction step of Mo6+ to Mo4+, W6+ to W4+, and the reduction in V species. The first reduction peak of the V3W3Mo3/TiO2 catalyst shifts towards lower temperatures and exhibits a higher hydrogen consumption, indicating better low-temperature reduction capability. This suggests strong interactions between V and W, affecting the properties, dispersion, or aggregation of V, W, and Mo species in the co-loaded catalyst, thereby influencing the reducibility of the catalyst. The peaks observed at 538 °C for V3W3Mo3/TiO2 and at 540 °C for V3Mo5/TiO2 are attributed to the reduction in polymeric octahedral Mo species from Mo6+ to Mo4+ with different degrees of aggregation [37]. The peak observed at 746 °C for V3W3Mo3/TiO2 corresponds to the reduction of amorphous tetrahedral Mo species and W species strongly interacting with the support (Mo6+ to Mo4+, W6+ to W4+), as well as the further reduction in Mo and W species (Mo4+ to Mo0, W4+ to W0) [38]. A similar peak is observed at 719 °C for V3Mo5/TiO2, corresponding to the reduction in Mo species. The shift of the third reduction peak towards higher temperatures for V3W3Mo3/TiO2 indicates stronger interactions between WO3 and the support. In summary, the stronger low-temperature redox capability of the V3W3Mo3/TiO2 catalyst is one of the key factors contributing to its optimal low-temperature catalytic activity.

2.3.5. Catalyst Surface Elemental Analysis

Figure 7a presents the XPS spectra of the O1s orbital for the catalyst samples. The characteristic peaks observed around 530.2 eV correspond to lattice oxygen (O2−, denoted as Oβ), while those around 531.5 eV correspond to chemisorbed oxygen (O, denoted as Oα). The positions and intensities of the O1s peaks are quite similar among the four catalyst samples. According to the data in Table 2, the ratios of surface Oα/(Oα + Oβ) for the V3W3Mo3/TiO2, V3W5Mo5/TiO2, V3Mo5/TiO2, and V3Mo10/TiO2 catalysts are 35.2%, 29.1%, 31.2%, and 29.4%, respectively. This indicates that excessive introduction of MoO3 and WO3 additives reduces the amount of surface chemisorbed oxygen, thereby lowering the denitrification activity of the catalyst. The V3W3Mo3/TiO2 catalyst exhibits the highest proportion of surface chemisorbed oxygen, thus demonstrating the best denitrification activity.
The XPS spectra of the V2p orbitals for catalysts with different MoO3 and WO3 loading amounts are presented in Figure 7b. Table 2 provides the peak areas and relative content of V, Mo, and O elements in different catalysts. As shown in Figure 7b, the V2p spectra for all four catalyst samples exhibit similar shapes, displaying distinct double-peak structures. Each peak can be fitted to represent two peaks corresponding to V4+ and V5+ ions. The peaks around 517.2 eV and 525.3 eV represent V5+ 2p3/2 and V5+ 2p1/2, respectively, while those around 516.3 eV and 523.6 eV represent V4+ 2p3/2 and V4+ 2p1/2, respectively. Although the V2O5 loading amounts are the same for all four catalysts, the peak areas differ. According to XPS analysis, the peak areas for the V3W3Mo3/TiO2, V3W5Mo5/TiO2, V3Mo5/TiO2, and V3Mo10/TiO2 catalysts are 6245.5, 5412.5, 5180.2, and 3684.6, respectively, indicating a decrease in vanadium content with increasing additive amounts. Catalysts containing only MoO3 show a greater decrease in surface vanadium content with increasing MoO3 loading, suggesting that excessive additive introduction leads to the aggregation of surface atoms, blocking catalyst pores. However, the addition of WO3 enables better dispersion of active components, thus increasing the surface vanadium atomic concentration. According to Table 2, the relative content ratios of surface V4+/(V4+ + V5+) for the V3W3Mo3/TiO2, V3W5Mo5/TiO2, V3Mo5/TiO2, and V3Mo10/TiO2 catalysts are 85.8%, 85.1%, 87.9%, and 72.9%, respectively. MoO3 promotes the generation of V4+ due to its higher work function. However, the V4+/(V4+ + V5+) ratio for the V3Mo10/TiO2 catalyst is the smallest, as further increases in loading lead to the formation of crystalline MoO3 and V2O5, resulting in spatial hindrance due to changes in spatial configuration, thereby increasing the V5+ content [39]. The absence of spatial hindrance effects in the V3W3Mo3/TiO2 and V3W5Mo5/TiO2 catalysts suggests that simultaneous increases in WO3 loading stabilize V4+. The ratio of V4+/(V4+ + V5+) is a crucial factor affecting the catalytic activity of SCR. The presence of non-stoichiometric vanadium species (Vn+, n ≤ 4) facilitates electron transfer, and under certain conditions, the V4+/(V4+ + V5+) ratio is positively correlated with the catalytic activity of SCR [40]. The combined influence of surface vanadium atomic concentration and V4+ content leads to higher SCR activity in the V3Mo5/TiO2 and V3W3Mo3/TiO2 catalysts compared to the V3Mo10/TiO2 and V3W5Mo5/TiO2 catalysts.
The XPS spectra of the Mo3d orbitals for the catalysts are shown in Figure 7c. From the figure, it can be observed that the Mo3d spectra of the catalysts exhibit two peaks, Mo3d5/2 (around 236.0 eV) and Mo3d3/2 (around 232.8 eV), and the peak areas increase with higher MoO3 loading. According to Table 2, the ratios of surface Mo6+/ (Mo5+ + Mo6+) for the V3W3Mo3/TiO2, V3W5Mo5/TiO2, V3Mo5/TiO2, and V3Mo10/TiO2 catalysts are 71.0%, 76.1%, 63.3%, and 74.7%, respectively. The Mo6+ content also increases with increasing MoO3 loading, but the denitrification activity of the catalysts decreases instead. This indicates that surface Mo species exhibit low or even no activity for the SCR reaction. The reactivity of surface V species is several orders of magnitude higher than that of W and Mo species. W and Mo species promote the reactivity of surface V species through oligomerization reactions rather than electron effects [41].

3. Experiment

3.1. Catalyst Preparation

This study employed the incipient wetness impregnation method to prepare catalysts with varying loadings of active components and promoters. The experimental reagents are shown in Table 3. The preparation procedures are as follows: Initially, the mass of active components and promoters (i.e., V2O5, WO3, MoO3) required for catalyst loading was calculated. Ammonium metavanadate and other necessary chemicals were weighed according to a 1:1 molar ratio of elements. These chemicals were then placed in a beaker and dissolved in an appropriate volume of deionized water. To ensure thorough dissolution of ammonium metavanadate powder, oxalic acid was added to the solution at a molar ratio of 1:2. The solution was stirred at 80 °C until complete dissolution was achieved. Subsequently, the desired amount of TiO2 support was added to the solution and stirred magnetically at 80 °C until a paste-like consistency was attained. Ultrasonication for 20 min was performed to ensure uniform dispersion of the active components. The resulting paste was then dried in an oven at 105 °C for 12 h, followed by calcination in a muffle furnace at 500 °C for 3 h. The resulting material was ground to a particle size of 45–80 mesh and stored for the following study. The prepared catalysts are denoted in Table 4.

3.2. Catalyst Characterization

BET characterization was carried out using the ASAP 2460 multi-station fully automated specific surface and pore size analyzer from Micromeritic Instruments, Inc. (Norcross, GA, USA) to test and analyze the catalysts using the low-temperature N2 physical adsorption method. XRD was carried out using a fully automated polycrystalline X-ray diffractometer, ULTIMA IV, Rigaku, Japan, with a test range of 2θ = 10–80°and a step size of 0.2°, and the data were obtained and then analyzed using Jade software for graphical analysis (Jade 2022.0.03. (SP2)); The NH3-TPD test was carried out by using a Mack AutoChem Ⅱ 2920 chemisorbent analyzer (Norcross, GA, USA). The H2-TPR test was carried out using a Mack AutoChem Ⅱ 2920 chemisorbent analyzer (Norcross, GA, USA). XPS spectroscopy was carried out on a Thermo Scientific Escalab 250Xi X-ray photoelectron spectrometer (Norcross, GA, USA), with the radiation source chosen to be Al Kα, and the charge correction was carried out by using C1s (284.8 eV).

3.3. Activity Test

The catalyst activity testing equipment depicted in Figure 8 consists of a quartz tube. The inside diameter of the quartz tube reactor is 19 mm. The total gas flow rate entering the quartz tube is 120 mL/min. The catalyst quantity is 4 mL, with gas hourly space velocity of 1800 h−1. The inlet concentrations of the simulated flue gas components are [NO2] = [NH3] = 10,000 ppm and [O2] = 13%, with the remaining N2 serving as the balance gas. Initially, under a nitrogen atmosphere, the catalyst is preheated to 200 °C. Subsequently, the mixed gas is introduced for the SCR reaction. The exhaust gas is then connected to a flue gas analyzer. Data is recorded once the readings stabilize. The procedure is repeated at increments of 25 °C until reaching 400 °C. The catalytic efficiency of the catalyst can be calculated using the following formula:
X = 1 N O X o u t N O X i n × 100 %

4. Conclusions

In this study, the conventional V-based catalysts were optimized for the high concentration of NO2, low space velocity, and low-temperature flue gas generated in the production process such as nuclear industry, and the following conclusions were obtained by optimizing the ratio of V-W-Mo to obtain the catalysts with good activity in a wide range of temperature range (200–400 °C), high concentration of NO2 (10,000 ppm), and low space velocity (1800 h−1):
(1) The NOx conversion rates of catalysts with different V2O5 loading concentrations generally increased with increasing temperature and within the temperature range of 200 °C–400 °C, the NOx conversion rates of the five catalysts were nearly identical. Excessive loading of MoO3 can decrease the denitrification performance of the catalyst, with an optimal loading amount observed at 5 wt%; the addition of WO3 significantly enhances the low-temperature activity of the catalyst. When both WO3 and MoO3 are loaded at 3 wt%, the catalyst exhibits the best denitrification performance, achieving a conversion rate of 98.8% at 250 °C.
(2) Increasing the loading of MoO3 to 10 wt% results in an increase in the catalyst’s specific surface area, but a deterioration in surface crystallinity and a decrease in the proportion of V4+ species on the surface. The reaction activity of surface V species, especially the reduced V4+, is significantly higher than that of Mo species. Therefore, the denitrification performance of the V3Mo10/TiO2 catalyst decreases.
(3) The low-temperature activity of the catalyst was significantly enhanced by WO3 modification. XRD and N2 adsorption-desorption experiments indicated that there was almost no change in the physical structure and morphology of the catalyst before and after modification, suggesting that this was not the main reason for the increased activity. XPS and H2-TPR results revealed that the proportion of chemisorbed oxygen (Oα) increased in the WO3-modified catalyst, exhibiting lower V reduction temperatures, which are favorable for low-temperature denitrification activity. NH3-TPD experiments showed that compared to MoOx species, surface WOx species could provide more acidic sites, resulting in stronger surface acidity of the catalyst.

Author Contributions

B.Y.: Data curation, Investigation, Methodology, Writing—original draft, Writing—review & editing. X.L.: Data curation, Investigation, Supervision, Writing—review & editing. S.W.: Conceptualization, Supervision, Writing—review & editing. H.Y.: Data curation, Visualization. S.Z.: Investigation, Resources. L.Y.: Data curation, Visualization. F.L.: Data curation, Formal analysis, Funding acquisition, Visualization, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key R&D Program of China [2023YFE0120600], Natural Science Foundation of Jiangsu Province [BK20220159], and Fundamental Research Funds for the Central Universities [2023KYJD1009].

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kim, V.; Shvab, A.; Brendakov, V. Mathematical Modeling of the Process of Decomposition of Ammonium Polyuranates. In Proceedings of the All-Russian Scientific Conference on Thermophysical Basis of Energy Technologies (TBET), Tomsk, Russia, 9–11 October 2019; National Research Tomsk Polytechnic University: Tomsk, Russia, 2019; Volume 2212. [Google Scholar]
  2. Dussoubs, B.; Jourde, J.; Patisson, F.; Houzelot, J.; Ablitzer, D. Mathematical modelling of uranium dioxide conversion in a moving bed furnace. Powder Technol. 2002, 128, 168–177. [Google Scholar] [CrossRef]
  3. Xie, W. Treatment technology and production practice of high concentrations nitrogen oxide. China Nonferrous Metall. 2010, 39, 47–49. [Google Scholar]
  4. Gao, F.Y.; Tang, X.L.; Yi, H.H.; Li, J.Y.; Zhao, S.Z.; Wang, J.G.; Chu, C.; Li, C.L. Promotional mechanisms of activity and SO2 tolerance of Co- or Ni-doped MnOx-CeO2 catalysts for SCR of NOx with NH3 at low temperature. Chem. Eng. J. 2017, 317, 20–31. [Google Scholar] [CrossRef]
  5. Huang, L.; Zeng, Y.Q.; Chang, Z.F.; Zong, Y.H.; Wang, H.; Zhang, S.L.; Yu, Y. Promotional effect of phosphorus modification on improving the Na resistance of V2O5-MoO3/TiO2 catalyst for selective catalytic reduction of NOx by NH3. Mol. Catal. 2021, 506, 111565. [Google Scholar] [CrossRef]
  6. Lee, G.M.Y.; Ye, B.R.; Kim, W.G.; Jung, J.I.; Park, K.Y.; Jeong, B.; Kim, H.D.; Kim, T. V2O5-WO3 catalysts treated with titanium isopropoxide using a one-step co-precipitation method for selective catalytic reduction with NH3. Catal. Today 2023, 411–412, 113924. [Google Scholar] [CrossRef]
  7. Jo, S.H.; Kim, K.; Seo, S.H.; Kim, T.H.; Yu, S.; Kim, T.H.; Son, Y.S. A study on additives to improve electron beam technology for NOx and SO2 reduction. Radiat. Phys. Chem. 2021, 183, 109397. [Google Scholar] [CrossRef]
  8. Jang, H.S.; Xing, S.L. A model to predict ammonia emission using a modified genetic artificial neural network: Analyzing cement mixed with fly ash from a coal-fired power plant. Constr. Build. Mater. 2020, 230, 117025. [Google Scholar] [CrossRef]
  9. Ye, B.; Jeong, B.; Lee, M.J.; Kim, T.H.; Park, S.S.; Jung, J.; Lee, S.; Kim, H.D. Recent trends in vanadium-based SCR catalysts for NOx reduction in industrial applications: Stationary sources. Nano Converg. 2022, 9, 51. [Google Scholar] [CrossRef]
  10. Kwon, D.W.; Park, K.H.; Hong, S.C. Enhancement of SCR activity and SO2 resistance on VOx/TiO2 catalyst by addition of molybdenum. Chem. Eng. J. 2016, 284, 315–324. [Google Scholar] [CrossRef]
  11. Zhu, X.; Zhang, L.Q.; Dong, Y.; Ma, C.Y. NO2-NH3 SCR over Activated Carbon: A Combination of NH4NO3 Formation and Consumption. Energy Fuels 2021, 35, 6167–6178. [Google Scholar] [CrossRef]
  12. Zhu, X.; Zhang, L.Q.; Chen, G.F.; Wang, T.; Dong, Y.; Ma, C.Y. Investigation of low-temperature NO2 reduction by NH3 over activated carbon: Effect of pore structure and flue gas conditions. J. Energy Inst. 2022, 101, 243–253. [Google Scholar] [CrossRef]
  13. Colombo, M.; Nova, I.; Tronconi, E.; Schmeisser, V.; Bandl-Konrad, B.; Zimmermann, L. NO/NO2/N2O-NH3 SCR reactions over a commercial Fe-zeolite catalyst for diesel exhaust aftertreatment: Intrinsic kinetics and monolith converter modelling. Appl. Catal. B-Environ. 2012, 111, 106–118. [Google Scholar] [CrossRef]
  14. Selleri, T.; Gramigni, F.; Nova, I.; Tronconi, E. NO oxidation on Fe- and Cu-zeolites mixed with BaO/Al2O3: Free oxidation regime and relevance for the NH3-SCR chemistry at low temperature. Appl. Catal. B-Environ. 2018, 225, 324–331. [Google Scholar] [CrossRef]
  15. Gao, X.; Liu, S.J.; Zhang, Y.; Du, X.S.; Luo, Z.Y.; Cen, K.F. Low temperature selective catalytic reduction of NO and NO2 with NH3 over activated carbon-supported vanadium oxide catalyst. Catal. Today 2011, 175, 164–170. [Google Scholar] [CrossRef]
  16. Chen, C.M.; Cao, Y.; Liu, S.T.; Chen, J.M.; Jia, W.B. Review on the latest developments in modified vanadium-titanium-based SCR catalysts. Chin. J. Catal. 2018, 39, 1347–1365. [Google Scholar] [CrossRef]
  17. Zhang, T.T.; Shi, T.; Wang, Y.; Hao, Y.H.; Gao, Y.H.; Li, H.R.; Jia, L.; Liu, F.R.; Zeng, S.H. Orchestrating dual adsorption sites and unravelling Ce-Mn interaction and reaction mechanisms for efficient NH3-SCR. J. Catal. 2024, 429, 115260. [Google Scholar] [CrossRef]
  18. Geng, Y.; Shan, W.P.; Yang, S.J.; Liu, F.D. W-Modified Mn-Ti Mixed Oxide Catalyst for the Selective Catalytic Reduction of NO with NH3. Ind. Eng. Chem. Res. 2018, 57, 9112–9119. [Google Scholar] [CrossRef]
  19. Zong, L.Y.; Dong, F.; Zhang, G.D.; Han, W.L.; Tang, Z.C.; Zhang, J.Y. Highly Efficient Mesoporous V2O5/WO3-TiO2 Catalyst for Selective Catalytic Reduction of NOx: Effect of the Valence of V on the Catalytic Performance. Catal. Surv. Asia 2017, 21, 103–113. [Google Scholar] [CrossRef]
  20. Guo, J.; Zhang, G.D.; Tang, Z.C.; Zhang, J.Y. Controlled Synthesis of Mesoporous CeO2-WO3/TiO2 Microspheres Catalysts for the Selective Catalytic Reduction of NOx with NH3. Catal. Surv. Asia 2019, 23, 311–321. [Google Scholar] [CrossRef]
  21. Chen, Y.; Guan, B.; Liu, Z.Q.; Wu, X.Z.; Guo, J.F.; Zheng, C.Z.; Zhou, J.F.; Su, T.X.; Han, P.C.; Yang, C.Z.; et al. Review on advances in structure-activity relationship, reaction & deactivation mechanism and rational improving design of selective catalytic reduction deNOx catalysts: Challenges and opportunities. Fuel 2023, 343, 127924. [Google Scholar]
  22. Qi, G.S.; Yang, R.T.; Chang, R. MnOx-CeO2 mixed oxides prepared by co-precipitation for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B-Environ. 2004, 51, 93–106. [Google Scholar] [CrossRef]
  23. Gao, C.R.; Li, J.W.; Zhang, J.; Sun, X.L. DFT Study on the Combined Catalytic Removal of N2O, NO, and NO2 over Binuclear Cu-ZSM-5. Catalysts 2022, 12, 438. [Google Scholar] [CrossRef]
  24. Van de Vyver, S.; Román-Leshkov, Y. Emerging catalytic processes for the production of adipic acid. Catal. Sci. Technol. 2013, 3, 1465–1479. [Google Scholar] [CrossRef]
  25. Fu, Y.; Zhong, X.; Chang, H.; Li, J. Mechanism study on CO-SCR over Ce-Co-Ox mixed oxides catalysts. China Environ. Sci. 2018, 38, 2934–2940. [Google Scholar]
  26. Seo, C.K. Improvement of NOx and CO Reduction Performance at Low Temperature of H2-SCR Catalyst. Int. J. Automot. Technol. 2022, 23, 1509–1515. [Google Scholar] [CrossRef]
  27. Liu, Z.M.; Jia, B.; Zhang, Y.Y.; Haneda, M. Engineering the Metal-Support Interaction on Pt/TiO2 Catalyst to Boost the H2-SCR of NOx. Ind. Eng. Chem. Res. 2020, 59, 13916–13922. [Google Scholar] [CrossRef]
  28. Liu, Z.M.; Li, Y.; Zhu, T.L.; Su, H.; Zhu, J.Z. Selective Catalytic Reduction of NOx by NH3 over Mn-Promoted V2O5/TiO2 Catalyst. Ind. Eng. Chem. Res. 2014, 53, 12964–12970. [Google Scholar] [CrossRef]
  29. Arfaoui, J.; Ghorbel, A.; Petitto, C.; Delahay, G. Novel V2O5-CeO2-TiO2-SO42- nanostructured aerogel catalyst for the low temperature selective catalytic reduction of NO by NH3 in excess O2. Appl. Catal. B-Environ. 2018, 224, 264–275. [Google Scholar] [CrossRef]
  30. Jung, M.G.; Shin, J.H.; Kwon, D.W.; Hong, S.C. Promotional effects of Me (Sb, La, Ce, Mo) additives on the NH3-SCR activity and SO2 durability of V2O5-WO3/TiO2 catalysts. Process Saf. Environ. Prot. 2024, 183, 911–924. [Google Scholar] [CrossRef]
  31. Su, C.X.; Zhu, L.; Xu, M.T.; Zhong, Z.P.; Wang, X.Y.; Gao, Y.; Zhu, Y.Z. Influences analysis of Sb/Si dopant in TiO2 on NH3-SCR activity and low temperature SO2 resistance of V2O5/TiO2 catalysts. Appl. Surf. Sci. 2023, 637, 157996. [Google Scholar] [CrossRef]
  32. Kim, D.H.; Kwon, D.W.; Hong, S.C. Structural characteristics of V-based catalyst with Sb on selective catalytic NOx reduction with NH3. Appl. Surf. Sci. 2021, 538, 148088. [Google Scholar] [CrossRef]
  33. Wang, P.; Guo, R.T. The promotion effect of copper doping on the potassium resistance of V/TiO2 catalyst for selective catalytic reduction of NO with NH3. Chem. Pap. 2017, 71, 2253–2259. [Google Scholar] [CrossRef]
  34. Wang, R.A.; Zhang, Y.L.; Fan, X.; Li, J. Cobalt-/pH-Modified V2O5-MoO3/TiO2 Catalyst with Enhanced Activity for the Low-Temperature Selective Catalytic Reduction Process. Catalysts 2023, 13, 844. [Google Scholar] [CrossRef]
  35. Kim, J.; Won, J.M.; Jeong, S.K.; Yu, K.S.; Shin, K.; Hwang, S.M. Fe-promoted V/W/TiO2 catalysts for enhanced low-temperature denitrification efficiency. Appl. Surf. Sci. 2022, 601, 154290. [Google Scholar] [CrossRef]
  36. Kang, T.H.; Kim, H.S.; Lee, H.; Kim, D.H. Synergistic effect of V2O5-WO3/TiO2 and H-ZSM-5 catalysts prepared by physical mixing on the selective catalytic reduction of NOx with NH3. Appl. Surf. Sci. 2023, 614, 156159. [Google Scholar] [CrossRef]
  37. Yuan, P.; Cui, C.S.; Han, W.; Bao, X.J. The preparation of Mo/γ-Al2O3 catalysts with controllable size and morphology via adjusting the metal-support interaction and their hydrodesulfurization performance. Appl. Catal. A-Gen. 2016, 524, 115–125. [Google Scholar] [CrossRef]
  38. Zhang, H.; Han, J.; Niu, X.W.; Han, X.; Wei, G.D.; Han, W. Study of synthesis and catalytic property of WO3/TiO2 catalysts for NO reduction at high temperatures. J. Mol. Catal. A-Chem. 2011, 350, 35–39. [Google Scholar] [CrossRef]
  39. Gao, Y.; Wu, X.D.; Ran, R.; Si, Z.C.; Ma, Z.R.; Wang, B.D.; Weng, D. Effects of MoOx on dispersion of vanadia and low-temperature NH3-SCR activity of titania supported catalysts: Liquid acidity and steric hindrance. Appl. Surf. Sci. 2022, 585, 152710. [Google Scholar] [CrossRef]
  40. Lázaro, M.J.; Boyano, A.; Herrera, C.; Larrubia, M.A.; Alemany, L.J.; Moliner, R. Vanadium loaded carbon-based monoliths for the on-board No reduction: Influence of vanadia and tungsten loadings. Chem. Eng. J. 2009, 155, 68–75. [Google Scholar] [CrossRef]
  41. Li, J.; Peng, Y.; Chang, H.; Li, X.; Crittenden, J.C.; Hao, J. Chemical poison and regeneration of SCR catalysts for NOx removal from stationary sources. Front. Environ. Sci. Eng. 2016, 10, 413–427. [Google Scholar] [CrossRef]
Figure 2. NO conversions of NH3-SCR over catalysts with different MoO3 and WO3 doping.
Figure 2. NO conversions of NH3-SCR over catalysts with different MoO3 and WO3 doping.
Catalysts 14 00406 g002
Figure 3. Pore parameter curves of V3Mo5/TiO2 and V3W3Mo3/TiO2 catalysts: (a) N2 adsorption-desorption isotherms; (b) pore size distribution.
Figure 3. Pore parameter curves of V3Mo5/TiO2 and V3W3Mo3/TiO2 catalysts: (a) N2 adsorption-desorption isotherms; (b) pore size distribution.
Catalysts 14 00406 g003
Figure 4. XRD patterns of catalysts with different MoO3, WO3 doping.
Figure 4. XRD patterns of catalysts with different MoO3, WO3 doping.
Catalysts 14 00406 g004
Figure 5. NH3-TPD curves of V3Mo5/TiO2 and V3W3Mo3/TiO2 catalysts.
Figure 5. NH3-TPD curves of V3Mo5/TiO2 and V3W3Mo3/TiO2 catalysts.
Catalysts 14 00406 g005
Figure 6. H2-TPR curves of V3Mo5/TiO2 and V3W3Mo3/TiO2 catalysts.
Figure 6. H2-TPR curves of V3Mo5/TiO2 and V3W3Mo3/TiO2 catalysts.
Catalysts 14 00406 g006
Figure 7. The XPS spectra of catalysts with varying MoO3 and WO3 loading amounts: (a) O1s; (b) V2p; (c) Mo3d.
Figure 7. The XPS spectra of catalysts with varying MoO3 and WO3 loading amounts: (a) O1s; (b) V2p; (c) Mo3d.
Catalysts 14 00406 g007
Figure 8. System diagram of a denitrification unit.
Figure 8. System diagram of a denitrification unit.
Catalysts 14 00406 g008
Table 1. BET test results of the four catalysts.
Table 1. BET test results of the four catalysts.
SampleBET (m2/g)
V3Mo3W3/TiO256.7
V3W5Mo5/TiO256.2
V3Mo10/TiO266.7
V3Mo5/TiO254.4
Table 2. Peak areas of V, Mo, and elements on the surface of catalysts and the proportions of different valence states.
Table 2. Peak areas of V, Mo, and elements on the surface of catalysts and the proportions of different valence states.
SamplePeak AreaAtomic Ratio %
VMoV4+/(V5+ + V4+)Mo6+/(Mo6+ + Mo5+)Oα/(Oα + Oβ)
V3W3Mo3/TiO26242.512,949.085.871.035.2
V3W5Mo5/TiO25412.515,172.285.176.129.1
V3Mo5/TiO25180.220,918.687.963.331.2
V3Mo10/TiO23684.624,661.472.974.729.4
Table 3. Experimental reagents.
Table 3. Experimental reagents.
NumberChemical FormulaSpecificationsManufacturer
1TiO2Nano gradeMacklin (Shanghai, China)
2(NH4)10H2(W2O7)6Analytical gradeMacklin (Shanghai, China)
3NH4VO3Analytical gradeMacklin (Shanghai, China)
4(NH4)6Mo7O24Analytical gradeMacklin (Shanghai, China)
5C2H2O4·2H2OAnalytical gradeSinopharm Chemical Reagent Co., Ltd. (Shanghai, China)
Table 4. Chemical composition and nomenclature of catalysts.
Table 4. Chemical composition and nomenclature of catalysts.
Chemical CompositionNamed
V: 1%, Mo: 3%, TiO2:96%V1Mo3/TiO2
V: 2%, Mo: 3%, TiO2:95%V2Mo3/TiO2
V: 3%, Mo: 3%, TiO2:94%V3Mo3/TiO2
V: 4%, Mo: 3%, TiO2:93%V4Mo3/TiO2
V: 5%, Mo: 3%, TiO2:92%V5Mo3/TiO2
V: 3%, Mo: 5%, TiO2:92%V3Mo5/TiO2
V: 3%, Mo: 10%, TiO2:87%V3Mo10/TiO2
V: 3%, W: 3%, Mo: 3%, TiO2:91%V3W3Mo3/TiO2
V: 3%, W: 5%, Mo: 5%, TiO2:87%V3W5Mo5/TiO2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yu, B.; Liu, X.; Wu, S.; Yang, H.; Zhou, S.; Yang, L.; Liu, F. Study on Novel SCR Catalysts for Denitration of High Concentrated Nitrogen Oxides and Their Reaction Mechanisms. Catalysts 2024, 14, 406. https://doi.org/10.3390/catal14070406

AMA Style

Yu B, Liu X, Wu S, Yang H, Zhou S, Yang L, Liu F. Study on Novel SCR Catalysts for Denitration of High Concentrated Nitrogen Oxides and Their Reaction Mechanisms. Catalysts. 2024; 14(7):406. https://doi.org/10.3390/catal14070406

Chicago/Turabian Style

Yu, Bo, Xingyu Liu, Shufeng Wu, Heng Yang, Shuran Zhou, Li Yang, and Fang Liu. 2024. "Study on Novel SCR Catalysts for Denitration of High Concentrated Nitrogen Oxides and Their Reaction Mechanisms" Catalysts 14, no. 7: 406. https://doi.org/10.3390/catal14070406

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