Next Article in Journal
Pattern-Dependent Mammalian Cell (Vero) Morphology on Tantalum/Silicon Oxide 3D Nanocomposites
Next Article in Special Issue
Tungsten-Embedded Graphene: Theoretical Study on a Potential High-Activity Catalyst toward CO Oxidation
Previous Article in Journal
Study on the Degradation of Optical Silicone Exposed to Harsh Environments
Previous Article in Special Issue
Metal/Carbon Hybrid Nanostructures Produced from Plasma-Enhanced Chemical Vapor Deposition over Nafion-Supported Electrochemically Deposited Cobalt Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ DRIFTS Studies of NH3-SCR Mechanism over V2O5-CeO2/TiO2-ZrO2 Catalysts for Selective Catalytic Reduction of NOx

Key Laboratory of Energy Thermal Conversion and Control of Ministry of Education, School of Energy and Environment, Southeast University, Nanjing 210096, China
*
Author to whom correspondence should be addressed.
Materials 2018, 11(8), 1307; https://doi.org/10.3390/ma11081307
Submission received: 19 June 2018 / Revised: 23 July 2018 / Accepted: 24 July 2018 / Published: 28 July 2018
(This article belongs to the Special Issue Supported Materials for Catalytic Application)

Abstract

:
TiO2-ZrO2 (Ti-Zr) carrier was prepared by a co-precipitation method and 1 wt. % V2O5 and 0.2 CeO2 (the Mole ratio of Ce to Ti-Zr) was impregnated to obtain the V2O5-CeO2/TiO2-ZrO2 catalyst for the selective catalytic reduction of NOx by NH3. The transient activity tests and the in situ DRIFTS (diffuse reflectance infrared Fourier transform spectroscopy) analyses were employed to explore the NH3-SCR (selective catalytic reduction) mechanism systematically, and by designing various conditions of single or mixing feeding gas and pre-treatment ways, a possible pathway of NOx reduction was proposed. It was found that NH3 exhibited a competitive advantage over NO in its adsorption on the catalyst surface, and could form an active intermediate substance of -NH2. More acid sites and intermediate reaction species (-NH2), at lower temperatures, significantly promoted the SCR activity of the V2O5-0.2CeO2/TiO2-ZrO2 catalyst. The presence of O2 could promote the conversion of NO to NO2, while NO2 was easier to reduce. The co-existence of NH3 and O2 resulted in the NH3 adsorption strength being lower, as compared to tests without O2, since O2 could occupy a part of the active site. Due to CeO2’s excellent oxygen storage-release capacity, NH3 adsorption was weakened, in comparison to the 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalyst. If NOx were to be pre-adsorbed in the catalyst, the formation of nitrate and nitro species would be difficult to desorb, which would greatly hinder the SCR reaction. All the findings concluded that NH3-SCR worked mainly through the Eley-Rideal (E-R) mechanism.

Graphical Abstract

1. Introduction

Generally, nitrogen oxides (NOx), which may cause environmental problems, such as: Photochemical smog; acid rain; ozone depletion; and, health hazards, are mainly emitted from the industrial combustion of fossil fuels. Therefore, the reduction of NOx has become an important research field for atmospheric environmental control. Currently, the selective catalytic reduction (SCR) is the most promising method to reduce the emissions of NOx [1,2]. The temperature window of the traditional V-W (Mo)/Ti catalyst is 300–400 °C, but in some coal-fired power plants, the temperature of exhaust gas is lower. In order to enhance the NOx conversion rate, the exhaust gases were reheated, which caused a large waste of energy. The narrow temperature window restrained its application. Thus, many researchers redirected their study to focus on the catalyst, which has superior low-temperature activity.
The VOx/TiO2 system for SCR has been studied extensively in the past and a number of reaction mechanisms have been proposed. It is generally accepted that the Brønsted and Lewis acid sites are essential for the reaction mechanism. Topsoe et al. [3] proposed a “Brønsted NH4+” mechanism over a V2O5-based catalyst, which has gained the majority of support in the literature. Arnarson et al. [4] observed the SCR reaction over the VO3H/TiO2 catalyst and demonstrated that the Brønsted acid site served to capture the NH3 and increased the NH4+ stability (increased Brønsted acid strength), which impacted the catalytic rate in a negative direction. Marberger et al. [5] had a similar conclusion for the V2O5-WO3/TiO2 catalyst. That is, the Brønsted acid sites hardly contributed to the SCR activity and mainly served as an NH3 pool to replenish the Lewis sites. NO reacted predominantly with NH3 adsorbed in the Lewis acid sites at low temperatures. SCR reactions over Ce-based catalysts mainly followed two mechanisms, one is the Eley–Rideal mechanism (i.e., the reaction of gaseous NO with adsorbed NH3 species), and the other is the Langmuir–Hinshelwood mechanism (i.e., the reaction of adsorbed NOx with adsorbed NH3 species on adjacent sites) [6]. While these two reaction pathways probably do not exclude each other, it is essential to understand whether either or both species are relevant. Vuong et al. [7] reported that NH3-SCR proceeded from a Langmuir–Hinshelwood mechanism on bare supports (TiO2), while an Eley–Rideal mechanism operated on V-containing catalysts.
In recent years, cerium oxides have attracted extensive attention due to their outstanding oxygen storage-release capacity and excellent redox properties in the low-temperature NH3-SCR reactions [8,9,10]. In V/Ce1−xTixO2 catalysts, Ce-O sites are effectively covered by VOx species, which hinder the formation of surface nitrates and cause the switch in the reaction mechanism. Zhang et al. [11] observed the adsorption and reaction processes in DRIFTS (diffuse reflectance infrared Fourier transform spectroscopy) spectra and concluded that the cis-N2O22− formed on CeO2 reacted more favorably with NH3 than with other nitrate species. Galvez et al. [12] demonstrated that the SCR reaction over activated carbon supported the V2O5 catalysts (V2O5/AC) that took place between the adsorbed species of NH3 on the Brønsted acid sites, and the NO molecules in the gaseous phase, following an Eley–Rideal (E–R) mechanism. In Yu et al. [13], the study proposed that the SCR reaction over Zr3 (PO4)2/CeO2-ZrO2 proceeded via the combination of the adjacent, surface NxOy species, and the ads-NH3 species by Langmuir-Hinshelwood (L-H) mechanism. Ma et al. [14] also observed the enhanced NH3 activation and NO3-formation. The latter promoted the reaction of ads-NH3 and ads-NO3-species for the SCR reaction over NbOx/CeO2-ZrO2 catalysts—according to the “L-H” mechanism. Getting to know the reaction pathway and proposing reaction mechanisms is helpful in guiding the design and preparation of the catalysts [15].
In our previous study [9,16], a series of 1 wt. % V2O5-CeO2/TiO2-ZrO2 catalysts with different contents of CeO2were prepared by an impregnation method. It was found that the sample of Ce/Ti = 0.2 (the molar ratio) exhibited a favorable performance with a 92% NOx conversion rate at 250 °C. In addition, the effect of Ce modification on microscopic properties and the catalytic performance of V2O5/TiO2-ZrO2 were investigated in more detail. It concluded that the promotional effect of adding Cemainly laid in the intensified interaction between the metal oxide components and the larger amount of Brønsted and Lewis acid sites, as well as the formation of active intermediates (-NH2). In this study, we further investigated the NH3-SCR mechanism over the optimal 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalyst, and by carrying out transient activity tests and in situ DRIFTS analyses under various conditions of single or mixing feeding gas and pre-treatment ways, proposed a possible reaction pathway.

2. Results and Discussion

2.1. Adsorption and Desorption Properties of NOx and NH3 on the Catalysts

The adsorption-desorption behavior of the catalyst is considered to be a crucial step to a heterogeneous catalysis system. To study the desorption status of the reactant gas on the catalyst surface, the desorption of NO on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 was studied. As shown in Figure 1, the band at 3670 cm−1 was attributed to O-H, and it decreased with the increased temperature until the negative peak appeared. The band at 3203 cm−1 was the result of the hydroxyl vibration. The catalysts contained a little bit of water at a normal temperature, and the water evaporated as the temperature rose; adsorption bands then disappeared. The bands (1618 cm−1, 1367–1378 cm−1, 1245–1288 cm−1, 1130 cm−1, and 1058 cm−1) were ascribed to the adsorbed NOx, especially the band of 1618 cm−1 which was related to weak adsorption of NO and NO2 [16,17,18]. In the case of cis-N2O22−, bands should appear in the 1300–1400 cm−1 [19]. When the temperature exceeded 200 °C, N2O22− appeared in the region of 1367–1378 cm−1 and the intensity of peaks increased as the temperature rose. This confirmed that it could exist stably on the surface of the catalyst. The band at 1245 cm−1 was due to bridging nitrate; the adsorption intensity receded as the temperature rose, and the band region moved to 1288 cm−1 with the generation of monodentate nitrate [17,20,21]. Subsequently, this peak disappeared as the temperature reached 400 °C. The band at 1130 cm−1 was assigned to nitrosyl NO, which could be oxidized to nitrite and nitrate with the existence of oxygen, and it sharply decreased as the temperature increased [22,23]. The band at 1054 cm−1 corresponded to nitrate species, which could exist on the surface of the catalyst stably and was hard to desorb even when the temperature was raised.
As shown in Figure 2a, as N2 was steadily purged on, it was clear that the adsorption of NO was very weak at 250 °C, and N2O22− and nitrate species only appeared at the band of 1371 cm−1 and 1052 cm−1. In addition, there was no significant change in peak intensity by increasing the adsorption and desorption time, indicating that it could exist stably on the surface of the catalyst. As shown in Figure 2b, it was observed that the presence of O2 obviously strengthened the adsorption intensity of NOx on the surface of catalysts. After being exposed to NO + O2 for 60 min, weak adsorption of NO and NO2 appeared at the band of 1630 cm−1. The bands at 1365 cm−1 and 1108 cm−1 could be assigned to cis- and trans-N2O22− [19,24], respectively. Simultaneously, the bands at 1284 cm−1 and 1038 cm−1 were attributed to monodentate nitrate and nitrate species, respectively.
As shown in Figure 3, the adsorption peaks of free O-H appeared at 3662 cm−1 and 3700 cm−1, and the band at 3100–3400 cm−1 was associated with the N-H stretching vibrations, which are linked to Lewis acid sites. The peaks at 1556 cm−1, 1548 cm−1, and 1505 cm−1 corresponded to the formation of intermediate species (-NH2) in SCR reactions [21]. In addition, there was no significant change in peak intensity as adsorption and desorption time increased. According to our previous study [16], intermediate species (-NH2) were detected above 300 °C over the V2O5/TiO2-ZrO2 catalyst. The V2O5-0.2CeO2/TiO2-ZrO2 catalyst exhibited more -NH2 at lower temperatures, which explained its higher activity in comparison to other catalysts. Vuong et al. [11] collected different DRIFTS spectra of bare supports (CeO2, TiO2 and CeO2-TiO2) and supported vanadium catalysts (V/CeO2, V/CeO2-TiO2 and V/TiO2) at 200 °C. They reported additional bands at 1510–1520 cm−1 of NH2, which were only observed on pure CeO2 and CeO2-TiO2. It could be speculated that the addition of Ce was the key factor to affect the surface adsorbed NH3 species. At the same time, the peaks at 1605 cm−1, 1357 cm−1, 1321 cm−1, 1282 cm−1, 1180 cm−1, and 1133 cm−1 were associated with NH3 cooperating vibration—linked to Lewis acid sites [25,26]. According to our previous study [16], with the addition of Ce, the acid sites of the catalysts increased and the optimal V2O5-0.2CeO2/TiO2-ZrO2 sample possessed the largest amount of surface acid sites, which greatly promoted the SCR reaction. The same trend was observed in Vuong et al. [7]. They demonstrated that the relative amount of Lewis acid sites in the V-containing catalysts decreased in the order V/Ce0.5Ti00.5O2 > V/CeO2 > V/TiO2. The band at 1180 cm−1 in Figure 3a split into two NH3 adsorption peaks (1085 cm−1 and 1044 cm−1), and the band at 1133 cm−1 in Figure 3b corresponded to the peak at 1085 cm−1. The band at 1678 cm−1 was associated with NH4+ symmetric vibration and is linked to Brønsted acid sites [27,28]. Comparing Figure 3a with Figure 3b, it can be found that the presence of O2 hindered the adsorption of NH3. However, in Figure 3b, after the feeding of NH3 + O2 was stopped, the intensity of the NH3 adsorption peak, linked to Lewis acid sites, was stronger than in Figure 3a. It might have been caused by the re-adsorption of desorbed ammonia or weak ammonia adsorption on Lewis acid sites, because CeO2 had the capacity of oxygen storage-release and O2 occupied some active sites.

2.2. Transient Response Experiment Analysis

In order to illuminate the difference between NOx species and explain the SCR reaction mechanism, transient reaction studies by in situ DRIFTS spectra were performed. As shown in Figure 4, the NH3 adsorption peak could be found after NH3 and NO were introduced for two min. After adsorption was saturated, the bands at 3400 cm−1, 3100 cm−1, and 1198 cm−1 were associated with NH3 adsorption and was linked to Lewis acid sites. The intermediate species (-NH2) appeared at 1591 cm−1, which implied more active intermediates for the NH3 oxidation reaction. No obvious NOx adsorption was observed; it was a preliminary inference that SCR reactions mainly followed from the Eley–Rideal mechanism.
As shown in Figure 5, catalysts were exposed to the flow of NO and NO + O2 at 250 °C for 60 min. The O-H adsorption peaks appeared at 3510 cm−1 and 3528 cm−1. The N-H stretching vibration peaks appeared in the range of 3400–3100 cm−1 after NO was introduced for two min. However, in Figure 5b, NH3 adsorption peaks appeared after NH3 was introduced for 10 min in the same region. NO2 asymmetric vibration adsorption peaks appeared at 1610 cm−1 and 1620 cm−1 in Figure 5a,b, respectively. The bands at 1583 cm−1, 1226 cm−1, and 1231 cm−1 were ascribed to bridging nitrates. cis-N2O22− appeared at 1353 cm−1 in Figure 5b, and it shifted to the region of 1335 cm−1 with the introduction of NH3, which then weakened the adsorption. When introducing NO + O2 again, the adsorption peak recovered to 1353 cm−1.
NH3 had no obvious influence on the NOx adsorption peak at 2000–1000 cm−1, especially after being exposed to NO + O2 where the influence became tinier. After the pre-adsorption of NO, the intensity of the NOx adsorption peak was obvious, but NH3 adsorption could barely be found. With the introduction of NH3, the N-H stretching vibration was present in the range of 3400–3100 cm−1, as seen in both Figure 5a,b. The results showed that when NO + O2 was injected separately, NO + O2 occupies SCR active reaction sites and restrains the adsorption of NH3, before hindering the SCR reaction. When NO and NO + O2 was reintroduced, respectively, peaks located at 3519 cm−1 and assigned to O-H adsorption were observed. The intensity of the NOx adsorption peak had no decrement; on the contrary, NH3 adsorption, which was linked to Lewis acid sites, disappeared. These results indicated that the gas-phase NOx had reacted with NH3 on Lewis acid sites, which verified the Eley–Rideal mechanism on catalysts. However, in Chen et al. [29], a Langmuir–Hinshelwood mechanism operated on the CeTi catalyst, and adsorbed NH3 and NH4+ that reacted with NO/O2 from the gas phase. Vuong et al. [7] demonstrated that the switch in reaction mechanisms has its roots in the structural differences of catalysts and supports. In V/Ce1−xTixO2 catalysts, Ce-O sites are effectively covered by VOx species, which hinders the formation of surface nitrates and causes the switch in the reaction mechanism.
As shown in Figure 6a, when NO + O2 is introduced, the adsorption peaks at the region of 3400–3100 cm−1 and 1189 cm−1 disappeared, while the O-H adsorption peak (1618 cm−1), the N2O22− adsorption peak (1371 cm−1 and 1112 cm−1), and the nitrate species peak (1024 cm−1) appeared. When NH3 was introduced again, the NO2 adsorption peak disappeared. Moreover, a strong adsorption of NH3 appeared at the region of 3400−3100 cm−1 and 1259 cm−1. Brønsted acid adsorption appeared at 1698 cm−1 and 1428 cm−1 and considerably intensified, while other NOx adsorption had no obvious change.
As shown in Figure 6b, NO was introduced after being exposed to NH3 + O2. Bridging nitrate and monodentate nitrate appeared at 1575 cm−1 and O-H vibration appeared at 3566 cm−1. When NH3 + O2were introduced again, the O-H vibration became stronger and the NH3 adsorption peak at the region of 3400–3100 cm−1 was heavily weakened, as compared with that in Figure 6a. As a result, it can be concluded that O2 reacted with NO first.
In Figure 7, NH3 was introduced first, and then NO was introduced in combination with NH3. Lastly, O2 was also introduced with NH3, NO, and O2 being presented at the same time. In these three different atmospheres, the intensity of NH3 adsorption on Lewis acid sites had no change. Meanwhile, the active intermediate species of -NH2 appeared at 1588 cm−1, indicating that NH3 molecules continued to be adsorbed on the catalytic surface with the process of reaction. The stable existence of intermediate species (-NH2) explained the high SCR activity of the V2O5-0.2CeO2/TiO2-ZrO2 catalyst at low temperatures. Simultaneously, the intensity of the O-H negative peak receded gradually, which might have been caused by the H2O produced in the SCR reaction.

2.3. Steady-State Response Experiments

As shown in Figure 8, catalysts were saturated at 25 °C after 60 min pre-adsorption. -NO2 adsorption appeared at 1839 cm−1 and 1843 cm−1. The bands at 1692 cm−1, 1682 cm−1, 1443 cm−1, and 1419 cm−1 were associated with NH4+ adsorption, linked to Brønsted acid sites, and the bands of 3400–3100 cm−1,1197 cm−1, and 1215 cm−1 were associated with NH3 adsorption, linked to Lewis acid sites. As shown in Figure 8b, N2O22− species appeared at 1106 cm−1 with the presence of O2, which indicated that the existence of O2 would promote NO adsorption. Comparing Figure 8a with Figure 8b, NH3 adsorption became much stronger with the existence of O2. At the same time, the combination of NOx and NH3 appeared at 1248 cm−1, and N2O22− species decreased with increasing temperatures, indicating that O2 is essential for SCR reactions.

2.4. Transient SCR Activity Test Experiments

As shown in Figure 9a, NO was introduced after the pre-adsorption of NH3 for 2 h. The initial conversion of NOx, NO, and NO2 was 61%, 56%, and 97%, respectively. With a steady flow of NOx, adsorbed ammonia was consumed gradually and the conversion of NOx and NO decreased, while NO2 conversion went down-up-down. According to the in situ DRIFTS results, it might be that NO2 is easier to be adsorbed on the catalysts surface, thus leading to the decrease of NH3 adsorption; the conversion of NO2 dropped correspondingly. Until NH3 was completely consumed, NO2 started to be adsorbed on catalysts and the conversion went up, and decreased again after adsorption saturation. By feeding NH3 and NO simultaneously, the conversion rate of NOx and NO was lower than if only NO was fed, suggesting that NO2 occupied active reaction sites resulting in its poor performance. In the case of feeding NH3 and NO at the same time, we found that all the three conversion rates showed the same trend, namely, that the conversion rate reduced after the first rose, which is associated with the promotion of NH3 for SCR reaction. When the three gases: NH3; NO; and, O2 were fed synchronously, the NOx conversion rate reached a stable level of 80%. Simultaneously, the conversion rate of NO and NO2 stabilized at 73% and 92%, respectively. What is more, both conversion rates obviously increased, indicating that NO2 was easier to be reduced. We can conclude that O2 was essential for the SCR reaction.
In Figure 9b, NO was introduced after NH3 and O2 was pre-adsorbed for 2 h. It was obvious that the conversion rate of NOx reduced compared to Figure 9a. However, the conversion rate of NO2 increased and the conversion rate of NO decreased. This might have been caused by the pre-adsorbed O2 reacting with NO and producing NO2, which was easier to react with, and be adsorbed by, the catalysts. The conversion rate increased rapidly when NH3, NO, and O2 was present at the same time.
In Figure 9c, after the pre-adsorption of NO + O2 for 2 h, the denitration efficiency declined continuously with the existence of NH3, NO, and O2. When introducing NH3 and NO together, the conversion rate of NOx and NO went down-up-down, while the conversion rate of NO2 went down-up-down-up. This could be ascribed to the oxygen storage-release capacity of CeO2. NO adsorbed on the catalysts, reacted with O2, and produced NO2, resulting in the ascended e-conversion rate of NO2. When O2 was completely consumed, the conversion rate went down again. In this process, the SCR reaction was very weak. As a result, most of the NO2 adsorbed on the catalysts, so its conversion rate went up. After this, O2 reacted with NO and produced more NO2, and its conversion rate declined after the adsorption of NO2 was saturated. When O2 reacted with NO completely, NO2 occupied the adsorption sites of O2, leading to the conversion rate going up. The combined effect of the NO2 and NO conversion rate resulted in the conversion of NOx going down-up-down-up.

2.5. Low-Temperature SCR Reaction Pathway

The above analyses of in situ DRIFTS have demonstrated the relatively high ability of 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalyst on NH3 adsorption and oxidation. At the reaction temperature (250 °C), the Lewis acid sites were much more stable than were the Brønsted acid sites and the quantity of coordinated NH3 was larger than that of the NH4+ ions. The gaseous NH3 was adsorbed on the catalytic surface, followed by a reaction with the gas phase NO to form the intermediate of NH2NO, which was unstable and would decompose into N2 and H2O (Eley–Rideal mechanism). Based on the combination of in situ DRIFTS experiments and transient SCR activity tests, the mechanism of NH3-SCR reaction over V2O5-CeO2/TiO2-ZrO2 catalysts are mainly as followed:
O 2 + 2 * 2 O * ( * : surface   activesites )
NH 3 ( g ) Ce 4 + NH 3 ( α ) ( Lewis   acid   site )
NH 3 ( α ) + O * NH 2 ( α ) + OH ( α )
NH 2 ( α ) + NO ( g ) NH 2 NO ( α ) N 2 ( g ) + H 2 O ( g )

3. Materials and Methods

3.1. Catalyst Preparation

The Ti-Zr support (molar ratio of Ti:Zr = 1:1) was prepared by a co-precipitation method. Typically, an equal molar amount of TiCl4 solution and ZrOCl2·8H2O was dissolved in the deionized water. NH3·H2O solution was dropped into a stoichiometric solution of TiCl4 and ZrOCl2·8H2O with steady stirring until the pH reached 10. The obtained precipitation solution was aged in air for 24h at room temperature, and then washed with deionized water until the supernatant was free from Cl. Subsequently, the resulting paste was dried at 110 °C for 12 h and then calcined at 450 °C for 4 h in a muffle stove.
1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 samples were prepared by the step-by-step impregnation of Ti-Zr and CeNO3∙6H2O (Ce/Ti = 0.2, molar ratio). The obtained mixture was stirred for 2 h at 25 °C, and then for about 4 h at 85 °C until the water boiled away. The resulting precipitate was dried at 110 °C for 12 h, followed by being calcined at 450 °C for 4 h in a muffle stove to obtain intermediate CeO2/Ti-Zr, which was then impregnated with a NH4VO3 solution. The obtained mixture was dried and calcined in the same process of preparing CeO2/Ti-Zr samples to finally acquire 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 samples.

3.2. In situ DRIFTS Experiments

In situ DRIFTS investigations were carried out on a Nicolet 6700 spectrometer (Thermo Electron Corporation, Waltham, MA, USA), running in the wavenumber range of 400–4000 cm−1 at a resolution of 4 cm−1. A thin, intact and self-supporting wafer of adsorbents were prepared and mounted inside a high temperature cell (HTC-3, Harrick Scientific Corporation, Ithaca, NY, USA). Prior to each experiment, the catalyst was heated to 400 °C under an N2 atmosphere for 1h to remove any adsorbed species, then cooled down to the reaction temperature. The background spectrum was recorded in N2 flow and was automatically subtracted from the sample spectrum during the experiment. Then the N2 flow was switched to a stream containing one or more reactants, such as NH3, NO, and O2. In situ DRIFTS experiments included transient response and steady-state response experiments. It should be noted that new catalyst samples pretreated under the same conditions and were used in each in situ DRIFTS experiments.

3.3. Transient SCR Activity Tests

As shown in Table 1, in order to coordinate the in situ DRIFTS experiments, catalyst activity test experiments were designed under different conditions of feeding gases. A total of 0.3 g of catalyst (screening through 40 to 60 mesh sieve) was tested on a fixed-bed quartz tube reactor (Nanjing University of Technology, Nanjing, China)with an internal diameter of 7 mm at the temperature of 250 °C. The total flow rate was 100 mL/min, which was pre-mixed in a gas mixer to obtain the simulated gas containing 0.08% NO, 0.08% NH3, and 5% O2, with a balance of N2, NO, NO2, and NOx in the outlet, which was continually monitored by a flue gas analyzer (Testo 330-2 LL, Shanghai, China). Typically, during the experiments, about 5% NO was converted to NO2. In other words, 5% NOx existed in the form of NO2.

4. Conclusions

In situ DRIFTS experiments and transient SCR activity tests were used coordinately to observe active and intermediate species and to describe the possible reaction path of 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 at low temperature. The results are as followed:
(1)
NH3 held a dominant position in the competitive adsorption between NH3 and NO. Transient SCR activity tests showed that the NH3 pre-adsorbed catalyst exhibited better SCR activity than its NOx pre-adsorbed counterpart.
(2)
NO might be adsorbed on the catalyst surface and be converted to monodentate nitrite and nitrate species, which is more obvious in the presence of O2, and dramatically restrains the adsorption of NH3, hindering the SCR reaction.
(3)
More acid sites and reaction intermediate species -NH2 at lower temperatures mainly led to the higher activity of the V2O5-0.2CeO2/TiO2-ZrO2 catalyst.
(4)
Transient SCR activity tests and steady-state response experiments both confirmed that NH3-SCR activity was enhanced by the presence of O2. NH3 adsorption intensity had no obvious difference, whether NO or O2 was introduced or not, indicating that the adsorption and consumption of NH3 was in dynamic equilibrium, which promoted SCR reaction.
(5)
NH3-SCR reaction over 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalyst mainly follows the E-R mechanism.

Author Contributions

S.K. and Y.Z. conceived and designed the experiments; X.Y. and T.H. performed the experiments; Y.Z. and B.L. analyzed the data and contributed reagents and materials; Y.Z. wrote the paper.

Funding

This work was supported by the Key Research and Development Projects of Jiangsu Province (BE2017716) and the National Key R&D Plan (2017YFB0603201).

Acknowledgments

This work was carried out in Key Laboratory of Energy Thermal Conversion and Control of Ministry of Education, based in Southeast University. The authors thank for the managers of the instruments in the laboratory.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nova, I.; Ciardelli, C.; Tronconi, E.; Chatterjee, D.; Bandlkonrad, B. NH3-SCR of NO over a V-based catalyst: Low-T redox kinetics with NH3 inhibition. AIChE J. 2006, 52, 3222. [Google Scholar] [CrossRef]
  2. Xu, H.; Shen, Y.; Shao, C.; Lin, F.; Zhu, S.; Qiu, T. A novel catalyst of silicon cerium complex oxides for selective catalytic reduction of NO by NH3. J. Rare Earths 2010, 28, 721–726. [Google Scholar] [CrossRef]
  3. Topsoe, N.Y.; Dumesic, J.A.; Topsoe, H. Vanadia-titania catalysts for selective catalytic reduction of nitric-oxide by ammonia: I.I. Studies of active sites and formulation of catalytic cycles. J. Catal. 1995, 151, 241–252. [Google Scholar] [CrossRef]
  4. Arnarson, L.; Falsig, H.; Rasmussen, S.B.; Lauritsen, J.V.; Moses, P.G. The reaction mechanism for the SCR process on monomer V5+ sites and the effect of modified Bronsted acidity. Phys. Chem. Chem. Phys. 2016, 18, 17071–17080. [Google Scholar] [CrossRef] [PubMed]
  5. Marberger, A.; Ferri, D.; Elsener, M.; Krocher, O. The Significance of Lewis Acid Sites for the Selective Catalytic Reduction of Nitric Oxide on Vanadium-Based Catalysts. Angew. Chem. Int. Ed. 2016, 128, 11989–11994. [Google Scholar] [CrossRef] [PubMed]
  6. Xiong, S.; Liao, Y.; Dang, H.; Qi, F.; Yang, S. Promotion mechanism of CeO2 addition on the low temperature SCR reaction over MnOx/TiO2: A new insight from the kinetic study. RSC Adv. 2015, 5, 27785–27793. [Google Scholar] [CrossRef]
  7. Vuong, T.H.; Radnik, J.; Rabeah, J.; Bentrup, U.; Schneider, M.; Atia, H.; Armbruster, U.; Grunert, W.; Bruckner, A. Efficient VOx/Ce1−xTixO2 catalysts for low-temperature NH3-SCR: Reaction mechanism and active sites assessed by in situ/operando spectroscop. ACS Catal. 2017, 7, 1693–1705. [Google Scholar] [CrossRef]
  8. Zhang, T.; Qu, R.; Su, W.; Li, J. A novel Ce–Ta mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2015, 176, 338–346. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Zhu, X.; Shen, K.; Xu, H.; Sun, K.; Zhou, C. Influence of ceria modification on the properties of TiO2–ZrO2 supported V2O5 catalysts for selective catalytic reduction of NO by NH3. J. Colloid Interface Sci. 2012, 376, 233–238. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, Z.; Zhu, J.; Li, J.; Ma, L.; Woo, S.I. Novel Mn–Ce–Ti Mixed-Oxide catalyst for the selective catalytic reduction of NOx with NH3. ACS Appl. Mater. Interfaces 2014, 6, 14500–14508. [Google Scholar] [CrossRef] [PubMed]
  11. Zhang, L.; Pierce, J.; Leung, V.L.; Wang, D.; Epling, W.S. Characterization of ceria’s interaction with NOx and NH3. J. Phys. Chem. C 2013, 117, 8282–8289. [Google Scholar] [CrossRef]
  12. Galvez, M.E.; Boyano, A.; Lazaro, M.J.; Moliner, R. A study of the mechanisms of NO reduction over vanadium loaded activated carbon catalysts. Chem. Eng. J. 2008, 144, 10–20. [Google Scholar] [CrossRef]
  13. Yu, J.; Si, Z.; Chen, L.; Wu, X.; Weng, D. Selective catalytic reduction of NOx by ammonia over phosphate-containing Ce0.75Zr0.25O2 solids. Appl. Catal. B Environ. 2015, 163, 223–232. [Google Scholar] [CrossRef]
  14. Ma, Z.; Wu, X.; Si, Z.; Weng, D.; Ma, J.; Xu, T. Impacts of niobia loading on active sites and surface acidity in NbOx/CeO2–ZrO2 NH3–SCR catalysts. Appl. Catal. B Environ. 2015, 179, 380–394. [Google Scholar] [CrossRef]
  15. Hu, H.; Zha, K.; Li, H.; Shi, L.; Zhang, D. In situ DRIFTs investigation of the reaction mechanism over MnOx-MOy/Ce0.75Zr0.25O2 (M=Fe, Co, Ni, Cu) for the selective catalytic reduction of NOx with NH3. Appl. Surf. Sci. 2016, 387, 921–928. [Google Scholar] [CrossRef]
  16. Zhang, Y.; Guo, W.; Wang, L.; Song, M.; Yang, L.; Shen, K.; Xu, H.; Zhou, C. Characterization and activity of V2O5-CeO2/TiO2-ZrO2 catalysts for NH3-selective catalytic reduction of NOx. Chin. J. Catal. 2015, 36, 1701–1710. [Google Scholar] [CrossRef]
  17. Underwood, G.M.; Miller, T.M.; Grassian, V.H. Transmission FT-IR and knudsen cell study of the heterogeneous reactivity of gaseous nitrogen dioxide on mineral oxide particles. J. Phys. Chem. A 1999, 103, 6184–6190. [Google Scholar] [CrossRef]
  18. Ramis, G.; Angeles Larrubia, M. An FT-IR study of the adsorption and oxidation of N-containing compounds over Fe2O3/Al2O3 SCR catalysts. J. Mol. Catal. A Chem. 2004, 215, 161–167. [Google Scholar] [CrossRef]
  19. Martínezarias, A.; Soria, J.; Conesa, J.C.; Seoane, X.L.; Arcoya, A.; Cataluna, R. NO reaction at surface oxygen vacancies generated in cerium oxide. J. Chem. Soc. Faraday Trans. 1995, 91, 1679–1687. [Google Scholar] [CrossRef]
  20. Kijlstra, W.S.; Brands, D.S.; Poels, E.K.; Bliek, A. Mechanism of the selective catalytic reduction of NO by NH3 over MnOx/Al2O3.1. Adsorption and desorption of the single reaction components. J. Catal. 1997, 171, 208–218. [Google Scholar] [CrossRef]
  21. Kijlstra, W.S.; Brands, D.S.; Smit, H.I.; Poels, E.K.; Bliek, A. Mechanism of the selective catalytic reduction of NO by NH3 over MnOx/Al2O3.2. Reactivity of adsorbed NH3 and NO complexes. J. Catal. 1997, 171, 219–230. [Google Scholar] [CrossRef]
  22. Hadjiivanov, K.; Knözinger, H. Species formed after NO adsorption and NO + O2 co-adsorption on TiO2: An FTIR spectroscopic study. Phys. Chem. Chem. Phys. 2000, 2, 2803–2806. [Google Scholar] [CrossRef]
  23. Kantcheva, M. Identification, stability, and reactivity of NOx species adsorbed on Titania-Supported manganese catalysts. J. Catal. 2001, 204, 479–494. [Google Scholar] [CrossRef]
  24. Qi, G.; 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]
  25. Long, R.Q.; Yang, R.T. Selective catalytic reduction of NO with Ammonia over Fe3+-Exchanged mordenite (Fe–MOR): Catalytic performance, characterization, and mechanistic Study. J. Catal. 2002, 207, 224–231. [Google Scholar] [CrossRef]
  26. Guan, B.; Lin, H.; Zhu, L.; Huang, Z. Selective catalytic reduction of NOx with NH3 over Mn, Ce substitution Ti0.9V0.1O2−δ nanocomposites catalysts prepared by self-propagating high-temperature synthesis method. J. Phys. Chem. C 2011, 115, 12850–12863. [Google Scholar] [CrossRef]
  27. Jung, S.M.; Grange, P. Investigation of the promotional effect of V2O5 on the SCR reaction and its mechanism on hybrid catalyst with V2O5 and TiO2-SO42− catalysts. Appl. Catal. B Environ. 2002, 36, 207–215. [Google Scholar] [CrossRef]
  28. Centeno, M.A.; Carrizosa, I.; Odriozola, J.A. In situ DRIFTS study of the SCR reaction of NO with NH3 in the presence of O2 over lanthanide doped V2O5/Al2O3 catalysts. Appl. Catal. B Environ. 1998, 19, 67–73. [Google Scholar] [CrossRef]
  29. 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]
Figure 1. In situ DRIFTS (diffuse reflectance infrared Fourier transform spectroscopy) spectra of NO desorption on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 as a function of temperature after the catalyst was exposed to a flow of 800 ppm NO for 60 min at 25 °C.
Figure 1. In situ DRIFTS (diffuse reflectance infrared Fourier transform spectroscopy) spectra of NO desorption on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 as a function of temperature after the catalyst was exposed to a flow of 800 ppm NO for 60 min at 25 °C.
Materials 11 01307 g001
Figure 2. In situ DRIFTS spectra of (a) NO adsorption and (b) NO + O2 adsorption on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2catalysts with N2 purging for various time at 250 °C after the catalysts were exposed to a flow of 800 ppm NO or 800 ppm NO + 5% O2 for 60 min.
Figure 2. In situ DRIFTS spectra of (a) NO adsorption and (b) NO + O2 adsorption on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2catalysts with N2 purging for various time at 250 °C after the catalysts were exposed to a flow of 800 ppm NO or 800 ppm NO + 5% O2 for 60 min.
Materials 11 01307 g002
Figure 3. In situ DRIFTS spectra of (a) NH3 adsorption and (b) NH3 + O2 adsorption on V2O5-0.2CeO2/TiO2-ZrO2 catalysts with N2 purging for various time at 250 °C after the catalysts were exposed to a flow of 800 ppm NH3 or 800 ppm NH3 + 5% O2 for 60 min.
Figure 3. In situ DRIFTS spectra of (a) NH3 adsorption and (b) NH3 + O2 adsorption on V2O5-0.2CeO2/TiO2-ZrO2 catalysts with N2 purging for various time at 250 °C after the catalysts were exposed to a flow of 800 ppm NH3 or 800 ppm NH3 + 5% O2 for 60 min.
Materials 11 01307 g003
Figure 4. In situ DRIFTS spectra of NH3 + NO adsorption on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalysts with N2 purging for various times at 250 °C after the catalysts were exposed to a flow of 800 ppm NH3 and 800 ppm NO for 60 min.
Figure 4. In situ DRIFTS spectra of NH3 + NO adsorption on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalysts with N2 purging for various times at 250 °C after the catalysts were exposed to a flow of 800 ppm NH3 and 800 ppm NO for 60 min.
Materials 11 01307 g004
Figure 5. In situ DRIFTS spectra of the transient reactions at 250 °C between (a) NO and pre-adsorbed NH3 + O2 and (b) NH3, and pre-adsorbed NO + O2 species over 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalysts recorded as a function of time.
Figure 5. In situ DRIFTS spectra of the transient reactions at 250 °C between (a) NO and pre-adsorbed NH3 + O2 and (b) NH3, and pre-adsorbed NO + O2 species over 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalysts recorded as a function of time.
Materials 11 01307 g005
Figure 6. In situ DRIFTS spectra of the transient reactions at 250 °C between (a) NO + O2 and pre-adsorbed NH3, and (b) NO and pre-adsorbed NH3 + O2 species over 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalysts recorded as a function of time.
Figure 6. In situ DRIFTS spectra of the transient reactions at 250 °C between (a) NO + O2 and pre-adsorbed NH3, and (b) NO and pre-adsorbed NH3 + O2 species over 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 catalysts recorded as a function of time.
Materials 11 01307 g006
Figure 7. In situ DRIFTS spectra of NH3 + NO + O2 pre-adsorption transient reaction at 250 °C over 1 wt. %V2O5-0.2CeO2/TiO2-ZrO2 catalysts recorded as a function of time.
Figure 7. In situ DRIFTS spectra of NH3 + NO + O2 pre-adsorption transient reaction at 250 °C over 1 wt. %V2O5-0.2CeO2/TiO2-ZrO2 catalysts recorded as a function of time.
Materials 11 01307 g007
Figure 8. In situ DRIFTS spectra of (a) NH3 + NO desorption and (b) NH3 + NO + O2 desorption on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 as a function of temperature after the catalyst was exposed to a flow of 800 ppm NO, 800 ppm NH3, and 5% O2 for 60 min at 25 °C.
Figure 8. In situ DRIFTS spectra of (a) NH3 + NO desorption and (b) NH3 + NO + O2 desorption on 1 wt. % V2O5-0.2CeO2/TiO2-ZrO2 as a function of temperature after the catalyst was exposed to a flow of 800 ppm NO, 800 ppm NH3, and 5% O2 for 60 min at 25 °C.
Materials 11 01307 g008
Figure 9. Transient SCR (selective catalytic reduction) activity tests (250 °C) under different pre-adsorption conditions: (a) pre-adsorption of NH3; (b) pre-adsorbed of NH3 and O2 and (c) pre-adsorption of NO + O2.
Figure 9. Transient SCR (selective catalytic reduction) activity tests (250 °C) under different pre-adsorption conditions: (a) pre-adsorption of NH3; (b) pre-adsorbed of NH3 and O2 and (c) pre-adsorption of NO + O2.
Materials 11 01307 g009
Table 1. The working conditions of transient SCR activity tests.
Table 1. The working conditions of transient SCR activity tests.
Gas Composition1234
INH3NONH3 + NONH3 + NO + O2
IINH3 + O2NONH3 + NO + O2-
IIINO + O2NH3 + NO + O2NH3 + NO-

Share and Cite

MDPI and ACS Style

Zhang, Y.; Yue, X.; Huang, T.; Shen, K.; Lu, B. In Situ DRIFTS Studies of NH3-SCR Mechanism over V2O5-CeO2/TiO2-ZrO2 Catalysts for Selective Catalytic Reduction of NOx. Materials 2018, 11, 1307. https://doi.org/10.3390/ma11081307

AMA Style

Zhang Y, Yue X, Huang T, Shen K, Lu B. In Situ DRIFTS Studies of NH3-SCR Mechanism over V2O5-CeO2/TiO2-ZrO2 Catalysts for Selective Catalytic Reduction of NOx. Materials. 2018; 11(8):1307. https://doi.org/10.3390/ma11081307

Chicago/Turabian Style

Zhang, Yaping, Xiupeng Yue, Tianjiao Huang, Kai Shen, and Bin Lu. 2018. "In Situ DRIFTS Studies of NH3-SCR Mechanism over V2O5-CeO2/TiO2-ZrO2 Catalysts for Selective Catalytic Reduction of NOx" Materials 11, no. 8: 1307. https://doi.org/10.3390/ma11081307

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