Next Article in Journal
Synthesis of Eugenol Ethyl Ether by Ethylation of Eugenol with Diethyl Carbonate over KF/γ-Al2O3 Catalyst
Next Article in Special Issue
The Catalyst of Ruthenium Nanoparticles Decorated Silicalite-1 Zeolite for Boosting Catalytic Soot Oxidation
Previous Article in Journal
Dual-MOFs-Derived Fe and Mn Species Anchored on Bamboo-like Carbon Nanotubes for Efficient Oxygen Reduction as Electrocatalysts
Previous Article in Special Issue
Research Progress on Metal Oxides for the Selective Catalytic Reduction of NOx with Ammonia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comparative Study on the Effect of Surface and Bulk Sulfates on the High-Temperature Selective Catalytic Reduction of NO with NH3 over CeO2

1
School of Environment, Nanjing Normal University, Nanjing 210023, China
2
School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023, China
3
College of Environmental and Chemical Engineering, Jiangsu University of Science and Technology, Zhenjiang 212000, China
4
Analysis and Testing Central Facility, Anhui University of Technology, Ma’anshan 243002, China
5
Jiangsu Province Engineering Research Center of Environmental Risk Prevention and Emergency Response Technology, Nanjing 210023, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(8), 1162; https://doi.org/10.3390/catal13081162
Submission received: 6 June 2023 / Revised: 20 July 2023 / Accepted: 21 July 2023 / Published: 28 July 2023

Abstract

:
Herein, two CeO2 samples dominantly decorated with surface and bulk sulfates were constructed and their distinct effects on high-temperature NH3-SCR were investigated by strictly controlling the sulfate content at a comparable level. The obtainment of surface and bulk sulfates was revealed using a designed leaching experiment, and further evidenced by the characterization results from XPS and H2-TPR. In comparison with CeO2 modified with bulk sulfates (B-CeS), sufficient acid sites with strong intensity were generated on CeO2 modified with surface sulfates (S-CeS). In addition, due to electron-withdrawing effect from S=O in sulfate species, NH3 oxidation over S-CeS was greatly suppressed, providing an additional contribution to enhanced performance in high-temperature NH3-SCR.

Graphical Abstract

1. Introduction

The large and uncontrolled emission of nitrogen oxides (NOx, x = 1,2) has caused serious environmental problems, and selective catalytic reduction of NOx with NH3 (NH3-SCR) is testified to be one of the most effective methods for NOx elimination from stationary sources and diesel vehicles [1]. Among the diverse catalysts developed for NH3-SCR, the vanadium–titanium system, represented by V2O5-WO3(MoO3)/TiO2, has been commercialized for SCR denitration for decades. The operation temperature for V2O5-WO3(MoO3)/TiO2 catalysts typically lies in the medium temperature zone (300–400 °C), which is compatible with the flue gas temperature at boiler economizer outlets and thus often used for NOx removal from coal-fired power plants [2]. However, the installation of this catalyst in high-temperature denitration is largely restrained, mainly due to the aggravated oxidation of ammonia by active V2O5 and inferior thermal stability of anatase TiO2 [3,4]. From the viewpoint of practical application, there is an urgent need for treatment of exhaust gases via high-temperature SCR in some workplaces, such as diesel engines, gas turbines and waste incinerators [5,6,7].
As early as the 1990s, Topsøe and coworkers pointed out that acid and redox sites are essential for NH3-SCR catalysts [8,9]. Whether for the Eley–Rideal (E-R) or the Langmuir–Hinshelwood (L-H) reaction pathway, NH3 adsorption on a catalyst surface constitutes the first step to initiate a reaction. With an increase in reaction temperature, NH3 adsorption will be disrupted and desorption tends to occur. As such, it is necessary that the catalysts designed for high-temperature SCR should acquire acid sites with sufficient intensity [10]. On the other hand, their redox properties also need to be carefully regulated, since excessive consumption of NH3 via heavy oxidation is bound to reduce the available NH3 for NO reduction. Following these two basic guidelines, a diverse array of catalysts has been explored and such typical combinations like metal ion-exchanged zeolite [11,12] and WO3/TiO2 [3,13] are found to exhibit excellent de-NOx performance even at temperatures higher than 500 °C.
Ceria-based materials, as an important kind of rare-earth oxide, have been extensively studied as efficient V-free NH3-SCR catalysts due to their abundant reserve, good redox property and tunable acidity [14,15]. Particularly, by taking advantage of the fascinating characteristics of rich oxygen vacancies and an easy Ce3+/Ce4+ redox cycle, substantial attention has been paid to CeO2 for application in low-temperature NH3-SCR [16]. Nevertheless, the catalytic behavior at high temperature is less considered. Actually, based on the experimental results of some recent studies, it was shown that by grafting strong acid sites onto the surface of ceria, the NO reduction efficiency at high temperatures (>400 °C) can also be significantly boosted. For example, Yang et al. reported that H3PO4-modified CeO2 exhibited more than 80% NO conversion in the broad temperature range of 250–550 °C [17]. Our recent study revealed that when applying excessive SO2 treatment (72 h) over CeO2, the obtained catalysts displayed excellent activity at 500 °C with NO conversion exceeding 90% [18].
It is well known that under actual flue gas conditions, SO2 is widely present. The interaction of SO2 with CeO2 will inevitably lead to the generation of sulfate species, and depending on the extent of the interaction, surface sulfates and bulk sulfates are expected to be formed [19,20]. Many previous studies have proved that irrespective of the specific structures or morphologies, sulfate species modification is conducive to enhance the acid strength of CeO2 and, simultaneously, disrupt the redox properties to a certain extent [21,22,23,24]. Obviously, this would endow a positive effect on the performance of high-temperature SCR. However, further details are still lacking. For example, what is the difference between surface sulfate and bulk sulfate species on the catalytic performance of CeO2 for high-temperature SCR, it is still obscure. Herein, two sulfate species-modified CeO2 catalysts with comparable sulfur content but distinct existing states were fabricated, and the unique function of surface sulfates and bulk sulfates in high-temperature SCR was explored in detail. Results showed that in comparison with bulk sulfates, surface sulfates played a more significant role in promoting NO conversion at high temperatures, which could be reasonably attributed to the much stronger acid properties and worse redox capacity due to close surface coverage of sulfate species on CeO2.

2. Results and Discussion

2.1. Construction of Target Catalysts

The CeO2 catalysts modified with sulfate species have been widely investigated in NH3-SCR, and conventional fabrication is fulfilled according to a bottom-up route, either through vapor sulfation of CeO2 with SO2 + O2 or via wet impregnation of CeO2 with H2SO4/(NH4)2SO4 [24,25,26]. Both methods are facile in operation and one can obtain surface or bulk sulfates by tuning the exposure time or modulating the H2SO4 concentration. However, it is worth to mention that to explicitly approach the goal of present study, we needed to control the sulfate content at a similar value. This is challenging for the traditional preparation, since bulk sulfates are usually formed at the expense of surface sulfates. In other words, in order to generate bulk sulfates, the introduced sulfur content should be controlled at a very high value. For this reason, an alternative top-down preparation methodology was used to construct the designated catalyst of CeO2 modified with bulk sulfates.
It is well reported that when CeO2 is exposed to a feed stream of SO2 + O2 for a short duration, its surface will be covered by an adsorption layer of sulfates [27]. In the present study, this method was utilized to construct the sample of CeO2 exclusively modified with surface sulfates (S-CeS). However, for the obtainment of CeO2 modified with bulk sulfates, the traditional method of treating CeO2 in a SO2 + O2 stream for a long duration or at a high temperature was not adopted. Alternatively, a novel but facile top-down strategy was performed. That is, directly calcining the sulfated precursor of Ce2(SO4)3·xH2O in air at a high temperature (i.e., 900 °C) to convert the majority of metal salts into CeO2, and the residual is expected to present mainly as bulk sulfates. Obviously, one notable merit of this method is that it can well resolve the dilemma between low sulfur content and bulk state of sulfates.
To confirm the effectiveness of this method, a TG analysis was first performed. It is clearly visible in Figure 1 that two distinct decomposition processes are displayed. The first step is related to a loss of crystalline water. Compared to desorption of crystalline water, a distinct delay is detected for the thermal decomposition of sulfate species. Evidently, decomposition of sulfate groups cannot be initiated until the temperature reaches 700 °C, and a sharp weight loss (29.7%) occurs in the narrow temperature range of 700–850 °C. Notably, by inspecting the TG profile at a much higher temperature (830–1000 °C, inset), it is found that the decomposition is still in progress even when the temperature surpasses 900 °C. This can satisfactorily explain the existence of minor sulfur in the B-CeS sample after the harsh thermal treatment at 900 °C for 1 h.
EDS analysis (Table 1) was performed to approach the actual sulfur content in B-CeS and S-CeS. It can be seen from SEM images (Figure 2a,b) that due to the employment of different preparation routes, variations in the macroscopic morphology of the two samples are present, and S-CeS exhibits a more porous appearance with fine particles and tight arrangement. From Table 1, a S signal is detected and the measured content displays a low but comparable value (Ce/S ratio: 20.44 vs. 20.78) for B-CeS and S-CeS. Since a variation in the S content is expected to bring great distinction in acid and redox properties [19,27], the construction of sulfate CeO2 with comparable sulfate loading provides a reliable comparison of the contribution from different sulfate species on the catalytic performance of ceria in high-temperature SCR.
In order to confirm the existence of different sulfate species in B-CeS and S-CeS, a designed water leaching experiment was carried out. According to previous studies, one notable difference between surface and bulk sulfates comes from their solubility in water [19,28]. Commonly, sulfates are tightly bound to a metal oxide surface and cannot be easily leached away by water. Contrarily, analogue to the salt of cerium sulfate, bulk sulfates interact less with the rigid framework of CeO2 and can thus be easily dissolved in water. Therefore, in the present work, we innovatively utilize this principle to distinguish bulk sulfates and surface sulfates, and with the aid of Raman characterization, the crucial information related to the existing state of sulfates is revealed.
As shown in Figure 3, a sharp band at 463 cm−1, which can be ascribed to the F2g vibrational mode of cubic fluorite CeO2, is dominated in the Raman spectra of B-CeS and S-CeS. In addition, no obvious difference in peak intensity is exhibited, suggesting insignificant disturbance on the bulk structure of CeO2 with sulfate species. By further inspecting the spectra, a weak and broad peak centered around 1170 cm−1 is detected and it actually represents the signal from sulfate species [19,29]. Unfortunately, no obvious variation in peak shape or position can be found for B-CeS and S-CeS. Nevertheless, we can still take advantage of the sensitivity of Raman characterization in revealing sulfate signals to approach the crucial information of S content changes. It is observed that the peak intensity of sulfate in B-CeS-W (water leached sample) is significantly lower than that of B-CeS. Interestingly, a negligible change in the peak intensity of the sulfate signal in S-CeS-W is detected as compared to the pristine S-CeS. This sharp difference in the solubility of the sulfate species can be attributed to the different existing states of the sulfate species. That is, bulk sulfates are widely present in B-CeS, while surface sulfates are dominant in S-CeS.

2.2. XRD Results

To explore any impact from sulfate introduction on the phase structure of CeO2, XRD measurements were performed. It is obvious from Figure 4 that the XRD patterns of B-CeS and S-CeS are represented by the characteristic diffractions from fluorite CeO2 (JCPDS 43-1002), and no peaks from sulfates are detected. Based on these results, we can conclude that in comparison with Raman, XRD shows less sensitivity in revealing the evolution of sulfate species. For S-CeS, the absence of reflection peaks from cerium sulfate is understandable, since the sulfates are well dispersed on CeO2. However, for B-CeS, we suppose the absence of signal from cerium sulfate is mainly due to the low loading or amorphous structure [29].
When it comes to the signal related to CeO2, a clear difference is exhibited. In general, the appearance of sharp peaks suggests high crystallinity of CeO2 in both catalysts. This is basically in agreement with the high-temperature treatment and small surface area shown by the two catalysts (B-CeS: 3.6 m2/g, S-CeS: 10.8 m2/g). Additionally, we can find that although both catalysts are calcined under the same temperature, B-CeS obtains diffraction peaks with a stronger intensity, and the calculated average grain sizes applying the Scherrer equation on the most intense peak at 2θ = 28.6° for B-CeS and S-Ce are 40.2 nm and 31.3 nm, respectively. The results clearly show the sintering behavior of ceria can be influenced by the accompanied anion linked to Ce3+ in the precursor.

2.3. XPS Measurement

To obtain a better understanding of the chemical state of elements on the catalyst surface, XPS characterization was performed. Figure 5a presents the S 2p spectra. No obvious difference in the binding energy of S is exhibited, and the appearance of a typical band at 168.7 eV conveys the sulfur element is in the oxidation state of +6, in accordance with the formation of SO42− [30]. In addition, a smaller surface atomic concentration of S is observed for B-CeS (2.13%) when compared to S-CeS (3.60%). This can be viewed as convincing evidence for the aggregated feature of sulfate species in B-CeS sample.
The XPS spectra of Ce 3d are shown in Figure 5b. The peaks denoted as u′ and v′ are related to Ce3+ species, and u, u″, u‴, v, v″ and v‴ are attributed to Ce4+ species. From Table 1, we can see that the introduction of surface sulfate species induces a significant augmentation of the surface concentration of Ce3+. This is not surprising since Waqif et al. reported that SO2 can act as a reducing agent and lead to the formation of Ce2(SO4)3 on a CeO2 surface [31]. Additionally, the obtainment of a much higher surface Ce3+ concentration can be attributed to the fact that most S in S-CeS are concentrated on the surface with CeO2, and therefore more surface Ce3+ are generated.
Regarding the O 1s spectra, a significant difference in the peak shape is displayed. For B-CeS, two well resolved peaks are detected, and they can be reasonably attributed to the signals from lattice oxygen Oα (529.0–530.0 eV) and surface oxygen Oβ (531.3–531.9 eV). Owing to insufficient contact of the CeO2 surface with bulk sulfates, Oβ shows a much weaker intensity than Oα. However, the case is very different for S-CeS. Oβ undergoes an obvious enhancement with the surface coverage of sulfates. In addition, a third peak, Oγ, located at the binding energy of 532.7–533.5 eV, springs out. According to a previous study, this peak can be related to the signal from surface hydroxyl species [32]. Based on the XPS analysis, we know that the different existing states of sulfate species induce a lot of alteration on the surface features of CeO2, and this is particularly true for the S-CeS sample. For B-CeS, lattice oxygen species from ceria are dominant, while the surface oxygen species from sulfate are enriched in S-CeS.

2.4. H2-TPR Results

To explore the impact from varied sulfate species on the reduction behavior of CeO2, H2-TPR was carried out and is shown in Figure 6. According to previous reports, three species have potential for interaction with H2 over the catalyst of sulfated ceria. They are the surface and bulk oxygen from CeO2 and the sulfur in sulfates [27,33]. In B-CeS, all of the three reduction peaks are displayed, with corresponding H2 consumption at low temperatures (peak α, 300–400 °C,), medium temperatures (peak β, 450–540 °C) and high temperatures (peak γ, 700–800 °C). For S-CeS, peak γ with a comparable intensity is found. This is reasonable since this peak is usually attributed to the reduction of bulk oxygen in the deeper interior of CeO2. However, for the other two peaks, significant changes are displayed. That is, peak α cannot be observed anymore and peak β shifts to higher temperatures, with its signal strongly intensified.
It is well known that for pure CeO2 free from any alien element modification, in addition to the reduction of bulk oxygen, the reduction of surface oxygen at a much lower temperature usually accompanies it [33]. This feature is shown by the appearance of a minor peak centered at 360 °C for the B-CeS sample (peak α). However, for S-CeS, the peak α disappears. In view of the fact that surface sulfate species are abundant, we suppose that the disappearance of the reduction peak from oxygen species on the ceria surface actually reflects that most of them reacted with SO2 in generating surface sulfate species. This can also explain the presence of the large β peak shown for the S-CeS sample.
Based on the comprehensive characterizations using Raman, XPS and H2-TPR, it is concluded that two kinds of sulfate species are established for CeO2. That is, through SO2 + O2 treatment, the surface of CeO2 is exclusively covered with a layer of sulfate. As a sharp contrast, the sulfate species are present mainly as bulk sulfates through directly calcining Ce2(SO4)3 at an elevated temperature.

2.5. NH3-SCR Performance

The NO conversion and N2 selectivity as a function of the reaction temperature for pure CeO2 and sulfated CeO2 at 200–500 °C is shown in Figure 7a,b. As expected, a much poorer activity was exhibited by pure CeO2, with a NO conversion of less than 50% across the entire test range. In addition, the appearance of a negative NO conversion at 500 °C reflects the occurrence of serious NH3 oxidation [32], which is also compatible with the very poor N2 selectivity (Figure 7b). Clearly, the introduction of sulfate species distinctly promotes the catalytic performance. S-CeS shows a more pronounced improvement in reaction efficiency than B-CeS, which acquires more than 90% NO conversion in the high-temperature range of 350–450 °C. When the reaction temperature is further increased, a distinct deactivation was observed for both samples. However, the negative NO conversion was no longer present, demonstrating efficient suppression of NH3 oxidation via sulfate modification. From the results of the activity test, we can tentatively propose that surface sulfates contribute more to promoting high-temperature SCR activity than bulk sulfates. To further confirm this, a direct comparison of the activity from B-CeS and B-CeS-W was achieved by taking advantage of the complete removal of bulk sulfates in the B-CeS-W sample. From Figure 7c, we can see that the catalytic performance at 400–500 °C shows negligible declining after removal of bulk sulfates, although obvious differences in low temperature range are displayed. This can be viewed as solid evidence that in comparison with surface sulfates, the contribution from bulk sulfates in boosting the catalytic performance of CeO2 at high temperatures is insignificant.
To further explore the reaction stability of catalysts operated at high temperatures, the results of NO conversion/N2 selection as a function of reaction time are presented in Figure 7d. Both catalysts exhibit no obvious activity and selectivity change after successive operation at 450 °C for 240 min, revealing a good reaction stability of the sulfated samples.

2.6. The Acidity of Prepared Catalysts (NH3-TPD)

To disclose possible reasons for the different catalytic behaviors of surface and bulk sulfates in promoting high-temperature NH3-SCR performance, the acid and redox properties were explored. Figure 8 presents the NH3-TPD results. In line with previous studies, both catalysts showed a wide range of NH3 desorption [22,24]. Through further examination, two desorption peaks can be differentiated. The low-temperature desorption with a peak centered at 150 °C can be attributed to NH3 adsorbed on the weak acid sites from CeO2. It is well reported that due to the nature of basicity, pure CeO2 possesses weak acidity. With introduction of sulfate species, the acidity is strongly enhanced, as evidenced by the appearance of another desorption peak at a temperature of 263 °C.
Interestingly, despite their comparable sulfate content, the distribution of weak and strong acid sites over the two catalysts varies sharply. For B-CeS, weak acidity is dominant, as can be deduced from the appearance of a maximum NH3 desorption signal at 150 °C (peak α). In view of the fact that insufficient contact is present between sulfate species and CeO2, more surface sites from ceria can be provided for NH3 adsorption. On the other hand, due to the aggregation state of sulfate in B-CeS, fewer sites from sulfates are available for NH3 adsorption. It is reasonable that this changes when bulk sulfates are substituted by surface sulfates. From the XPS and H2-TPR results, we know that abundant sulfates are bound to the CeO2 surface. With this significant difference in sulfate species distribution, we can reasonably conclude that the amount of strong acid sites (peak β) increases at the expense of the weak sites provided by CeO2.

2.7. The Oxidation Properties with Respect to NO and NH3 (NO Oxidation and NH3 Oxidation)

The oxidation properties of catalysts with respect to reactants are investigated in terms of NO oxidation and NH3 oxidation. The NO oxidation performance is shown in Figure 9a. In general, both catalysts show poor activity in converting NO to NO2, which is evidenced by the high temperature required to initiate the reaction (250 °C and the obtained poor maximum conversion efficiency (<20%). It has been previously reported that although ceria has excellent redox properties, it shows poor NO oxidation performance and, due to the strong electron-withdrawing effect from S=O in sulfate group, the modification of sulfate species can induce a further deterioration in NO oxidation performance [30]. From this point of view, it is reasonable that the B-CeS displayed a better NO oxidation performance than the S-CeS, since the main body of ceria is less disrupted by bulk sulfates in B-CeS. By comparing the NO oxidation performance with the high-temperature NH3-SCR results, we can conclude that the ability to oxidize NO is not responsible for the different SCR behaviors at high temperatures.
The results of the NH3 + O2 test are shown in Figure 9b. In line with the NO oxidation performance, the S-CeS catalyst presents a worse oxidation capacity for NH3. It is possible that due to the strong coverage of sulfate species on CeO2, the inherent surface oxygen on ceria is not available for S-CeS. As a result, a poor oxidation ability is exhibited, which can suppress ammonia oxidation to some extent at high temperatures. Based on the above analysis, we can conclude that in addition to the change in acid sites, the reduced oxidation capacity of reductant NH3 constitutes another reason for the better performance of the S-CeS sample in high-temperature NH3-SCR.

3. Materials and Methods

3.1. Catalyst Preparation

CeO2 modified with bulk sulfates was constructed through a novel top-down strategy. Ce2(SO4)3·xH2O was employed directly as the precursor, and with help of a harsh thermal treatment (static air, 900 °C, 1 h), most of the metal salt was converted into CeO2. Notably, a minute amount of precursor was not decomposed and still existed as bulk phase in the mixture. For simplicity, the obtained sample is denoted as B-CeS.
The CeO2 modified with surface sulfates was obtained via a vapor sulfation method. Pure CeO2 derived from thermal treatment of cerium nitrate hydrate (900 °C, 1 h) was exposed to a stream of 2000 ppm SO2 + O2 to obtain the designated sample of CeO2 modified with surface sulfate (S-CeS). By controlling the treatment time, a similar value in sulfate content as B-CeS is guaranteed.
A designed experiment of water leaching was performed to discriminate bulk sulfates and surface sulfates, by taking advantage of their different behaviors in water solubility. The B-CeS and S-CeS catalysts were dropped into distilled water, washed to remove the soluble sulfates, sonicated, filtered and dried to obtain the water-washed samples, which are named B-CeS-W and S-CeS-W, respectively.

3.2. Catalyst Characterization

The specific surface area was characterized using the N2 sorption technique at −196 °C on an Micromeritics ASAP-2020 adsorption apparatus (Norcross, GA, USA). The samples were vacuum-pretreated at 300 °C for 3 h before measurement.
TG was performed on a Netzsch thermal analyzer STA 449C (Erich NETZSCH GmbH & Co.Holding KG, Selb, Germany) with N2 as carrier gas and a test temperature range of 0–1000 °C with a temperature rise rate of 10 °C/min.
Surface morphologies of the catalyst samples were investigated using a FEI NANO SEM430 (Fei company, Lausanne, Switzerland) scanning electron microscope. All SEM images were obtained from secondary electrons with an accelerating voltage of 15 kV. The elemental composition and distribution were determined using an Oxford X-Max20 EDS equipment (Oxford Instruments Private Limited, Oxford, UK).
Raman spectra of samples were collected on a LabRAM Aramis (Japan Hobiba, Kyoto, Japan) Laser Raman spectroscopy with a 532 nm Ar+ laser beam.
XRD was performed on a Philips X’pert Pro diffractometer (Philips Analytical, Overijssel, The Netherlands) with a Ni filter and a Cu target Kα radiation source (wavelength 0.15418 nm). The operating voltage and current were set to 40 kV and 40 mA, and data were collected in the range 2θ = 10-80° with a scanning speed of 10 °C/min.
XPS was performed on a PHI 5000 VersaProbe X-ray photoelectron spectrometer (PHI Corporation, Kanagawa, Japan) with a monochromatic Al Kα radiation source (1486.6 eV) and an acceleration power of 15 kW. Prior to testing, the sample was degassed to below 5 × 10−7 Pa at room temperature in an ultra-high vacuum chamber. Sample charge effects were compensated for by correcting the C 1s spectrum to 284.6 eV. The accuracy of the binding energy was 0.1 eV.
Temperature-programmed reduction by H2 (H2-TPR) was carried out in a quartz U-tube reactor connected to a thermal conduct detector (TCD) with an H2/Ar mixture (7% H2 by volume) as a reductant. A total of 10 mg sample was used for each measurement. Before introducing it to the H2-Ar stream, the sample was pretreated in the N2 stream at 200 °C for 1 h. The H2 consumption profile was collected from room temperature to 900 °C at a rate of 10 °C/min.
The acid properties were determined from temperature-programmed desorption (TPD) of ammonia using a Nicolet IS10 FT-IR spectrometer (Thermo Scientific, Waltham, MA, USA) equipped with a 2 m path-length gas cell (2 L volume). Catalysts were evaluated in a fixed-bed quartz reactor (5 mm internal diameter) operating with a steady-state flow. About 100 mg sample was primarily pretreated using argon (100 mL/min) at 200 °C for 1 h, then it was cooled to room temperature and exposed to 500 ppm NH3/Ar (100 mL/min) for 1 h. Thereafter, excessive physical absorbed ammonia was removed by purging with argon at room temperature for 1 h. All NH3-TPD tests were carried out by increasing the temperature from 30 to 600 °C at a rate of 10 °C/min and an argon flow rate of 100 mL/min.
The NH3/NO oxidation experiments were carried out on a fixed-bed reactor with a gas composition of 500 ppm NH3 (or NO), 5 vol.% O2, and N2 as equilibrium gas at a gas flow rate of 100 mL/min. A certain amount (200 mg) of the sample was loaded into the quartz tube reactor, purged with N2 at 200 °C for 1 h and then cooled to room temperature. After the reaction gas was adsorbed to saturation, the reactions were carried out at different temperatures. The concentrations of NO and NH3 were recorded online using a Thermofisher IS10 FTIR spectrometer, and the catalytic performance data as a function of temperature were collected after being kept stable for 30 min at each temperature point.

3.3. Activity Measurement

The catalytic performance was evaluated in a fixed-bed quartz reactor with 5 mm internal diameter under steady-state flow. The feed gas was 500 ppm NO, 500 ppm NH3, 5% O2, and total gas flow was controlled at 100 mL/min. The WHSV was 60,000 mL/(gh). A total of 100 mg catalyst was sieved through a 40–60 mesh. Before the tests, the catalyst was pretreated in a purified Ar stream at 200 °C for 30 min to avoid surface impurities. Then, the mixed gases were switched on and activity data were collected at every target temperature after stabilizing for 30 min in the range of 200–500 °C. The concentration of effluent gases was analyzed on a Nicolet IS10 FT-IR spectrometer equipped with a 2 m path-length gas cell (2 L volume). The NO conversion was calculated based on the following equations:
NO   conversion   ( % ) = [ NO ] in [ NO ] out [ NO ] in × 100 %
N 2   selectivity   ( % ) = [ NO ] in [ NO ] out + [ NH 3 ] in [ NH 3 ] out [ NO 2 ] out 2 × [ N 2 O ] out [ NO ] in [ NO ] out + [ NH 3 ] in [ NH 3 ] out × 100 %

4. Conclusions

By explicitly controlling the existing states of sulfate species, the effect of surface and bulk sulfates on high-temperature catalytic performance of ceria in NH3-SCR was investigated in the present study. It was found that introduction of sulfate species induced little modification on the bulk structure of CeO2. However, both the adsorption and reaction behaviors were significantly altered. In comparison with pure CeO2, introduction of sulfates notably enhanced the high-temperature NH3-SCR performance. In addition, an apparent difference in reaction efficiency between surface and bulk sulfates was exhibited. Due to close contact between sulfate species and ceria, the active surface oxygen species from ceria were largely eliminated, resulting in alleviated NH3 oxidation for S-CeS. On the other hand, the strong interaction between SO2 and surface oxygen species from ceria created sufficient sulfate species. As a result, more strong acid sites were generated in S-CeS. With the obtainment of these unique properties (more strong acid sites and reduced NH3 oxidation capacity), the catalyst of ceria modified with surface sulfates displayed enhanced catalytic performance in high-temperature NH3-SCR.

Author Contributions

Conceptualization, C.T. (Changjin Tang) and C.T. (Chong Tan); methodology, C.T. (Chong Tan), Z.G., L.C. and J.J.; validation, C.T. (Chong Tan), Z.G., Y.W., B.Z. and L.C.; formal analysis, C.T. (Chong Tan), Z.G. and S.S.; investigation, C.T. (Chong Tan), L.C., H.Z., S.S., C.H. and L.L.; data curation, C.T. (Chong Tan), S.S., Y.W., H.Z. and B.Z.; writing—original draft preparation, C.T. (Chong Tan); writing—review and editing, C.T. (Chong Tan), C.T. (Changjin Tang) and S.S.; supervision, H.Z. and C.T. (Changjin Tang); project administration, C.T. (Changjin Tang); funding acquisition, C.T. (Changjin Tang). All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Science Foundation of China (21976081, 21972062) and the Major Scientific and Technological Project of Bingtuan (2018AA002).

Data Availability Statement

Not applicable.

Acknowledgments

The financial support from the National Science Foundation of China (21976081, 22276097) is greatly acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

Glossary

S-CeSCeO2 modified with surface sulfates
B-CeSCeO2 modified with bulk sulfates
B-CeS-WThe B-CeS catalyst treated using water leaching
S-CeS-WThe S-CeS catalyst treated using water leaching

References

  1. Han, L.; Cai, S.; Gao, M.; Hasegawa, J.-Y.; Wang, P.; Zhang, J.; Shi, L.; Zhang, D. Selective Catalytic Reduction of NOx with NH3 by Using Novel Catalysts: State of the Art and Future Prospects. Chem. Rev. 2019, 119, 10916–10976. [Google Scholar] [CrossRef]
  2. Zhao, S.; Peng, J.; Ge, R.; Wu, S.; Zeng, K.; Huang, H.; Yang, K.; Sun, Z. Research progress on selective catalytic reduction (SCR) catalysts for NO removal from coal-fired flue gas. Fuel Process. Technol. 2022, 236, 107432. [Google Scholar] [CrossRef]
  3. Kobayashi, M.; Miyoshi, K. WO3–TiO2 monolithic catalysts for high temperature SCR of NO by NH3: Influence of preparation method on structural and physico-chemical properties, activity and durability. Appl. Catal. B 2007, 72, 253–261. [Google Scholar] [CrossRef]
  4. Lian, Z.; Deng, H.; Xin, S.; Shan, W.; Wang, Q.; Xu, J.; He, H. Significant promotion effect of the rutile phase on V2O5/TiO2 catalysts for NH3-SCR. Chem. Commun. 2021, 57, 355–358. [Google Scholar] [CrossRef] [PubMed]
  5. Feng, S.; Li, Z.; Shen, B.; Yuan, P.; Wang, B.; Liu, L.; Wang, Z.; Ma, J.; Kong, W. High activity of NH3-SCR at high temperature over W-Zr/ZSM-5 in the exhaust gas of diesel engine. Fuel 2022, 323, 124337. [Google Scholar] [CrossRef]
  6. Park, S.; Choi, G.M.; Tanahashi, M. Demonstration of a gas turbine combustion-tuning method and sensitivity analysis of the combustion-tuning parameters with regard to NOx emissions. Fuel 2019, 239, 1134–1142. [Google Scholar] [CrossRef]
  7. Cao, L.; Wu, X.; Xu, Y.; Lin, Q.; Hu, J.; Chen, Y.; Ran, R.; Weng, D. Ceria-modified WO3-TiO2-SiO2 monolithic catalyst for high-temperature NH3-SCR. Catal. Commun. 2019, 120, 55–58. [Google Scholar] [CrossRef]
  8. Topsoe, N.Y. Mechanism of the selective catalytic reduction of nitric oxide by ammonia elucidated by in situ on-line fourier transform infrared spectroscopy. Science 1994, 265, 1217–1219. [Google Scholar] [CrossRef]
  9. Topsoe, N.Y.; Dumesic, J.A.; Topsoe, H. Vanadia-Titania Catalysts for Selective Catalytic Reduction of Nitric-Oxide by Ammonia. J. Catal. 1995, 151, 241–252. [Google Scholar] [CrossRef]
  10. Imanari, M.; Watanabe, Y. Reduction of NOx in combustion flue gases on tungsten based catalysts. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 1981; Volume 7, pp. 841–852. [Google Scholar]
  11. Byrne, J.; Chen, J.; Speronello, B. Selective catalytic reduction of NOx using zeolitic catalysts for high temperature applications. Catal. Today 1992, 13, 33–42. [Google Scholar] [CrossRef]
  12. Xie, L.; Liu, F.; Ren, L.; Shi, X.; Xiao, F.-S.; He, H. Excellent performance of one-pot synthesized Cu-SSZ-13 catalyst for the selective catalytic reduction of NOx with NH3. Environ. Sci. Technol. 2014, 48, 566–572. [Google Scholar] [CrossRef] [PubMed]
  13. Moon Lee, S.; Su Kim, S.; Chang Hong, S. Systematic mechanism study of the high temperature SCR of NO by NH3 over a W/TiO2 catalyst. Chem. Eng. Sci. 2012, 79, 177–185. [Google Scholar] [CrossRef]
  14. Shan, W.; Liu, F.; Yu, Y.; He, H. The use of ceria for the selective catalytic reduction of NOx with NH3. Chin. J. Catal. 2014, 35, 1251–1259. [Google Scholar] [CrossRef]
  15. Wu, T.; Guo, R.-t.; Li, C.-f.; Pan, W.-g. Recent progress of CeO2-based catalysts with special morphologies applied in air pollutants abatement: A review. J. Environ. Chem. Eng. 2022, 11, 109136. [Google Scholar]
  16. Tang, C.; Zhang, H.; Dong, L. Ceria-based catalysts for low-temperature selective catalytic reduction of NO with NH3. Catal. Sci. Technol. 2016, 6, 1248–1264. [Google Scholar] [CrossRef]
  17. Yi, T.; Zhang, Y.; Li, J.; Yang, X. Promotional effect of H3PO4 on ceria catalyst for selective catalytic reduction of NO by NH3. Chin. J. Catal. 2016, 37, 300–307. [Google Scholar] [CrossRef]
  18. Ji, J.; Han, L.; Song, W.; Sun, J.; Zou, W.; Tang, C.; Dong, L. Promotion effect of bulk sulfates over CeO2 for selective catalytic reduction of NO by NH3 at high temperatures. Chin. Chem. Lett. 2023, 34, 107769. [Google Scholar] [CrossRef]
  19. Zhang, L.; Zou, W.X.; Ma, K.L.; Cao, Y.; Xiong, Y.; Wu, S.G.; Tang, C.J.; Gao, F.; Dong, L. Sulfated Temperature Effects on the Catalytic Activity of CeO2 in NH3-Selective Catalytic Reduction Conditions. J. Phys. Chem. C 2015, 119, 1155–1163. [Google Scholar] [CrossRef]
  20. Xie, R.; Ma, L.; Li, Z.; Qu, Z.; Yan, N.; Li, J. Review of sulfur promotion effects on metal oxide catalysts for NOx emission control. ACS Catal. 2021, 11, 13119–13139. [Google Scholar] [CrossRef]
  21. Lian, Z.; Shan, W.; Wang, M.; He, H.; Feng, Q. The balance of acidity and redox capability over modified CeO2 catalyst for the selective catalytic reduction of NO with NH3. J. Environ. Sci. 2019, 79, 273–279. [Google Scholar] [CrossRef]
  22. Chen, J.; Zhao, W.; Wu, Q.; Mi, J.; Wang, X.; Ma, L.; Jiang, L.; Au, C.; Li, J. Effects of anaerobic SO2 treatment on nano-CeO2 of different morphologies for selective catalytic reduction of NOx with NH3. Chem. Eng. J. 2020, 382, 122910. [Google Scholar] [CrossRef]
  23. Tan, W.; Wang, J.; Yu, S.; Liu, A.; Li, L.; Guo, K.; Luo, Y.; Xie, S.; Gao, F.; Liu, F.; et al. Morphology-Sensitive Sulfation Effect on Ceria Catalysts for NH3-SCR. Top. Catal. 2020, 63, 932–943. [Google Scholar] [CrossRef]
  24. Ji, J.; Jing, M.; Wang, X.; Tan, W.; Guo, K.; Li, L.; Wang, X.; Song, W.; Cheng, L.; Sun, J.; et al. Activating low-temperature NH3-SCR catalyst by breaking the strong interface between acid and redox sites: A case of model Ce2(SO4)3-CeO2 study. J. Catal. 2021, 399, 212–223. [Google Scholar] [CrossRef]
  25. Yao, X.; Wang, Z.; Yu, S.; Yang, F.; Dong, L. Acid pretreatment effect on the physicochemical property and catalytic performance of CeO2 for NH3-SCR. Appl. Catal. A 2017, 542, 282–288. [Google Scholar] [CrossRef]
  26. Chen, S.; Xie, R.; Liu, Z.; Ma, L.; Yan, N. Efficient NOx Reduction against Alkali Poisoning over a Self-Protection Armor by Fabricating Surface Ce2(SO4)3 Species: Comparison to Commercial Vanadia Catalysts. Environ. Sci. Technol. 2023, 57, 2949–2957. [Google Scholar] [CrossRef] [PubMed]
  27. Xie, Q.; Cai, Y.; Zhang, L.; Hu, Z.; Li, T.; Wang, X.; Zeng, Q.; Wang, X.; Sun, J.; Dong, L. Relationships between Adsorption Amount of Surface Sulfate and NH3-SCR Performance over CeO2. J. Phys. Chem. C 2021, 125, 21964–21974. [Google Scholar] [CrossRef]
  28. Jin, R.; Liu, Y.; Wu, Z.; Wang, H.; Gu, T. Relationship between SO2 poisoning effects and reaction temperature for selective catalytic reduction of NO over Mn–Ce/TiO2 catalyst. Catal. Today 2010, 153, 84–89. [Google Scholar] [CrossRef]
  29. Shu, Y.; Aikebaier, T.; Quan, X.; Chen, S.; Yu, H. Selective catalytic reaction of NOx with NH3 over Ce–Fe/TiO2-loaded wire-mesh honeycomb: Resistance to SO2 poisoning. Appl. Catal. B 2014, 150, 630–635. [Google Scholar] [CrossRef]
  30. Ji, J.; Gao, N.; Song, W.; Tang, Y.; Cai, Y.; Han, L.; Cheng, L.; Sun, J.; Ma, S.; Chu, Y. Understanding the temperature-dependent H2O promotion effect on SO2 resistance of MnOx-CeO2 catalyst for SCR denitration. Appl. Catal. B 2023, 324, 122263. [Google Scholar] [CrossRef]
  31. Waqif, M.; Bazin, P.; Saur, O.; Lavalley, J.C.; Blanchard, G.; Touret, O. Study of ceria sulfation. Appl. Catal. B 1997, 11, 193–205. [Google Scholar] [CrossRef]
  32. Gu, T.; Liu, Y.; Weng, X.; Wang, H.; Wu, Z. The enhanced performance of ceria with surface sulfation for selective catalytic reduction of NO by NH3. Catal. Commun. 2010, 12, 310–313. [Google Scholar] [CrossRef]
  33. Giordano, F.; Trovarelli, A.; de Leitenburg, C.; Giona, M. A Model for the Temperature-Programmed Reduction of Low and High Surface Area Ceria. J. Catal. 2000, 193, 273–282. [Google Scholar] [CrossRef]
Figure 1. The TG curve of Ce2(SO4)3·xH2O under static air atmosphere. Inset pattern shows the continued process of thermal degradation at 800–1000 °C.
Figure 1. The TG curve of Ce2(SO4)3·xH2O under static air atmosphere. Inset pattern shows the continued process of thermal degradation at 800–1000 °C.
Catalysts 13 01162 g001
Figure 2. Typical SEM images of (a) B-CeS and (b) S-CeS.
Figure 2. Typical SEM images of (a) B-CeS and (b) S-CeS.
Catalysts 13 01162 g002
Figure 3. (a) Influence of water leaching treatment on the signal change in Raman spectra for B-CeS and S-CeS. (b) Amplification of Raman signal changes.
Figure 3. (a) Influence of water leaching treatment on the signal change in Raman spectra for B-CeS and S-CeS. (b) Amplification of Raman signal changes.
Catalysts 13 01162 g003
Figure 4. The XRD patterns of S-CeS and B-CeS.
Figure 4. The XRD patterns of S-CeS and B-CeS.
Catalysts 13 01162 g004
Figure 5. The XPS spectra of (a) S 2p, (b) Ce 3d and (c) O1s for B-CeS and S-CeS.
Figure 5. The XPS spectra of (a) S 2p, (b) Ce 3d and (c) O1s for B-CeS and S-CeS.
Catalysts 13 01162 g005
Figure 6. The H2-TPR patterns of B-CeS and S-CeS.
Figure 6. The H2-TPR patterns of B-CeS and S-CeS.
Catalysts 13 01162 g006
Figure 7. Effect of different sulfate species on the catalytic performance of CeO2 in NH3-SCR. (a) NO conversion of S-CeS, B-CeS and CeO2, (b) N2 conversion of S-CeS, B-CeS and CeO2, (c) NO conversion of B-CeS and B-CeS-W, (d) N2 conversion of B-CeS and B-CeS-W.
Figure 7. Effect of different sulfate species on the catalytic performance of CeO2 in NH3-SCR. (a) NO conversion of S-CeS, B-CeS and CeO2, (b) N2 conversion of S-CeS, B-CeS and CeO2, (c) NO conversion of B-CeS and B-CeS-W, (d) N2 conversion of B-CeS and B-CeS-W.
Catalysts 13 01162 g007
Figure 8. The NH3-TPD profiles of B-CeS and S-CeS.
Figure 8. The NH3-TPD profiles of B-CeS and S-CeS.
Catalysts 13 01162 g008
Figure 9. (a) NO oxidation and (b) NH3 oxidation results of B-CeS and S-CeS.
Figure 9. (a) NO oxidation and (b) NH3 oxidation results of B-CeS and S-CeS.
Catalysts 13 01162 g009
Table 1. Summary of element analysis from EDS and XPS for B-CeS and S-CeS.
Table 1. Summary of element analysis from EDS and XPS for B-CeS and S-CeS.
SamplesAtomic Concentration (mol.%)Atomic Ratio (%)
CeOSCCe3+/(Ce3+ + Ce4+)Oβ/(Oα + Oβ)Ce/S *
B-CeS18.4258.522.1320.9322.0833.0620.44
S-CeS16.4754.323.625.6140.9774.0720.78
* Ce/S ratio determined from EDS analysis.
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

Tan, C.; Gu, Z.; Sin, S.; Ji, J.; Wang, Y.; Zhu, B.; Cheng, L.; Huang, C.; Li, L.; Zhang, H.; et al. A Comparative Study on the Effect of Surface and Bulk Sulfates on the High-Temperature Selective Catalytic Reduction of NO with NH3 over CeO2. Catalysts 2023, 13, 1162. https://doi.org/10.3390/catal13081162

AMA Style

Tan C, Gu Z, Sin S, Ji J, Wang Y, Zhu B, Cheng L, Huang C, Li L, Zhang H, et al. A Comparative Study on the Effect of Surface and Bulk Sulfates on the High-Temperature Selective Catalytic Reduction of NO with NH3 over CeO2. Catalysts. 2023; 13(8):1162. https://doi.org/10.3390/catal13081162

Chicago/Turabian Style

Tan, Chong, Zhiwen Gu, Songil Sin, Jiawei Ji, Yan Wang, Baiyun Zhu, Lijun Cheng, Chunkai Huang, Lulu Li, Hongliang Zhang, and et al. 2023. "A Comparative Study on the Effect of Surface and Bulk Sulfates on the High-Temperature Selective Catalytic Reduction of NO with NH3 over CeO2" Catalysts 13, no. 8: 1162. https://doi.org/10.3390/catal13081162

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