Next Article in Journal
Supported Ni Single-Atom Catalysts: Synthesis, Structure, and Applications in Thermocatalytic Reactions
Next Article in Special Issue
Engineering Surface Properties of CuO/Ce0.6Zr0.4O2 Catalysts for Efficient Low-Temperature Toluene Oxidation
Previous Article in Journal
Production of Fructooligosaccharides Using a Commercial Heterologously Expressed Aspergillus sp. Fructosyltransferase
Previous Article in Special Issue
Low-Temperature NH3-SCR Performance and In Situ DRIFTS Study on Zeolite X-Supported Different Crystal Phases of MnO2 Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cobalt-/pH-Modified V2O5-MoO3/TiO2 Catalyst with Enhanced Activity for the Low-Temperature Selective Catalytic Reduction Process

Key Laboratory of Beijing on Regional Air Pollution Control, Beijing University of Technology, Beijing 100124, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(5), 844; https://doi.org/10.3390/catal13050844
Submission received: 8 March 2023 / Revised: 20 April 2023 / Accepted: 4 May 2023 / Published: 6 May 2023

Abstract

:
Currently, the elimination of gaseous pollutants—particularly nitrogen oxides—has emerged as a significant concern. Among various deNOx technologies, selective catalytic reduction (SCR) has gained prominence as the primary approach for NOx abatement, owing to its superior performance. In this study, novel low-temperature SCR catalysts were developed by regulating the pH value and doping cobalt based on a V2O5-MoO3/TiO2 (VMT) catalyst. The results show an increased SCR performance with 82.8% and 91.1% for catalysts after pH (=10) modification (VMT-10) and (1 wt%) Co/pH (=10) modification (1CoVMT-10), respectively. H2-TPR, NH3-TPD, XPS and DRIFTS confirmed that the pH regulation transformed polymerization V species into isolated V5+=O, thus leading to an increase in the number of acid sites, which enhanced the NH3 and NO2 adsorption capacity. Furthermore, the DRIFTS study indicated that the NH3-SCR reaction over 1CoVMT-10 followed the E–R and L–H mechanism.

1. Introduction

As a main contaminant in steel and coal-fired industries, photochemical smog, acid rain, and the damage to human health caused by NOx urgently requires solutions. Selective catalytic reduction (SCR) is the main method for NOx removal due to its high performance among various deNOx technologies. V2O5-MoO3(WO3)/TiO2 catalysts have been widely commercialized in industrial NOx removal, operated at temperatures ranging from 280 °C to 400 °C. V2O5-MoO3(WO3)/TiO2 (VM(W)T) is one of the most widely used catalysts for the NH3-SCR process at present. The market for high-temperature catalysts has matured due to the installation of denitration devices in the front of dust and sulfur removal equipment [1]. However, the existing high amount of dust delivered a negative effect on catalysts; high-concentration dust contains a large number of harmful elements, such as Na, K, Ca, and As, which will poison the catalyst. Therefore, novel SCR units are commonly installed downstream of the airflow (approximately 180–200 °C) to the dust collector, necessitating the use of low-temperature SCR catalysts capable of performing denitrification. Moreover, anatase TiO2, which plays an important carrier role in VM(W)T catalysts, is susceptible to sintering at high temperatures, and the resultant crystal phase conversion can lead to catalytic activity reduction [2,3]. Hence, developing highly efficient catalysts for low-temperature NOx conversion seems imperative.
The properties of V2O5-MoO3(WO3)/TiO2 catalysts have been extensively investigated. Denitrification activity can be effectively enhanced by including WO3 and MoO3 (especially at amounts of 10% and 6%, respectively) in the composite [3,4,5]. Modifying SCR catalysts with other additives, such as transition metals, is another way to develop better catalytic behavior [6,7,8,9].
Cobalt oxide (CoOx) has proven to be one of the most efficient metal oxide catalysts for increasing the removal of N2O, CH4, and NO due to its excellent reduction ability and abundant oxygen vacancy [9,10]. For example, the effect of Co/ZSM proved that the ammonia species adsorbed on the Brønsted acid site and that –NH2 may take part in the medium/low-temperature region in the NH3-SCR process [11]. A catalyst consisting of K and Co3O4 could promote the availability of Co2+ by effectively regenerating the Co3+ oxidized by N2O [12]. For Co–Ce binary metal oxide catalysts, the redox pair of Ce3+/Ce4+ in CeO2 facilitates the storage and release of lattice oxygen species, re-oxidizing Co2+ to Co3+ [13]. In addition, the participation of Co in the catalyst can simultaneously increase the activity of both Oα and Oβ and L–H (Langmuir–Hinshelwood mechanism) was also proposed according to the source of active oxygen species [14].
Apart from modifying catalysts with various elements, pH regulation could be regarded as another effective way to improve SCR activity. A one-step ion exchange method was used to increase the amount of Cu2+ by changing the pH of the precursor mixture with HNO3, in which the H+ concentration has a significant influence on the Cu loading and distribution [15]. A series of Mn-doped perovskite La–Mn oxides were prepared during the synthesis with ammonia, while Mn4+ content was dominant when regulating the pH to 1.4, which displayed excellent NH3-SCR activity [16]. In another study, ammonium hydroxide was successfully used to reduce aggregated polymerization in V2O5/TiO2 catalysts, and the as-obtained isolated vanadium species mainly worked on the adsorption and oxidation of NH3, resulting in higher NOx conversion [17].
In previous tests, a VMT catalyst prepared in our laboratory could maintain a denitrification efficiency of over 85% under factory conditions, but this is not enough for low-temperature SCR catalysts with better efficiency and stability. In this work, modified V2O5-MoO3/TiO2 (VMT) catalysts generated by pH (=10) regulation (VMT-10) and recombined (1 wt%) cobalt doping (1CoVMT-10) were investigated. The purpose of the present work is to clarify the impact of Co/pH affected by the physicochemical properties and catalytic performance of VMT catalysts; the structures of the catalysts were characterized by XRD, BET, SEM, H2-TPR, NH3-TPD, and in situ DRIFTS.

2. Results and Discussion

2.1. NH3-SCR Activity

The parameters that varied during the experiments were the amount of Co doping and the pH regulation value. The samples were herein denoted as xCoVMT-y, where x is the mass fraction of Co (x = 1, 3, 5), while y is the pH (y = 9, 9.5, 10, 10.5, 11).
The addition of Co was initiated during impregnation. Based on Figure 1a–c, the denitration efficiency of the catalysts at different pH values changed as a function of Co addition (1, 3, and 5 wt%). Figure 1a shows that the NOx conversion rate was 71.4% at 160 °C for unmodified VMT. For the 1CoVMT sample series, at the investigated pH 9–11 interval, the best catalyst (1CoVMT-10) developed to 91.1% denitrification efficiency was observed at pH 10, which was 19.7% higher than that for the VMT reference. By increasing the amount of Co (Figure 1b,c), the most efficient activity decreased to 89.3% (3CoVMT-10) and 85.8% (5CoVMT-10).
The NOx removal efficiency decreased with the increased Co addition after pH adjustment (Figure 1d), suggesting that excessive Co has an inhibitory effect on the catalyst. Probably, an increase in active components unavoidably resulted in increasing particle sizes and the formation of polymeric constituents or large grains on the surface. This further reduced the capacity of the catalyst to activate molecular oxygen and remove lattice oxygen [18].
For further investigation, the NOx conversion curves of VMT, VMT-10, and 1CoVMT-10 are compared in Figure 1e. The catalytic efficiency could be improved both by increasing the pH value and the amount of Co. Two parameters synergistically promoted the NOx conversion efficiency of the catalysts. Table S1 lists the NO removal efficiency under laboratory conditions of Co-doped or pH-regulated catalysts. Compared to previous studies, it can be seen that current research has a wider temperature range under low-temperature testing conditions and has an advantage in denitrification efficiency at higher NOx concentration.

2.2. Textural Properties

The catalysts prepared by adjusting the pH and adding different amounts of Co were compared with VMT. The results were evaluated using the BET and BJH methods, as shown in Table 1, and the calculated value of C = 71 falls within the applicable range for oxides (Figure S1).
Adjusting the pH value resulted in larger specific surface areas, pore volumes, and pore diameters. However, simultaneous Co/pH modification reduced the specific surface area, which might be ascribed to the blockage of partial pores caused by cobalt impregnation.
Figure 2a–c show the adsorption isotherms and pore size distributions of VMT, VMT-10, and 1CoVMT-10, wherein all samples could be characterized by a type IV isotherm [19]. Mesopores were also present according to the hysteresis loops. The pore structure provided numerous active sites for gas molecules. For 1CoVMT-10, the enlarged hysteresis loop indicates that gas adsorption and diffusion are facilitated in this sample, thus increasing the catalytic reaction (Figure 2c). In Figure 2d, the pore distributions of the three catalysts are compared. As a result of pH modification, the number of micropores increased (both VMT-10 and 1CoVMT-10), whereas the addition of Co had little effect on the pore size distribution. It can be speculated that the influence of the Co content on the pore structure damage is limited. Variations in pore structure caused by adjusting the pH were likely to be one of the most important reasons for the improved denitrification efficiency.
Based on the SEM images of VMT (Figure 3a), the rough surface with large particles developed a relatively flat appearance with small pores after pH regulation (VMT-10; Figure 3b). For the visibly deeper cracks and pores detected on the surface of VMT, the particles in VMT-10 are more compact, and the cracks become narrower. This might be due to the transformation of polymeric V–O–V bonds into isolated V=O bands [20]. Then, after Co/pH modification, the surface particles became even finer, indicating that Co doping also had an isolated dispersion influence on the catalyst (Figure 3c). The labeled area consists of small particles and is distributed more uniformly on the surface. As the 1CoVMT-10 catalyst had the best efficiency, it is suggested that partially isolated species might play an advantageous role in terms of catalyst activation [6].

2.3. XRD Measurement

The diffraction peaks at 25.3°, 37.8°, 48°, 54°, 55°, 62.8°, 69°, 70.3°, and 75° were attributed to anatase TiO2 (Figure 4). No rutile crystal phase and active components (CoOx, VOx, MoOx) were observed during XRD analysis, suggesting that they were well-dispersed on the support structure, or the amount of rutile crystal phase and active components were too low to be detected.

2.4. H2-TPR Experiments

H2-TPR measurements were carried out for the samples. As shown in Figure 5, 3 Tmax types could be observed: isolated monomeric species (≤497–507 °C), polymeric species (537 °C), and bulk amorphous V2O5 (579 °C) [20]. The difference observed for VMT and VMT-10 at 413 °C and 429 °C, respectively, in Figure 5a, indicated that isolated monomeric species could be caused by ammonium hydroxide. For polymeric and bulk amorphous species, due to overlapping characteristic peaks, we conducted peak separation analysis for the two peaks at 537 °C and 579 °C, as seen in Figure S2, and filled the peak area of vanadyl species from the fitting data present in Table S2. The results showed that the monomeric species of the VMT catalyst accounted for approximately 60.7%, and after the addition of NH4OH, the proportion of monomeric species in the catalyst increased to 78.6%. Furthermore, the monomeric species of the 1CoVMT-10 catalyst reached the maximum proportion to 79.4%, and its bulk amorphous species decreased to the lowest value among the three catalysts, indicating that the modification indeed increased the quantity of the monomeric species on the catalyst surface. According to the research of S. Besselmann, polymeric and crystal vanadiums on V/Ti catalysts can be selectively removed after washing with aqueous ammonia, resulting in monomeric species [17]. It is generally believed that monomeric V species activate catalysts better than polymerized V2O5 [21]. The reaction of H2 with V5+ eventually generates V3+ [22]. The peak at 700 °C is characteristic of the transformation of V2O5 to V6O13 [21], while others believe that it could also be due to the decomposition of bulk V2O5 [20].
Vanadium can form many different compounds depending on the chemical properties of the solution. For example, VO43– and VO3(OH)2– ions mainly exist in strong alkaline media [6]. For VMT-10, the decreased height in the 480–560 °C range indicates the lack of polymerization (Figure 5a). Furthermore, the generated isolated V=O on Brønsted acid on V/Ti catalysts has proven to play a significant role in SCR [20,21].
At 288 °C and 366 °C (Figure 5a), cobalt provided new reduction sites, while the former belongs to the transformation of Co3+ to Co2+, and the latter belongs to the transformation of Co2+ to metallic Co. [12]. For the Co reduction peaks, the peak at 366 °C for Co2+ is more pronounced for 1CoVMT-10, while the peak of Co3+ is relatively weak. With an increase in Co%, the characteristic peaks of both Co3+ and Co2+ shifted to higher temperatures, gradually approaching the characteristic peak of vanadyl monomeric species until partially overlapping, as shown in Figure 5b. When Co% was added excessively, vanadyl species reduction peaks gradually shifted from 420 °C to 428 °C and 453 °C for 1Co, 3Co, and 5Co, respectively. The peak shift here is probably due to the transformation of vanadyl species from monomeric species to polymeric species. The shift toward higher temperatures caused by excess Co addition is detrimental to the conversion of NO to NO2 [23], as low-temperature reducibility was one of the key factors affecting catalytic activity [24].
During H2-TPR measurements, the effects of Co doping were also investigated in Figure 5c. After 1 wt% cobalt doping, a peak with a broad shoulder was observed at 507 °C, which indicates that the presence of isolated V species and a few polymeric V species consumes large amounts of hydrogen. It was found that appropriate polymerization can improve reaction activity [6], and this heterogeneous structure is formed to produce with a unique effect to catalytic performance improvement.

2.5. XPS Measurement

The XPS corresponding spectra of Co 2p3/2, O 1s, and V 2p3/2 are illustrated in Figure 6, and the surface binding energies and ratio of the valence state are summarized in Table 2. For 1CoVMT-10, Co2+ (782.2 eV) shows a dominant amount compared to Co3+ (779.9 eV) [25], as seen in Figure 6a. The ratio of Co2+/Co3+ is large, which is 5.00 to 1CoVMT-10, indicating that Co2+ has high activity in the SCR reaction. As the percentage of Co in the catalyst increases, a decrease in the proportion of Co2+ and a concomitant increase in the proportion of Co3+ was observed.
From Figure 6b and Table 2, the V 2p3/2 peaks are concentrated at 514–518 eV, which are attributed to V5+ and V4+ characteristic peaks. V4+ ions were abundant in VMT, but the growing ratio of V5+/V4+ brought about pH regulation and Co doping in VMT-10 and 1CoVMT-10 indicates that V5+ has higher activity in the SCR reaction. However, the amount of V shows a certain decrease in VMT-10. When Co was added, the total amount of V was readded. The presence of Co2+ and V5+ enables the Co2+ + V5+ → Co3+ + V4+ reaction. NH4+ adsorbed on V=O reacts with Co3+ to form NH3+ by dehydrogenation, and the obtained NH3+ combined NO to produce H3N+NO, which finally decomposed to H2O and N2. Then, the increasing synergized Co2+ could be oxidized by V5+, while V4+ was oxidized by O2, respectively.
Modifying the pH affected the distribution of lattice oxygen and adsorbed oxygen significantly. In Figure 6c, the O 1s spectrum is commonly divided into two components: oxygen species at 529–531 eV (lattice oxygen species; Oβ), and 531–534 eV (adsorbed oxygen species; Oα). From Table 2, affected by the ammonia solution, the ratio of Oα/(Oβ + Oα) in VMT-10 greatly increased from 14.11% to 58.34%.
Considering the high mobility of Oα, the larger ratio of Oα/(Oα + Oβ) is beneficial for yielding improved catalytic activity [26]. VMT-10 exhibited the highest proportion of adsorbed oxygen (Oα/(Oβ + Oα)) at 58.34%. Massive surface adsorbed oxygen was favorable to the recirculation of reactive oxygen species, which was also an advantage for the improvement of catalyst activity [24].
Co doping restored the amount of Oβ in 1CoVMT-10 (Figure 6c). Since oxygen vacancies were generated to balance the decreasing charge of Co3+, the ratio of Co2+/Co3+ for 5.00 was considered to be indicative of the creation of oxygen vacancies on the surface. It is possible that the increase in crystal oxygen is caused by the addition of Co oxides, which leads to the formation of surface oxygen vacancies, as Co3+ is converted to Co2+ [27]. The generation of V5+, together with the observation of Co2+ species, was demonstrative of the synergetic effect between vanadium and cobalt via charge transfer (V5+ + Co2+ ↔ V4+ + Co3+), which further confirmed the earlier deduction from H2-TPR.
In addition, studies have shown that both Oβ and Oα could promote the participation of active substances in the process of oxidizing NO to NO2, theoretically making the rapid SCR cycling reaction (2NH3 + NO + NO2 → 2N2 + 3H2O) easier to occur [27,28], related to higher SCR activity. Oβ and Oα play different roles throughout the entire SCR process. For the VMT-10 catalyst, the advantages introduced by Oα dominated and chemisorbed oxygen had a significant effect on the deep oxidation reactions of the reducing substances [29]. For the 1CoVMT-10 catalyst with the added Co element, Oβ is more advantageous. Maintaining the continuity of surface oxygen vacancies in the catalyst is one of the main factors affecting the catalyst activity [30].

2.6. NH3-TPD Measurement

For the NH3-SCR method, the basic reaction gas of NH3 first adsorbs on the acid sites and then participates in SCR [31].
Acid sites can be classified according to their thermal stability: weakly (physically) adsorbed NH3 (100–250 °C), NH3 adsorbed on Lewis acid sites (250–350 °C), and NH3 adsorbed on Brønsted acid sites (over 350 °C) for chemisorption [32,33]. As illustrated in Figure 7a, all of the samples exhibited similar peaks at around 93 °C, which were attributed to weakly (physically) adsorbed NH3. The adsorption peaks at around 350 °C and 435 °C were attributed to NH3 adsorbed at the Lewis acid site, and the peak at 655 °C corresponded to the NH3 bound at Brønsted acid sites (NH4+) [34].
In Figure 7b, a gradual increase in the content of basic sites brought about through pH regulation and Co doping also indicated the strengthened ability for NH3 adsorbed on the catalyst surface. The increased number of strong acid sites on VMT-10 was attributed to the formation of strong isolated V=O bonds, suggesting that VMT-10 and 1CoVMT-10 possessed stronger acid sites. Compared to VMT-10, the 1CoVMT-10 catalyst had more sites at 350 °C and 435 °C, indicating that Co species improved the ability to bind NH3 at Lewis acid sites in the medium region. Additionally, the result goes some way toward reflecting that abundant acid sites had better efficiency in terms of catalytic behavior.

2.7. In Situ Drifts Measurement

2.7.1. Adsorption of NH3

Previous studies proved that NH4+ species adsorbed on the catalyst surface react with gaseous NO species during SCR [35]. As shown in Figure 8, after N2 purging pretreatment at 300 °C, NH3 was inlet and adsorbed on catalysts at ambient temperature. When the temperature gradually increased from 100 to 300 °C at a rate of 4 °C/min, infrared bands were recorded at 100 °C, 150 °C, 200 °C, 250 °C, and 300 °C, respectively.
The bands at 1435, 1445, 1672, and 1660 cm–1 were ascribed to the bending and symmetric bending vibrations of NH4+ at Brønsted acid sites [36], or the bending vibrations of N–H bonds at Brønsted acid sites originating from the chemisorption of NH3 on the V–OH groups of the catalyst [21]. In addition, the gradually decreased peak intensities of 1445, 1454, and 1660 cm–1 in VMT-10 and 1CoVMT-10 indicated that the modification due to pH regulation and cobalt doping both reduced the formation of polymeric HO–V–OH bonds.
The infrared absorption bands at 1242 cm–1 (Figure 8a), 1227 cm–1 (Figure 8b), and 1239 cm–1 (Figure 8c) were similar to those at 1623 cm–1 (Figure 8a) and 1620 cm–1 (Figure 8c). All of them can be attributed to the asymmetric bending vibrations of the N–H bonds during the N–H chemisorption at the Brønsted acidic sites [21,37], which appeared at 150 °C and could be detected at high temperatures. The band at 3352 cm–1 can be ascribed to asymmetric N–H stretching vibrations at the Lewis acidic sites (Figure 8a) [37,38,39]. The absorption bands at 3556, 3670, and 3737 cm–1 were attributed to the O–H tensile vibrations caused by the interaction between surface hydroxyl groups and NH3 [39,40]. From Figure 8a–c, the bands observed at 2820, 2833, 3016, 3020, 3027, 3204, and 3215 cm–1 can be attributed to NH4+ species [41]; meanwhile, these species could be easily desorbed at 150 °C on the catalyst surface, which was in accordance with the results of NH3-TPD.
The bands observed at 1415 cm–1 and 1445 cm–1 on VMT can be attributed to the adsorption capacity of V2O5 on the catalyst surface were also related to the asymmetric bending vibrations of NH4+ chemisorbed on the Brønsted acid sites [26]. It is evident that among the three catalysts, 1CoVMT-10 exhibited a faster disappearance of NH4+ groups at these sites, indicating that the NH4+ groups on the surface of 1CoVMT-10 are more active. The absorption bands at 1020 and 1096 cm–1 (Figure 8a), 1044 and 1102 cm–1 (Figure 8b), and 1025 and 1096 cm–1 (Figure 8c) were attributed to δS(NH) symmetric bending vibrations at the Brønsted acidic sites. After pH regulation, the intensity of the bands on VMT-10 decreased, indicating that the transformation of –NH2 into –NH was inhibited under alkaline conditions. In conclusion, Co/pH modification improved the reductive properties of the catalysts by restricting the growth of –NH groups.
The bands at 1553, 2066 (Figure 8b), 1507, and 2066 cm–1 (Figure 8c) can be ascribed to N–H bonds chemisorbed on V=O Lewis acidic sites, and the lattice oxygen on V=O can react with adsorbed NH3 [21]—these benefited from Co/pH modification. It was only VMT on which any obvious infrared absorption band could not be observed in these regions due to an insufficient number of V=O sites.

2.7.2. Reaction of NH3 with NO+O2

The reaction of catalysts with NH3 and NO during SCR was also simulated by in situ DRIFTS. Following surface NH3 saturation and purging with N2, a 0.07 vol% NO and 5 vol% O2 mixture was introduced into the reaction chamber. The spectra were recorded at different temperatures, as in Figure 9.
The bands at 1670, 2835, 3041 cm–1 (Figure 9a), 1674, 2817, 3022 cm–1 (Figure 9b), and 1661, 2827, 3033, 3212 cm–1 (Figure 9c) were attributed to the presence of NH4+. When reaching 120 °C, the bands from VMT-10 and 1CoVMT-10 disappeared rapidly after NO+O2 exposure. This could be due to their well-developed desorption capacity to NH3 [41]. NH4+ groups on the surface of VMT did not sufficiently react with NO+O2 in air, as their characteristic band could still be observed at 200 °C.
Adding NO+O2, NH4+, and NH3 groups on the surface can generate active intermediates that react with NO2 during SCR [37]. Characteristic bands of NO2 were observed at 1629 (Figure 9a), 1637 (Figure 9b), and 1632 cm–1 (Figure 9c). It was deduced that pH regulation increased the number of Lewis acidic sites on the catalyst surface and improved its adsorption capacity for NO2 [37,42].
We have zoomed in on the 1250–1500 cm−1 range of the infrared spectrum in Figure 9a′–c′. The asymmetric bending vibrations of NH4+ chemisorbed on Brønsted acid can be observed at 1445 cm−1 and 1415 cm−1 in Figure S2 [6,26,43]. The 1341 cm−1 band was attributed to NOx species, while the 1372 cm−1 and 1367 cm−1 bands corresponded to nitrate species [37,44].
The peak at 3547 cm−1 was attributed to the vibration of O–H in H2O. As the temperature increased, physically adsorbed water gave rise to strong infrared bands at approximately 3550 cm−1 [39,45]. The VMT and VMT-10 exhibit a more prominent peak at this location, whereas the 1CoVMT-10 shows a weak and broad peak, suggesting that the adsorption of surface H2O was suppressed at this location.
For VMT, the absorption bands at 1013 and 1629 cm–1 can be attributed to the NOx species (Figure 9a); the one at 1013 cm–1 corresponds to the N2O2 group υas, while the one at 1629 cm–1 can be ascribed either to nitrite (NO2) or nitrate (NO3) [44,46,47]. The intensity of all these bands increased slightly with increasing temperature.
VMT-10 possessed the highest NO2 band (1637 cm–1) intensity (Figure 9b), indicating the formation of a strong bond between NO2 and the Lewis acidic site during adsorption. However, too strong of adsorption inhibited the conversion of NOx. As shown in Figure 9c, the band at 1632 cm–1 for 1CoVMT-10 reduced significantly compared to that for VMT-10. This proves that Co addition could compensate for the disadvantage caused by the strong adsorption [10].

2.7.3. Reaction of NH3 and NO+O2 on 1CoVMT-10

During the SCR process, the mixed atmosphere of NH3, NO, and O2 exposed upon Lewis and Brønsted acidic sites could be provided by V2O5. As shown in Figure 10, a higher adsorption strength was provided by NH3 than for NO, indicating that the dominant reaction in SCR occurred via the E–R (Eley–Rideal) reaction mechanism [48,49].
Studies have shown that a limited amount of NO can be adsorbed on V/Ti catalysts [50,51]. Whereas, in Figure 10, the adsorption bands of NO (1902, 1845 cm–1), NO2 (1635 cm–1), and –NO2 (1350 cm–1) could still be observed. Based on the contents mentioned above, the reaction of oxidized intermediates of NO2 and adsorbed NH3 on the surface of VMT-10 and 1CoVMT-10 promoted SCR, proving that this process might also follow the L–H (Langmuir–Hinshelwood) mechanism [39].
In this way, NH3 was adsorbed on the SCR catalysts and was activated through H abstraction to form NH2, which then reacted with the gas phase NO to form a nitroamide (NH2NO) species. The nitrous acid was produced through a reoxidation reaction. The nitrous acid reacted with ammonia and then produced ammonium nitrite. Both ammonium nitrite and nitrosamine were unstable and would decompose into N2 and H2O.
The Eley–Rideal mechanism (reactions between the coordinated NH3 and gaseous NO over the Lewis acid sites):
NH 3 g + O a NH 2 a + OH a
NO g + NH 2 a NH 2 NO a
NH 2 NO a N 2 g + H 2 O
The SCR of NO with NH3 on the V2O5/TiO2 catalyst proceeds through the oxidation of NO to NO2. NH3 is adsorbed in the form of coordinated NH3, NH4+, and –NH2. V2O5 promotes the formation of –NH2, which is the main intermediate of the SCR reaction.
Catalysts in the SCR reaction follow the Langmuir–Hinshelwood mechanism (reactions between the absorbed NO2 and coordinated NH3):
2 NO + O 2 g 2 NO 2 a
NH 3 g NH 3 a
NO 2 a + 2 NH 3 a NO 2 NH 3 2
NO a + NO 2 NH 3 2 2 N 2 + H 2 O
In conclusion, this catalytic reaction could be predominantly performed using the E–R mechanism and the L–H mechanism.

3. Materials and Methods

3.1. Catalyst Preparation

The pH-modified catalysts were prepared as follows. Ammonium hydroxide (25–28%) (AR, 99.5%, Tianjin Fuchen Chemicals Co., Ltd., Tianjin, China) was used to adjust the pH between 2.5 and 11 under continuous stirring for 30 min. The original pH of the synthesis mixture was 2.5, which was also considered for comparison. Finally, the container was put in an oven at 110 °C for 4–5 h and calcined in air at 500 °C for 3 h. The original V2O5-MoO3/TiO2 catalyst is prepared using the wet impregnation method. NH4VO3 (AR, 99.0%, Tianjin Fuchen Chemicals Co., Ltd., China) was added to an oxalic acid (H2C2O4, AR, 99.5%, Tianjin Fuchen Chemicals Co., Ltd., China) solution under stirring until the color of the solution changed gradually from orange to dark blue. The mole ratio of NH4VO3 and H2C2O4 was 1:2. (NH4)2MoO4 (AR, 99.5%, Tianjin Fuchen Chemical Co., Ltd., China) was then added, followed by TiO2 (98%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), and the obtained catalyst slurry was then continued to stir for 2 h. Finally, the catalyst slurry was placed into an oven, dried at 110 °C for 4–5 h, and calcined under air at 500 °C for 3 h. The mass fraction of V2O5 and MoO3 was 3 and 6 wt%, respectively. After cooling, the catalysts were sieved with a 20–40 mesh sieve and prepared for the test.
The Co-modified catalysts were also prepared using the impregnation method. The primary distinction between the cobalt-modified catalyst and the VMT catalyst lies in the former’s introduction of a requisite mass percentage of (CH3COO)2Co (AR, 99.5%, Tianjin Fuchen Chemicals Co., Ltd., China) subsequent to the addition of (NH4)2MoO4, but prior to the inclusion of TiO2.
For the pH-modified catalysts prepared in this paper, after preparing the catalyst precursor solution the pH of the resulting solution was adjusted with ammonium hydroxide (NH4OH, 25–28%, AR, Tianjin Fuchen Chemicals Co., Ltd., Tianjin, China). The level selected for the pH was between 9 and 11. The original pH of the synthesis mixture was 2.5. The drying and calcination of the modified catalyst slurries were conducted in the same manner as the treatment of the original catalyst. In addition, to ensure experimental reproducibility, the modified catalysts were dried and calcined under the same conditions as the original catalyst (dried at 110 °C for 4–5 h and calcined in air at 500 °C for 3 h).

3.2. Catalytic Activity

NO conversion was carried out in a fixed-bed glass reactor (φ = 20 mm), and the required reaction temperature was controlled using an electric furnace. The amount of catalyst in each test was 2 g, and the total amount of flue gas was 1.50 L/min, with a gas hourly space velocity of 30,000 h–1. The reaction gas composition was 0.07 vol% NOx, 0.07 vol% NH3, and 5 vol% O2, while N2 was used as the carrier gas. The following reaction in the flue gas occured:
4NO + 4NH3 → O2 + 4N2 + 6H2O
A nitrogen oxide analyzer (High Level 42i, Thermal Electron Co., Waltham, MA, USA) was used to test the concentration of NOx and calculate its conversion rate. A pH tester (WTW Multi3620, Burladingen, Germany) was used to control the pH of the catalyst slurry. The conversion of NOx was calculated as follows:
NO x   conversion   %   = NO x in NO x out NO x in × 100 %   Assume   Q 1 = Q 2
where [NOx]in and [NOx]out represent the concentrations of NOx of the inlet and outlet gas streams, respectively, while Q1 and Q2 represent the flow rates of the inlet and outlet, respectively.

3.3. Structure Characterizations

The Brunauer–Emmett–Teller (BET) and Barret–Joyner–Halenda (BJH) methods were used to measure the specific surface area by N2 adsorption with a Micromeritics Gemini V instrument (Norcross, GA, USA). After preparation, the samples were degassed under a vacuum at 110 °C for 1 h.
X-ray diffraction (XRD) patterns were obtained with a Bruker D8 Advance instrument (Billerica, MA, USA) employing a Ni-filtered Cu Κα (λ = 0.15406) radiation. The spectra were registered between 10 and 90° with a step of 0.02°.
X-ray photoelectron spectroscopy (XPS) measurements were carried out on a Thermo–Fisher Scientific K-Alpha+ instrument (Waltham, MA, USA) under an ultravacuum at a pressure lower than 5 × 10–6 Pa. A monochromatic Al Kα X-ray source (1486.6 eV) at a power of 15 kV (15 mA) was used for the analysis.
Temperature-programmed reduction with hydrogen (H2-TPR) was carried out using a Micrometritics Auto-Chem II 2920 instrument (Norcross, GA, USA). For the measurements, a 10% H2 and 90% Ar mixture was used, the heating rate was 10 °C/min, and the samples were pretreated at 300 °C. The consumption of H2 was detected with a thermal conductivity detector.
Temperature-programmed desorption with ammonia (NH3-TPD) was carried out with a Chem-BET Pulsar TPR/TPD instrument (Quantachrome, Boynton Beach, FL, USA). After pretreatment at 300 °C for 1 h, the samples were saturated with pure NH3 vapor at ambient temperature, then purged with He as the temperature was increased at a rate of 10 °C/min.
In situ diffuse reflectance infrared Fourier-transform spectroscopy (In situ DRIFTS) data were collected by accumulating 64 scans at a resolution of 4 cm–1. The spectra were acquired using an in situ DRIFTS cell equipped with a gas flow system (Bruker Tensor II, Billerica, MA, USA). The samples were pretreated at 300 °C under N2 for 1 h and then cooled to room temperature.

4. Conclusions

Modified catalysts with pH regulation and Co doping can be synthesized using the wet impregnation method. The 1CoVMT-10 catalyst exhibited optimal denitrification activity when the pH was 10 and the molar fraction of Co doping was 1 wt%.
The pH regulation method could improve the dispersion of the particles and the uniformity on TiO2. Both pH regulation and cobalt doping methods can convert polymeric vanadium to an isolated form, and the increased V5+=O species resulted in more Brønsted and Lewis acid sites for NH3 and NOx adsorbed species on the surface of the catalyst, promoting SCR activity. The ratios of Co2+/Co3+ and V5+/V4+ developed, and abundant V5+ and Co2+ were beneficial for facilitating the SCR reaction cycle in the Co2+ + V5+ → Co3+ + V4+ process. In addition, both Oβ and Oα could promote the participation of active substances in the process of oxidizing NO to NO2, theoretically making the rapid SCR cycling reaction easier to occur in terms of VMT-10 and 1CoVMT-10, respectively. Combined with the NH3-TPD and DRIFTS studies, the results suggest that the NH3 and NO2 adsorbed species on the surface of 1CoVMT-10 have higher activity in the SCR reaction than VMT. The developed amount of medium and strong Lewis acidic sites showed an intensive effect in terms of NH3 adsorption. Furthermore, the DRIFTS study indicated that the NH3-SCR reaction over 1CoVMT-10 followed the E–R and L–H mechanism.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13050844/s1, Figure S1: BET fitting curve of VMT catalyst; Figure S2: H2-TPR profiles of the catalyst samples: (a) VMT; (b) VMT-10; (c) 1CoVMT-10; Table S1: NO removal efficiency under laboratory conditions of Co-doped or pH-regulated catalysts; Table S2: Peak area of vanadyl species species from fitting data.

Author Contributions

Conceptualization, R.W. and Y.Z.; methodology, R.W.; software, R.W. and Y.Z.; validation, R.W.; formal analysis, R.W.; investigation, R.W.; resources, R.W. and Y.Z.; data curation, R.W.; writing—original draft preparation, R.W.; writing—review and editing, J.L.; visualization, X.F.; supervision, X.F. and J.L.; project administration, J.L.; funding acquisition, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (No.2017YFC0210303).

Data Availability Statement

The data presented in this study are available in article and Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, Y.X.; Liang, H.L.; Chen, X.; Chen, C.; Wang, X.Z.; Dai, C.Y.; Hu, L.M.; Chen, Y.F. Effect of preparation methods on denitration performance of V-Mo/TiO2 catalyst. J. Fuel Chem. Technol. 2020, 48, 189–196. [Google Scholar] [CrossRef]
  2. Zheng, Y.J.; Jensen, A.D.; Johnsson, J.E. Deactivation of V2O5-WO3-TiO2 SCR catalyst at a biomass-fired combined heat and power plant. Appl. Catal. B-Environ. 2005, 60, 253–264. [Google Scholar] [CrossRef]
  3. Nova, I.; Dall Acqua, L.; Lietti, L.; Giamello, E.; Forzatti, P. Study of thermal deactivation of a de-NOx commercial catalyst. Appl. Catal. B-Environ. 2001, 35, 31–42. [Google Scholar] [CrossRef]
  4. Zhang, S.; Li, H.Y.; Zhong, Q. Promotional effect of F-doped V2O5-WO3/TiO2 catalyst for NH3-SCR of NO at low-temperature. Appl. Catal. A.-Gen. 2012, 435–436, 156–162. [Google Scholar] [CrossRef]
  5. Zhao, W.; Zhong, Q.; Pan, Y.X.; Zhang, R. Systematic effects of S-doping on the activity of V2O5/TiO2 catalyst for low-temperature NH3-SCR. Chem. Eng. J. 2013, 228, 815–823. [Google Scholar] [CrossRef]
  6. Gao, Y.; Wu, X.D.; Ran, R.; Si, Z.C.; Ma, Z.R.; Wang, B.D.O.; 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]
  7. Jae, G.H.; Mahboob, U.; Chun, M.; Yong, S.C.; Seong, G.S.; Min, C.S.; Young, S.C.; Kim, D. Low-temperature shift DeNOx activity of Nanoflake V2O5 loaded WO3/TiO2 as NH3-SCR catalyst. Inorg. Chem. Commun. 2022, 137, 109191. [Google Scholar] [CrossRef]
  8. Liu, F.; He, H.; Lian, Z.; Shan, W.; Xie, L.; Asakura, K.; Yang, W.; Deng, H. Highly dispersed iron vanadate catalyst supported on TiO2 for the selective catalytic reduction of NOx with NH3. J. Catal. 2013, 307, 340–351. [Google Scholar] [CrossRef]
  9. Lónyi, F.; Solt, H.E.; Valyon, J.; Boix, A.; Gutierrez, L.B. The SCR of NO with methane over In,H- and Co,In,H-ZSM-5 catalysts: The promotional effect of cobalt. Appl. Catal. B-Environ. 2012, 117–118, 212–223. [Google Scholar] [CrossRef]
  10. Kubacka, A.; Janas, J.; Sulikowski, B. In/Co-ferrierite: A highly active catalyst for the CH4-SCR NO process under presence of steam. Appl. Catal. B-Environ. 2006, 69, 43–48. [Google Scholar] [CrossRef]
  11. Xue, H.Y.; Guo, X.M.; Meng, T.; Mao, D.S.; Ma, Z. NH3-SCR of NO over M/ZSM-5 (M = Mn, Co, Cu) catalysts: An in-situ DRIFTS study. Surf. Interfaces 2022, 29, 101722. [Google Scholar] [CrossRef]
  12. Kimihiro, A.; Chie, O.; Shinji, I.; Yasushi, S.; Masashi, I. Potassium-doped Co3O4 catalyst for direct decomposition of N2O. Appl. Catal. B-Environ. 2008, 78, 242–249. [Google Scholar] [CrossRef]
  13. Bo, Z.; Zhu, J.; Yang, S.; Yang, H.; Yan, J.; Cen, K. Enhanced plasma-catalytic decomposition of toluene over Co-Ce binary metal oxide catalysts with high energy efficiency. RSC Adv. 2019, 9, 7447–7456. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, H.; Han, W.; Tang, Z. Tailored design of high-stability CoMn1.5Ox@TiO2 double-wall nanocages derived from Prussian blue analogue for catalytic combustion of o-dichlorobenzene. Appl. Catal. B-Environ. 2020, 276, 119133. [Google Scholar] [CrossRef]
  15. Zhao, H.W.; Yang, G.P.; Hill, A.J.; Luo, B.E.; Jing, G.H. One-step ion-exchange from Na-SSZ-13 to Cu-SSZ-13 for NH3-SCR by adjusting the pH value of Cu-exchange solution: The effect of H+ ions on activity and hydrothermal stability. Microporous Mesoporous Mat. 2021, 324, 111271. [Google Scholar] [CrossRef]
  16. Fan, A.; Jing, Y.; Guo, J.; Shi, X.; Yuan, S.; Li, J. Investigation of Mn doped perovskite La-Mn oxides for NH3-SCR activity and SO2/H2O resistance. Fuel 2022, 310, 122237. [Google Scholar] [CrossRef]
  17. Besselmann, S.; Löffler, E.; Muhler, M. On the role of monomeric vanadyl species in toluene adsorption and oxidation on V2O5/TiO2 catalysts: A Raman and in situ DRIFTS study. Mol. Catal. 2000, 162, 401–411. [Google Scholar] [CrossRef]
  18. Qiu, Y.; Liu, B.; Du, J.; Tang, Q.; Liu, Z.H.; Liu, R.L.; Tao, C.Y. The monolithic cordierite supported V2O5-MoO3/TiO2 catalyst for NH3-SCR. Chem. Eng. J. 2016, 294, 264–272. [Google Scholar] [CrossRef]
  19. Pulido Melián, E.; González Díaz, O.; Ortega Méndez, A.; López, C.R.; Nereida Suárez, M.; Doña Rodríguez, J.M.; Navío, J.A.; Fernández Hevia, D.; Pérez Peña, J. Efficient and affordable hydrogen production by water photo-splitting using TiO2-based photocatalysts. Int. J. Hydrogen Energy 2013, 38, 2144–2155. [Google Scholar] [CrossRef]
  20. Bulushev, D.A.; Kiwi-Minsker, L.; Rainone, F.; Renken, A. Characterization of surface vanadia forms on V/Ti-oxide catalyst via temperature-programmed reduction in hydrogen and spectroscopic methods. J. Catal. 2002, 205, 115–122. [Google Scholar] [CrossRef]
  21. Tang, F.S.; Zhuang, K.; Yang, F.; Yang, L.L.; Xu, B.L.; Qiu, J.H.; Fan, Y.N. Effect of dispersion state and surface properties of supported vanadia on the activity of V2O5/TiO2 catalysts for the selective catalytic reduction of NO by NH3. Chin. J. Catal. 2012, 33, 933–940. [Google Scholar] [CrossRef]
  22. Topsoe, N.Y.; Topsoe, H.; Dumesic, J.A. Vanadia/titania catalysts for selective catalytic reduction (SCR) of nitric-oxide by ammonia: I. combined temperature-programmed in-situ FTIR and on-line mass-spectroscopy studies. J. Catal. 1995, 151, 226–240. [Google Scholar] [CrossRef]
  23. Yu, C.; Huang, B.; Dong, L.; Chen, F.; Yang, Y.; Fan, Y.; Yang, Y.; Liu, X.; Wang, X. Effect of Pr/Ce addition on the catalytic performance and SO2 resistance of highly dispersed MnOx/SAPO-34 catalyst for NH3-SCR at low temperature. Chem. Eng. J. 2017, 316, 1059–1068. [Google Scholar] [CrossRef]
  24. Li, X.; Niu, Y.; Li, J.; Yang, M.; Chen, R.; Shao, D.; Zheng, X.; Zhang, C.; Qi, Y. Trace Co doping improves NH3-SCR performance and poisoning resistance of Ce-Mn-based catalysts. Chem. Eng. J. 2023, 454, 140180. [Google Scholar] [CrossRef]
  25. Chen, Q.; Guo, R.; Wang, Q.; Pan, W.; Yang, N.; Lu, C.; Wang, S. The promotion effect of Co doping on the K resistance of Mn/TiO2 catalyst for NH3-SCR of NO. J. Taiwan Inst. Chem. Eng. 2016, 64, 116–123. [Google Scholar] [CrossRef]
  26. Wu, R.; Li, L.; Zhang, N.; He, J.; Song, L.; Zhang, G.; Zhang, Z.; He, H. Enhancement of low-temperature NH3-SCR catalytic activity and H2O & SO2 resistance over commercial V2O5-MoO3/TiO2 catalyst by high shear-induced doping of expanded graphite. Catal. Today 2021, 376, 302–310. [Google Scholar] [CrossRef]
  27. Chen, Y.; Chen, Z.; Zhang, C.; Chen, L.; Tang, J.; Liao, Y.; Ma, X. Multiple pollutants control of NO, benzene and toluene from coal-fired plant by Mo/Ni impregnated TiO2-based NH3-SCR catalyst: A DFT supported experimental study. Appl. Surf. Sci. 2022, 599, 153986. [Google Scholar] [CrossRef]
  28. Gao, F.; Tang, X.; Yi, H.; Li, J.; Zhao, S.; Wang, J.; Chu, C.; Li, C. Promotional mechanisms of activity and SO2 tolerance of Co-or Ni-doped MnOx-CeO2 catalysts for SCR of NOx with NH3 at low temperature. Chem. Eng. J. 2017, 317, 20–31. [Google Scholar] [CrossRef]
  29. Chen, M.; Zheng, X. Effect of promoter thallium for a novel selectivity oxidation catalyst studied by X-ray photoelectron spectroscopy. J. Mol. Catal. 2003, 201, 161–166. [Google Scholar] [CrossRef]
  30. Xi, X.Y.; Zeng, F.; Zhang, H.; Wu, X.F.; Ren, J.; Bisswanger, T.; Stampfer, C.; Hofmann, J.P.; Palkovits, R.; Heeres, H.J. CO2 hydrogenation to higher alcohols over K-promoted bimetallic Fe-In catalysts on a Ce-ZrO2 support. ACS Sustain. Chem. Eng. 2021, 9, 6235–6249. [Google Scholar] [CrossRef]
  31. Qiu, L.; Pang, D.D.; Zhang, C.L.; Meng, J.J.; Zhu, R.S.; Ouyang, F. In situ IR studies of Co and Ce doped Mn/TiO2 catalyst for low-temperature selective catalytic reduction of NO with NH3. Appl. Surf. Sci. 2015, 357, 189–196. [Google Scholar] [CrossRef]
  32. Kang, S.H.; Ryu, J.H.; Kim, J.H.; Sai Prasad, P.S.; Bae, J.W.; Cheon, J.Y.; Jun, K.W. ZSM-5 supported cobalt catalyst for the direct production of gasoline range hydrocarbons by Fischer-Tropsch synthesis. Catal. Lett. 2011, 141, 1464–1471. [Google Scholar] [CrossRef]
  33. Lonyi, F.; Valyon, J. On the interpretation of the NH3-TPD patterns of H-ZSM-5 and H-mordenite. Microporous Mesoporous Mat. 2001, 47, 293–301. [Google Scholar] [CrossRef]
  34. Dong, G.; Zhang, Y.F.; Zhao, Y.; Bai, Y. 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]
  35. Centeno, M.A.; Carrizosa, I.; Odriozola, J.A. In situ DRIFTS study of the SCR reaction of NO with ammonia over a high loading (15% weight) vanadia-titania catalyst. Phys. Chem. Chem. Phys. 1999, 1, 349–354. [Google Scholar] [CrossRef]
  36. Chen, Z.D.; Li, N.; Zhang, K.; Hou, L.M.; Wu, W.F. In-situ infrared spectroscopic study of the mechanism of the low temperature selective catalytic reduction of NO surface by Mn/bastnaesite concentrate. Int. J. Hydrogen Energy 2022, 47, 24777–24795. [Google Scholar] [CrossRef]
  37. Wang, J.P.; Yan, Z.; Liu, L.L.; Chen, Y.; Zhang, Z.T.; Wang, X.D. In situ DRIFTS investigation on the SCR of NO with NH3 over V2O5 catalyst supported by activated semi-coke. Appl. Surf. Sci. 2014, 313, 660–669. [Google Scholar] [CrossRef]
  38. Dall’Acqua, L.; Nova, I.; Lietti, L.; Ramis, G.; Busca, G.; Giamello, E. Spectroscopic characterisation of MoO3/TiO2 denox-scr catalysts: Redox and coordination properties. Phys. Chem. Chem. Phys. 2000, 2, 4991–4998. [Google Scholar] [CrossRef]
  39. Sun, D.; Liu, Q.; Liu, Z.; Gui, G.; Huang, Z. An In Situ DRIFTS Study on SCR of NO with NH3 Over V2O5/AC Surface. Catal. Lett. 2009, 132, 122–126. [Google Scholar] [CrossRef]
  40. Primet, M.; Pichat, P.; Mathieu, M.V. Infrared study of the surface of titanium dioxides. I. Hydroxyl groups. J. Phys. Chem. 1971, 75, 1216–1220. [Google Scholar] [CrossRef]
  41. Liu, K.; Liu, F.; Xie, L.; Shan, W.; He, H. DRIFTS study of a Ce-W mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Catal. Sci. Technol. 2015, 5, 2290–2299. [Google Scholar] [CrossRef]
  42. Karami, A.; Salehi, V. The influence of chromium substitution on an iron-titanium catalyst used in the selective catalytic reduction of NO. J. Catal. 2012, 292, 32–43. [Google Scholar] [CrossRef]
  43. Chen, L.; Li, J.; Ge, M. DRIFT Study on Cerium-Tungsten/Titiania Catalyst for Selective Catalytic Reduction of NOx with NH3. Environ. Sci. Technol. 2010, 44, 9590–9596. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, H.Y.; Voskoboinikov, T.; Sachtler, W.M.H. Reduction of NOx over Fe/ZSM-5 catalysts: Mechanistic causes of activity differences between alkanes. Catal. Today 1999, 54, 483–494. [Google Scholar] [CrossRef]
  45. Frost, R.L. An infrared and Raman spectroscopic study of the uranyl micas. Spectroc. Acta Part A-Molec. Biomolec. Spectr. 2004, 60, 1469–1480. [Google Scholar] [CrossRef] [PubMed]
  46. Li, Y.J.; Slager, T.L.; Armor, J.N. Selective reduction of NOx by methane on Co-Ferrierites: II. catalyst characterization. J. Catal. 1994, 150, 388–399. [Google Scholar] [CrossRef]
  47. Aylor, A.W.; Lobree, L.J.; Reimer, J.A.; Bell, A.T. An infrared study of NO reduction by CH4 over Co-ZSM-5. Stud. Surf. Sci. Catal. 1996, 101, 661–670. [Google Scholar] [CrossRef]
  48. Amblard, M.; Burch, R.; Southward, B.W.L. A study of the mechanism of selective conversion of ammonia to nitrogen on Ni/γ-Al2O3 under strongly oxidising conditions. Catal. Today 2000, 59, 365–371. [Google Scholar] [CrossRef]
  49. Zhu, Z.P.; Liu, Z.Y.; Niu, H.X.; Liu, S.J. Promoting effect of SO2 on activated carbon-supported vanadia catalyst for NO reduction by NH3 at low temperatures. J. Catal. 1999, 187, 245–248. [Google Scholar] [CrossRef]
  50. Abello, L.; Husson, E.; Repelin, Y.; Lucazeau, G. Vibrational spectra and valence force field of crystalline V2O5. Spectroc. Acta Part A-Molec. Biomolec. Spectr. 1983, 39, 641–651. [Google Scholar] [CrossRef]
  51. Topsøe, N.; Topsøe, H. Combined in-situ FTIR and on-line activity studies: Applications to vanadia-titania DeNOx catalyst. Catal. Today 1991, 9, 77–82. [Google Scholar] [CrossRef]
Figure 1. NH3-SCR activity: (a) 1CoVMT series; (b) 3CoVMT series; (c) 5CoVMT series; (d) 1Co/3Co/5CoVMT-10; (e) VMT/VMT-10/1CoVMT-10.
Figure 1. NH3-SCR activity: (a) 1CoVMT series; (b) 3CoVMT series; (c) 5CoVMT series; (d) 1Co/3Co/5CoVMT-10; (e) VMT/VMT-10/1CoVMT-10.
Catalysts 13 00844 g001
Figure 2. N2 adsorption–desorption isotherms and pore size distribution: (a) VMT; (b) VMT-10; (c) 1CoVMT-10; (d) Sample comparison of pore size distribution.
Figure 2. N2 adsorption–desorption isotherms and pore size distribution: (a) VMT; (b) VMT-10; (c) 1CoVMT-10; (d) Sample comparison of pore size distribution.
Catalysts 13 00844 g002
Figure 3. SEM images of catalyst surface: (a) VMT; (b) VMT-10; (c) 1CoVMT-10.
Figure 3. SEM images of catalyst surface: (a) VMT; (b) VMT-10; (c) 1CoVMT-10.
Catalysts 13 00844 g003
Figure 4. XRD patterns of the catalyst samples: (a) VMT; (b) 1CoVMT; (c) 3CoVMT; (d) 5CoVMT; (e) VMT-10; (f) 1CoVMT-10; (g) 3CoVMT-10; (h) 5CoVMT-10.
Figure 4. XRD patterns of the catalyst samples: (a) VMT; (b) 1CoVMT; (c) 3CoVMT; (d) 5CoVMT; (e) VMT-10; (f) 1CoVMT-10; (g) 3CoVMT-10; (h) 5CoVMT-10.
Catalysts 13 00844 g004
Figure 5. H2-TPR profiles of the catalyst samples: (a) VMT/VMT-10/1CoVMT-10; (b) 1Co/3Co/5CoVMT-10; (c) VMT/1CoVMT.
Figure 5. H2-TPR profiles of the catalyst samples: (a) VMT/VMT-10/1CoVMT-10; (b) 1Co/3Co/5CoVMT-10; (c) VMT/1CoVMT.
Catalysts 13 00844 g005
Figure 6. XPS profiles of catalyst samples: (a) Co; (b) V; (c) O.
Figure 6. XPS profiles of catalyst samples: (a) Co; (b) V; (c) O.
Catalysts 13 00844 g006
Figure 7. (a) TPD profiles of catalyst samples; (b) Distribution of various acidic sites.
Figure 7. (a) TPD profiles of catalyst samples; (b) Distribution of various acidic sites.
Catalysts 13 00844 g007
Figure 8. In situ DRIFTS patterns of catalysts exposed to 0.07 vol% NH3 and desorption under N2: (a) VMT; (b) VMT-10; (c) 1CoVMT-10.
Figure 8. In situ DRIFTS patterns of catalysts exposed to 0.07 vol% NH3 and desorption under N2: (a) VMT; (b) VMT-10; (c) 1CoVMT-10.
Catalysts 13 00844 g008
Figure 9. In situ DRIFTS patterns of catalysts exposed to 0.07 vol% NH3 and 0.07 vol% NO + 5 vol% O2: (a) VMT; (b) VMT-10; (c) 1CoVMT-10. For 1250–1500 cm−1: (a′) VMT; (b′) VMT-10; (c′) 1CoVMT-10.
Figure 9. In situ DRIFTS patterns of catalysts exposed to 0.07 vol% NH3 and 0.07 vol% NO + 5 vol% O2: (a) VMT; (b) VMT-10; (c) 1CoVMT-10. For 1250–1500 cm−1: (a′) VMT; (b′) VMT-10; (c′) 1CoVMT-10.
Catalysts 13 00844 g009
Figure 10. Comparison between the in situ DRIFTS patterns of 0.07 vol% NH3 and 0.07 vol% NO + 5 vol% O2 adsorbed on 1CoVMT-10.
Figure 10. Comparison between the in situ DRIFTS patterns of 0.07 vol% NH3 and 0.07 vol% NO + 5 vol% O2 adsorbed on 1CoVMT-10.
Catalysts 13 00844 g010
Table 1. Pore structure parameters.
Table 1. Pore structure parameters.
SamplepH
Value
Specific Surface Area
A/(m2·g−1)
Pore Volume
v/(cm3·g–1)
Average Pore Diameter
d/(nm)
VMT2.5720.3117.1
VMT-1010780.3718.8
1CoVMT-1010750.3719.9
Table 2. Surface binding energies and ratio of valence state.
Table 2. Surface binding energies and ratio of valence state.
SampleBinding Energy (eV)Ratio of Valence State
Co 2p
Co2+ Co3+
Oα OβV 2p
V4+ V5+
V5+/
V4+
Oα/(Oβ + Oα)Co2+/
Co3+
VMT532.5 529.6515.7 516.81.0214.11
VMT-10532.0 529.8515.7 516.61.1258.34
1CoVMT-10782.2 779.9532.3 529.5515.6 516.71.4916.305.00
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

Wang, R.; Zhang, Y.; 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. https://doi.org/10.3390/catal13050844

AMA Style

Wang R, Zhang Y, 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(5):844. https://doi.org/10.3390/catal13050844

Chicago/Turabian Style

Wang, Ruonan, Yanli Zhang, Xing Fan, and Jian Li. 2023. "Cobalt-/pH-Modified V2O5-MoO3/TiO2 Catalyst with Enhanced Activity for the Low-Temperature Selective Catalytic Reduction Process" Catalysts 13, no. 5: 844. https://doi.org/10.3390/catal13050844

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