Next Article in Journal
Production of Sustainable Synthetic Natural Gas from Carbon Dioxide and Renewable Energy Catalyzed by Carbon-Nanotube-Supported Ni and ZrO2 Nanoparticles
Next Article in Special Issue
FeCo-LDH/CF Cathode-Based Electrocatalysts Applied to a Flow-Through Electro-Fenton System: Iron Cycling and Radical Transformation
Previous Article in Journal
Efficient Visible-Light-Driven Photocatalysis of BiVO4@Diatomite for Degradation of Methoxychlor
Previous Article in Special Issue
Current Developments in Ozone Catalyst Preparation Techniques and Their Catalytic Oxidation Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ag-Cu Synergism-Driven Oxygen Structure Modulation Promotes Low-Temperature NOx and CO Abatement

1
State Key Laboratory of Clean and Efficient Coal Utilization, College of Chemistry and Chemical Engineering, Taiyuan University of Technology, Taiyuan 030024, China
2
College of Chemistry and Chemical Engineering, Taiyuan University of Technology, Taiyuan 030024, China
3
National Engineering Laboratory for Multi Flue Gas Pollution Control Technology and Equipment, School of Environment, Tsinghua University, Beijing 100084, China
4
School of Environment and Safety Engineering, North University of China, Taiyuan 030051, China
5
Key Laboratory of Coal Science and Technology, Ministry of Education, Taiyuan University of Technology, Taiyuan 030024, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(7), 674; https://doi.org/10.3390/catal15070674
Submission received: 21 June 2025 / Revised: 5 July 2025 / Accepted: 6 July 2025 / Published: 11 July 2025
(This article belongs to the Special Issue Environmentally Friendly Catalysis for Green Future)

Abstract

The efficient simultaneous removal of NOx and CO from sintering flue gas under low-temperature conditions (110–180 °C) in iron and steel enterprises remains a significant challenge in the field of environmental catalysis. In this study, we present an innovative strategy to enhance the performance of CuSmTi catalysts through silver modification, yielding a bifunctional system capable of oxygen structure regulation and demonstrating superior activity for the combined NH3-SCR and CO oxidation reactions under low-temperature, oxygen-rich conditions. The modified AgCuSmTi catalyst achieves complete NO conversion at 150 °C, representing a 50 °C reduction compared to the unmodified CuSmTi catalyst (T100% = 200 °C). Moreover, the catalyst exhibits over 90% N2 selectivity across a broad temperature range of 150–300 °C, while achieving full CO oxidation at 175 °C. A series of characterization techniques, including XRD, Raman spectroscopy, N2 adsorption, XPS, and O2-TPD, were employed to elucidate the Ag-Cu interaction. These modifications effectively optimize the surface physical structure, modulate the distribution of acid sites, increase the proportion of Lewis acid sites, and enhance the activity of lattice oxygen species. As a result, they effectively promote the adsorption and activation of reactants, as well as electron transfer between active species, thereby significantly enhancing the low-temperature performance of the catalyst. Furthermore, in situ DRIFTS investigations reveal the reaction mechanisms involved in NH3-SCR and CO oxidation over the Ag-modified CuSmTi catalyst. The NH3-SCR process predominantly follows the L-H mechanism, with partial contribution from the E-R mechanism, whereas CO oxidation proceeds via the MvK mechanism. This work demonstrates that Ag modification is an effective approach for enhancing the low-temperature performance of CuSmTi-based catalysts, offering a promising technical solution for the simultaneous control of NOx and CO emissions in industrial flue gases.

1. Introduction

Flue gas generated during the fossil fuel sintering process in iron and steel enterprises contains harmful atmospheric pollutants, such as nitrogen oxides (NOx) and carbon monoxide (CO) [1], which pose serious threats to both environmental quality and human health [2,3]. Among various NOx removal technologies, ammonia-based selective catalytic reduction (NH3-SCR) is widely recognized as the most effective denitrification method [4]. The V2O5-WO3/TiO2 catalyst, known for its excellent catalytic performance, has been extensively applied in industrial NH3-SCR processes [5]. However, this catalyst operates optimally within a temperature range of 320–450 °C and involves the use of V2O5, which becomes highly toxic at elevated temperatures [6]. As a result, its application faces significant challenges when treating sintering flue gas at much lower temperatures (110–180 °C) [7]. In light of increasingly stringent environmental regulations, the development of high-performance denitrification catalysts suitable for low-temperature conditions has become an urgent priority. Meanwhile, catalytic CO oxidation, as a well-established technology for CO control, has been widely studied and implemented [8]. Nevertheless, the design and development of a bifunctional catalyst capable of simultaneously removing both NOx and CO remains a key research focus and challenge in the field of environmental catalysis.
In theory, the selective catalytic reduction of NOx using CO as a reductant (CO-SCR) offers a promising strategy for addressing the simultaneous removal of NOx and CO [9]. However, in practical applications, the presence of a large excess of oxygen in flue gas causes CO to preferentially react with O2 rather than NOx [10], significantly limiting the effectiveness and applicability of the CO-SCR process. A promising alternative is the development of bifunctional catalysts capable of simultaneously promoting both CO oxidation and NH3-SCR reactions under oxygen-rich conditions. These catalysts can utilize the heat generated during CO oxidation to enhance the performance of the NH3-SCR reaction, thereby improving overall energy efficiency [11]. This dual-function strategy not only reduces energy consumption and operational costs but also provides a feasible technical pathway for the synergistic control of NOx and CO emissions in low-temperature flue gas environments.
In recent years, transition metal catalysts such as Cu, Mn, and Fe have demonstrated significant potential in NH3-SCR coupled with CO oxidation due to their excellent catalytic activity, superior thermal stability, and relatively low cost [12]. Among these, copper-based catalysts exhibit particular potential, owing to their multiple redox states and strong redox capabilities, which confer significant advantages in both NH3-SCR [13] and CO oxidation [14] reactions. To further enhance the catalytic performance of copper-based catalysts, extensive modification and optimization efforts have been undertaken by researchers. Liu et al. [15] introduced phosphorus into a CuCeZr (CCZ) catalyst to achieve a balanced combination of acidity and redox properties, significantly enhancing the catalytic activity of the CCZ-P1 catalyst within the temperature range of 250–350 °C. Shen et al. [16] synthesized a molybdenum-doped CuCeOx bifunctional catalyst and found that Mo incorporation enhanced the catalyst’s reducibility, promoted the adsorption and activation of reactant gases, and enabled the CCM3 catalyst to achieve a NOx conversion rate of 91.2% at 225 °C, along with complete CO oxidation over a broad temperature window of 150–300 °C. Lv et al. [17] modified a Cu/TiO2 catalyst with Sm, constructing asymmetric active sites of Cu+-Sm3+-Ov-Ti4+, which effectively facilitated full NO conversion at 200 °C and efficient CO oxidation in the temperature range of 175–300 °C.
Although modified copper-based catalysts such as CuSmTi have exhibited relatively superior low-temperature activity, their performance still does not fully meet the practical requirements for the treatment of sintering flue gas at temperatures below 180 °C. Therefore, further modification strategies are required to enhance their catalytic performance at such conditions. Liu et al. [18] prepared Ag-modified α-MnO2 catalysts for efficient benzene oxidation. The Ag modification enriched the variety of oxygen species and increased the quantity and enhanced the reactivity of lattice oxygen, thereby promoting the reaction efficiency. Similarly, Rao et al. [19] reported that the incorporation of Ag could activate a substantial amount of lattice oxygen in CeO2 nanoparticles. The lattice oxygen activated at the Ag-CeO2 interface significantly improved the catalytic performance. These studies demonstrate that Ag modification plays a crucial role in regulating the structure and distribution of oxygen species, thereby effectively enhancing the overall catalytic performance of the material.
In this study, an Ag-modified AgCuSmTi bifunctional catalyst was successfully synthesized using the sol–gel method, which facilitated effective oxygen structure regulation. The catalyst exhibits excellent NO and CO conversion efficiencies at low temperatures while maintaining favorable N2 selectivity. A suite of characterization techniques was employed to investigate the surface structure and internal chemical valence states, revealing pronounced Ag-Cu interactions within the AgCuSmTi catalyst. Further analysis indicated that these Ag-Cu interactions promote the optimization of oxygen structure and modulate the distribution of surface acid sites, thereby enhancing the low-temperature activity of the AgCuSmTi bifunctional catalyst. This study provides new insights into catalytic materials and a theoretical basis for the synergistic removal of multiple pollutants in sintering flue gas.

2. Results and Discussion

2.1. Evaluation of Catalytic Performance

To regulate the performance of the CuSmTi catalyst in the NH3-SCR coupled with CO oxidation reaction, Ag species were introduced as dopants, and the influence of varying Ag doping levels on catalytic activity was systematically investigated. The catalytic performance of AgxCuSmTi catalysts with different Ag contents (x = 0.5, 0.8, 1.0, 1.2, 1.5) was evaluated to determine the optimal doping ratio, where x denotes the molar ratio of AgNO3 to Cu(NO3)2·3H2O. As shown in Figure S1a, the NOx conversion at 150 °C gradually increases with increasing Ag content, reaching 100% for the Ag1.2CuSmTi catalyst. However, further increasing the Ag doping ratio to x = 1.5 results in a slight decline in catalytic activity. Figure S1b presents the variation in N2 selectivity with Ag loading. With increasing Ag content, the N2 selectivity decreases gradually, and this trend becomes more pronounced at higher doping levels. Nevertheless, the Ag1.2CuSmTi catalyst still maintains a N2 selectivity exceeding 90%, demonstrating a favorable combination of high activity and selectivity. As depicted in Figure S1c, all catalysts achieve complete CO conversion at 175 °C. Based on the overall performance analysis, the Ag1.2CuSmTi catalyst exhibits the best catalytic performance among the series. For simplicity and consistency in subsequent discussions, the Ag1.2CuSmTi catalyst is hereafter referred to as AgCuSmTi.
After determining the optimal silver doping ratio, the effect of Ag modification on catalytic performance was further investigated by comparing the NH3-SCR and CO oxidation activities of AgCuSmTi, CuSmTi, and AgSmTi catalysts. Figure 1a presents the NOx conversion profiles of the three catalysts. The CuSmTi catalyst achieves complete NOx conversion at 200 °C; however, its activity declines significantly at temperatures above 250 °C. Notably, the introduction of Ag markedly enhances low-temperature performance, enabling the AgCuSmTi catalyst to reach 100% NOx conversion at 150 °C (T100% = 150 °C). In contrast, the AgSmTi catalyst, which lacks Cu species, exhibits poor performance, achieving full NOx conversion only between 225 and 250 °C. These results indicate that Ag species strongly interact with CuSmTi and play a crucial role in promoting the NH3-SCR reaction. Figure 1b shows the N2 selectivity of the three catalysts. The AgSmTi catalyst exhibits a sharp decrease in N2 selectivity above 150 °C. Although the addition of Ag slightly lowers the N2 selectivity compared to CuSmTi, the AgCuSmTi catalyst maintains over 90% selectivity across the entire temperature range of 50–300 °C. This suggests that the interaction between Ag and Cu not only enhances catalytic activity but also contributes to the preservation of high N2 selectivity. The CO oxidation performance, shown in Figure 1c, further confirms the beneficial effect of Ag modification. Among the three catalysts, AgCuSmTi exhibits the lowest T100% value. At any given temperature within the tested range, its CO conversion rate is slightly higher than that of CuSmTi, whereas AgSmTi shows relatively poor performance. This indicates that the synergistic interaction between Ag and CuSmTi plays a key role in enhancing CO oxidation activity as well. The TOF values for NO and CO conversion over the AgCuSmTi and CuSmTi catalysts are presented in Figure S2a and Figure S2b, respectively. The results clearly demonstrate that the incorporation of Ag significantly enhances the TOF values associated with the Cu species. Consequently, the AgCuSmTi catalyst exhibits markedly higher catalytic efficiency compared to the CuSmTi catalyst.
To further highlight the outstanding catalytic performance of the AgCuSmTi catalyst, Table S1 summarizes recent studies on Cu-based catalysts applied in NH3-SCR coupled with CO oxidation [15,16,17,19,20,21,22,23,24]. As shown, the α-MnO2-Cu catalyst achieves a NOx conversion rate of 85% at 150 °C and complete CO conversion at 175 °C [25]. In comparison, the AgCuSmTi catalyst developed in this study exhibits a 100% NOx conversion rate at the same temperature of 150 °C. Moreover, the NO conversion rate over the CuOx/α-MnO2 catalyst reaches only 81% at 150 °C. Several other well-reported catalysts, such as MnCuCeOx/γ-Al2O3 [26], CuMgFeO [20], V2O5-CuO/TiO2 [22], CuO/VWTi [11], and CuW/CeZr, exhibit higher temperature requirements for achieving maximum NO and CO conversion. The results demonstrate that the AgCuSmTi catalyst not only exhibits competitive low-temperature catalytic activity compared to traditional systems, but in fact surpasses many of them, highlighting its significant potential for practical applications in low-temperature flue gas treatment.
To evaluate the stability of the AgCuSmTi catalyst, stability tests were carried out at 150 °C and 175 °C over a period of 20 h. The results presented in Figure S3a,b clearly indicate that during the 20 h reaction duration, the conversion rates of NO and CO, as well as the N2 selectivity, remained relatively constant. This demonstrates that the AgCuSmTi catalyst exhibits excellent stability under reaction conditions. In summary, the synergistic interaction between Ag and CuSmTi significantly enhances the catalytic performance in both the NH3-SCR and CO oxidation reactions. The AgCuSmTi catalyst displays outstanding low-temperature activity, excellent N2 selectivity, strong CO conversion capability, and superior stability, highlighting its significant potential for practical applications.

2.2. Structural and Chemical Characterization of the Catalyst

2.2.1. Morphological Characterization of the Catalysts

The SEM images (Figure 2a–f) reveal that both the AgCuSmTi and CuSmTi catalysts exhibit irregular nanoparticle morphologies. The incorporation of Ag effectively suppresses nanoparticle agglomeration and promotes the uniform dispersion of active species, thereby enhancing the overall catalytic performance [27]. The particle size distribution of the AgCuSmTi catalyst is presented in Figure 2g. The results indicate that the particle sizes are predominantly within the range of 14–18 nm and are relatively uniformly distributed. However, the average particle size of AgCuSmTi is 16.77 nm, which is significantly larger than that of CuSmTi (6.22 nm; see Figure S4). This suggests that the introduction of Ag has somewhat altered the particle size characteristics of the catalyst. TEM images (Figure 2h,i) show a clear lattice fringe with an interplanar spacing of approximately 0.35 nm, consistent with the (101) plane of anatase TiO2 [28]. Notably, no distinct lattice fringes corresponding to Ag, Cu, or Sm species were observed, indicating that these metal dopants are highly dispersed on the catalyst surface rather than forming large crystalline domains or aggregates.
To further investigate the elemental composition and surface distribution characteristics of the AgCuSmTi catalyst, energy-dispersive X-ray spectroscopy (EDX) mapping analysis was conducted. As shown in Figure 2j, elements such as Ag, Cu, Sm, Ti, and O are clearly present and exhibit a uniform spatial distribution, with no signs of agglomeration. This result indicates that all metallic components are highly dispersed on the catalyst surface. This uniform elemental distribution not only confirms the successful integration and dispersion of multiple metal components but also indicates the presence of well-distributed active sites, which are crucial for efficient catalytic performance. Additionally, the high dispersion of various elements suggests that there may be strong synergistic interactions between different active components, which could help enhance catalytic activity.

2.2.2. Evidence for Ag-Cu Interaction

To further investigate the structural characteristics of the catalysts, XRD analysis was conducted, and the results are presented in Figure 3a. All samples exhibited diffraction peaks corresponding to the anatase phase of TiO2. Notably, no characteristic diffraction peaks associated with Sm, Cu, or Ag species were observed, indicating that these elements are either amorphous or highly dispersed within the matrix [29]. Compared to pure TiO2, the intensity of the diffraction peaks in the SmTi sample was significantly reduced, suggesting that Sm doping inhibits the growth of TiO2 crystallites. Upon further introduction of Cu to form the CuSmTi catalyst, the peak intensities remained nearly unchanged relative to SmTi. However, upon co-doping with Ag, Cu, and Sm to obtain AgCuSmTi, the diffraction peak intensities decreased further, implying a further reduction in the average grain size of TiO2. The smaller grain size facilitates the dispersion of active species, leading to an increased number of active sites and thereby enhancing the catalytic performance [30]. This phenomenon can be attributed to the lattice distortion and increased defect density resulting from the synergistic interaction among Ag, Cu, and Sm species, which collectively suppress the growth of TiO2 grains.
The crystal structures of the samples were further analyzed using Raman spectroscopy. As shown in Figure 3b, all catalysts exhibit the characteristic Raman bands of anatase TiO2 at approximately 143 cm−1 (Eg), 396 cm−1 (B1g), 515 cm−1 (A1g), and 638 cm−1 (Eg) [31]. These results confirm that doping does not induce any phase transformation of TiO2 or lead to the formation of detectable secondary oxide phases, consistent with the XRD findings. However, after doping with Sm, Cu, and Ag, the Eg mode of TiO2 undergoes noticeable blue shifts, exhibits reduced peak intensity, and shows an increased full width at half maximum (FWHM), reflecting enhanced local lattice distortions and internal stress within the TiO2 lattice [32]. Further analysis of the changes in the Eg peak position, intensity, and FWHM reveals that the degree of lattice disturbance follows the order TiO2 > SmTi > CuSmTi > AgCuSmTi. This trend indicates that the co-doping of Ag and Cu introduces the most significant structural perturbation. This effect is likely due to the large differences in ionic radii between Cu2+ (0.73 Å) and Ag+ (1.26 Å) compared to Ti4+ (0.64 Å), which generate considerable lattice strain when incorporated into the TiO2 structure [33]. Moreover, Cu doping tends to create localized stress fields, while the addition of Ag further weakens the Ti-O bond strength, thereby suppressing and broadening the Raman-active vibrational modes [33,34]. In particular, the most pronounced Eg peak broadening is observed for the AgCuSmTi sample. Therefore, the combined interaction of Ag, Cu, and Sm significantly disturbs the TiO2 lattice, resulting in blue shifting of the Eg peak, reduced intensity, and increased peak width. These structural modifications provide a fundamental basis for the enhanced catalytic performance observed in the AgCuSmTi catalyst.
Figure S5 presents the nitrogen adsorption–desorption isotherms and corresponding pore size distribution curves of the samples. As shown in Figure S5a, all catalysts display IV-type nitrogen adsorption isotherms accompanied by H2-type hysteresis loops, indicating that they possess typical mesoporous structural characteristics. The pore size distributions calculated from the desorption branches further support this conclusion [35]. As illustrated in Figure S5b, the pore size distribution of the SmTi catalyst is only slightly affected by Cu modification. However, upon co-doping with Ag and Cu, a noticeable change in the pore size distribution is observed. The data summarized in Table S2 indicate that the incorporation of Ag and Cu leads to an increase in average pore size and a broader pore size distribution. This suggests that the synergistic interaction among Ag, Cu, and Sm alters the aggregation behavior of the particles, thereby influencing both the pore structure and specific surface area. Moreover, by comparing the catalytic performance with the specific surface area data, it can be inferred that the specific surface area is not the dominant factor affecting catalytic performance.
To further investigate the chemical states of the elements in the AgCuSmTi catalyst and its reference samples, XPS analysis was carried out. As shown in Figure 3c, the Ag 3d spectrum exhibits two main peaks corresponding to the Ag 3d3/2 and Ag 3d5/2 spin–orbit components. The binding energies at 368.3 eV and 367.6 eV are assigned to metallic Ag0 and oxidized Ag+ species in the Ag 3d3/2 orbital, respectively [36]. According to the atomic ratio calculations presented in Table S3, the Ag+/(Ag0 + Ag+) ratio in the AgCuSmTi catalyst is higher than that in the AgSmTi catalyst upon Cu doping. Moreover, the binding energy of Ag+ in the AgCuSmTi sample shifts toward a higher value compared to that in AgSmTi. These observations suggest that the introduction of Cu promotes the oxidation of Ag0 to Ag+, indicating possible electron transfer interactions among Ag, Cu, and Sm, which may enhance the electronic affinity of Ag.
The Cu 2p spectrum is displayed in Figure 3d. The peaks at 934.2 eV and 932.5 eV correspond to Cu2+ and Cu+ species in the Cu 2p3/2 orbital, respectively [36,37]. Compared with the CuSmTi catalyst, the proportion of Cu2+ in the AgCuSmTi catalyst is significantly increased, as summarized in Table 1. This indicates that the incorporation of Ag improves the stability of the Cu in its respective oxidation states, further supporting the presence of synergistic electronic interactions between Ag, Cu, and Sm.
As shown in the Sm 3d spectrum (Figure 3e), all catalysts exhibit two fitted peaks at 1083.7 eV and 1081.4 eV, corresponding to Sm3+ and Sm2+ species, respectively [38]. Compared to the CuSmTi catalyst, the Sm3+/(Sm2+ + Sm3+) ratio in the AgCuSmTi catalyst remains nearly unchanged. This result suggests that the incorporation of Ag has little influence on the oxidation state of Sm, implying the absence of a strong direct interaction between Ag and Sm. Instead, the primary effect of Ag doping appears to be on its interaction with Cu species, rather than forming a direct synergistic relationship with Sm.

2.2.3. Structural Regulation of Oxygen

To gain deeper insight into the regulation of surface oxygen species and their role in catalytic performance, a comprehensive analysis of the oxygen structure in the catalyst was conducted. First, O 1s XPS was employed to qualitatively and quantitatively distinguish between lattice oxygen (Olatt) and surface-adsorbed oxygen (Oads). As shown in Figure 4a, the O 1s spectrum can be deconvoluted into two characteristic peaks at 529.7–529.9 eV and 531.3–531.5 eV, corresponding to lattice oxygen (Oβ) and surface-adsorbed oxygen (Oα), respectively [38]. The quantitative results summarized in Table 1 reveal that the AgCuSmTi catalyst exhibits a higher proportion of lattice oxygen compared to the CuSmTi catalyst. This enhancement may be attributed to the electronic synergy between Ag and Cu, which strengthens metal–oxygen bonding and increases the content of lattice oxygen within the crystal structure. The increased lattice oxygen content suggests that the AgCuSmTi catalyst possesses a greater reservoir of active oxygen species available during the initial reaction stage. This facilitates the rapid oxidation of reactant molecules and enhances the overall reaction rate. Concurrently, the consumption of lattice oxygen leads to the formation of oxygen vacancies, which can be readily replenished by gaseous oxygen, thereby sustaining efficient and continuous catalytic cycles [29].
Subsequently, O2-TPD analysis was carried out on both the CuSmTi and AgCuSmTi catalysts to evaluate the mobility and desorption behavior of surface oxygen species, reflecting their potential involvement in oxidation reactions. As presented in Figure 4b, both catalysts exhibit two distinct desorption peaks: one in the low-temperature region (<300 °C), assigned to surface-adsorbed oxygen (Oα), and another in the medium-to-high temperature range (300–600 °C), corresponding to lattice oxygen (Oβ) [20]. Notably, the AgCuSmTi catalyst shows a significant increase in lattice oxygen content, from 33.48% in CuSmTi to 48.25% after the introduction of Ag, indicating that Ag promotes the enrichment of lattice oxygen (Table S4). Therefore, the incorporation of Ag likely activates lattice oxygen through electronic interactions with Cu, enhancing both the reaction rate and stability of the catalyst.
To further validate changes in oxygen vacancy concentration, EPR spectroscopy was employed for sample analysis. As illustrated in Figure 4c, all catalysts display a distinct EPR signal at g = 2.003, which is characteristic of oxygen vacancies [39]. Among them, the CuSmTi catalyst exhibits the strongest signal intensity, indicating a high concentration of oxygen vacancies on its surface. In contrast, the AgCuSmTi catalyst shows a significantly weakened EPR signal, suggesting a reduction in oxygen vacancy density. It can be inferred that the interaction between Ag and Cu contributes to the suppression of oxygen defects while simultaneously facilitating the mobilization and enrichment of lattice oxygen.
In summary, the synergistic effect between Ag and Cu drives the structural regulation of oxygen species, not only creating favorable conditions for oxygen migration and release but also improving the accessibility and utilization efficiency of reactive oxygen species during the reaction. These effects collectively enhance the catalytic activity of the AgCuSmTi system in low-temperature NH3-SCR processes.

2.2.4. Redox Properties and Surface Acidity Characterization

To further investigate the influence of active species on the catalyst surface and their reduction behavior in relation to catalytic performance, H2-TPR characterization was carried out on all catalysts, with the results presented in Figure 5a. The reduction profile of the CuSmTi catalyst can be divided into three distinct stages: the low-temperature peak (131 °C) corresponds to the reduction of Cu2+ to Cu+; the medium-temperature peak (149 °C) is attributed to the reduction of Cu+ to Cu0; and the high-temperature peak (178 °C) is associated with the reduction of CuO clusters [40]. These findings indicate that Cu species exist in multiple oxidation states and undergo progressive reduction to the metallic state through sequential electron transfer processes. Upon Ag doping, all three reduction peaks in the AgCuSmTi catalyst shift toward lower temperatures, suggesting a notable enhancement in the reducibility of Cu species. This improvement can be attributed to the synergistic interaction between Ag and Cu, which facilitates electron transfer and promotes the activation of oxygen species. Consequently, Cu2+ becomes more readily reduced to Cu+ and ultimately to Cu0. Additionally, a new reduction peak appears at approximately 228 °C for the AgCuSmTi catalyst, which is assigned to the reduction of Ag2O to metallic Ag0 [41]. The enhanced Ag–Cu interaction accelerates the redox reaction kinetics at lower temperatures, thereby improving the overall redox activity of the catalyst.
The acid characteristics of the catalyst surfaces were investigated using NH3-TPD analysis for the AgCuSmTi and CuSmTi samples. The results are shown in Figure 5b, with the desorption peak areas quantitatively analyzed and summarized in Table 2. Both catalysts exhibit three distinct NH3 desorption peaks: the α peak (100–150 °C) corresponds to weak Lewis acid sites; the β peak (150–300 °C) represents strong Lewis acid sites; and the γ peak (>300 °C) is associated with Brønsted acid sites [42]. Comparative analysis reveals a significant increase in the proportion of the β peak in the AgCuSmTi catalyst, from 68.15% in CuSmTi to 76.86%, indicating that Ag incorporation substantially enhances the density of strong Lewis acid sites. This facilitates NH3 adsorption and activation, providing abundant active sites for subsequent surface reactions. In contrast, the proportion of the γ peak decreases in AgCuSmTi, reflecting a reduction in Brønsted acidity. These observations suggest that the Ag-Cu interaction plays a crucial role in modulating the surface acidic structure, favoring the formation of strong Lewis acid sites that are more conducive to the NH3-SCR reaction mechanism [43]. As a result, the catalytic performance is significantly improved.

2.3. Proposed Reaction Mechanism Investigation

2.3.1. Proposed Reaction Mechanism of CO Oxidation

To elucidate the intermediate species and reaction mechanisms involved in CO oxidation over the AgCuSmTi composite metal oxide catalyst, in situ DRIFTS experiments were carried out under CO + O2 co-adsorption conditions at various temperatures. The results are presented in Figure 6a. Upon exposure to CO and O2, two main carbonate species formed via interactions between CO and surface oxygen species were observed: monodentate carbonate (1329 cm−1 and 1463 cm−1) [16,44] and bidentate carbonate (1042 cm−1 and 1581 cm−1) [45,46]. A distinct peak at 2113 cm−1 was assigned to Cu+-CO, whose intensity decreased with increasing temperature and vanished entirely above 150 °C. A peak at 2163 cm−1 was attributed to Cu2+-CO [47,48]. Additionally, broad bands in the range of 2285-2390 cm−1 corresponded to CO2 molecules, indicating the formation of gaseous CO2 as a result of CO oxidation [49]. As shown in the spectra, the band at 1581 cm−1, associated with bidentate carbonate, along with the peak at 1329 cm−1, representing monodentate carbonate, both diminished significantly with rising temperature. However, these features remained detectable even at 300 °C. In contrast, the peaks at 1463 cm−1 and 1042 cm−1 exhibited an opposite trend, showing increased intensity at higher temperatures. Considering the evolution of CO2 and Cu+-CO signals, it can be inferred that the thermal decomposition of carbonate species contributes significantly to the oxidation of CO at elevated temperatures.
The involvement of molecular oxygen in the CO oxidation mechanism was examined by comparing DRIFTS spectra obtained at 50 °C under conditions with and without O2, as shown in Figure 6b. Notably, the intensity of the Cu+-CO peak at 2113 cm−1 is significantly lower under CO + O2 co-adsorption compared to CO-only adsorption, suggesting that O2 partially occupies the active sites responsible for CO adsorption. Meanwhile, the spectral region between 1200 and 1700 cm−1, which includes the carbonate-related peaks, remains largely unchanged regardless of O2 presence. This indicates that O2 does not directly influence the formation of carbonate species [15]. Notably, without molecular oxygen present, CO reacts with lattice oxygen, resulting in the production of carbonate intermediates. Based on the above observations, the CO oxidation reaction over the AgCuSmTi catalyst proceeds primarily through the Mars–van Krevelen (MvK) mechanism.

2.3.2. Proposed Reaction Mechanism of NH3-SCR

To investigate the acid sites and intermediate species formed on the catalyst upon incorporation of Ag species, NH3 adsorption–desorption DRIFTS measurements were carried out at various temperatures. Figure 7a presents the in situ DRIFTS spectra of NH3 adsorption on the AgCuSmTi catalyst as a function of surface temperature from 50 to 300 °C. The broad band observed in the range of 3000–3400 cm−1 is attributed to the N-H stretching of adsorbed NH3 [50,51]. Additionally, two distinct peaks at 1668 cm−1 and 1479 cm−1 are assigned to NH4+ species formed via NH3 protonation on Brønsted acid sites [52], whereas the bands at 1195 cm−1 and 1606 cm−1 correspond to NH3 coordinated to Lewis acid sites [50,53]. It is evident that the intensity of the Lewis acid site-related peaks is significantly higher than that of the Brønsted acid site peaks, indicating the dominant role of Lewis acidity in NH3 adsorption on the AgCuSmTi catalyst. As the temperature increases, the intensities of these NH3 adsorption features gradually decrease, suggesting progressive desorption and activation of NH3. Notably, the spectral at 1308 cm−1 and 1520 cm−1 can be assigned to the formation of -NH2 species generated via dehydrogenation of NH3 [54,55]. Furthermore, as the temperature increases, a new band emerges at 1429 cm−1 in the DRIFTS spectrum of the AgCuSmTi catalyst (Figure 7a), which is absent in the CuSmTi catalyst (Figure S6). This spectral feature can be attributed to the formation of -NH species on the Ag-doped catalyst surface [56]. -NH intermediates are likely involved in subsequent redox reactions with NO molecules, potentially leading to the formation of N2O. This observation provides a plausible explanation for the reduced N2 selectivity observed upon Ag doping.
To further explore the interaction between NO and O2 on the catalyst surface, in situ DRIFTS analysis of NO + O2 adsorption was also conducted under the same temperature conditions. The results are shown in Figure 7b. Several nitrate species were identified based on their characteristic vibrational frequencies. The bands at 1600 cm−1 and 1243 cm−1 are attributed to bridged nitrate species [57,58], while the peak at 1575 cm−1 corresponds to bidentate nitrates [59]. The signals at 1439 cm−1 and 1278 cm−1 are associated with monodentate nitrate species [15,60]. As clearly observed in the spectra, the intensity of the monodentate nitrate peaks is significantly higher compared to that of bridged and bidentate nitrates. According to previous studies, the preferential formation of monodentate nitrates is beneficial for enhancing low-temperature NH3-SCR performance [61,62]. This finding supports the conclusion that the AgCuSmTi catalyst exhibits excellent catalytic activity at low temperatures.
To elucidate the reaction mechanism of NH3-SCR over the AgCuSmTi catalyst, in situ DRIFTS transient experiments were carried out. In the first set of experiments, NO + O2 was introduced after NH3 had been pre-adsorbed on the catalyst surface. The results are presented in Figure 7c. Following NH3 pre-adsorption on the AgCuSmTi catalyst, characteristic peaks corresponding to Lewis acid sites (L-acid) were observed at 1177 cm−1 and 1605 cm−1, along with a peak at 1518 cm−1 attributed to -NH2 species and another at 1428 cm−1 assigned to -NH species. After 30 min of N2 purging, the intensities of the L-acid-related peaks slightly decreased, indicating that weakly physisorbed NH3 was partially desorbed. Upon introduction of NO + O2 for 1 min, slight decreases in the intensities of the L-acid and -NH3 peaks were observed, suggesting that gaseous NO and O2 reacted with adsorbed NH3 species, consistent with an Eley–Rideal (E-R) reaction mechanism. At 5 min after NO + O2 exposure, new spectral features emerged corresponding to nitrate species, bridged nitrate (1255 cm−1), monodentate nitrate (1481 cm−1), and bidentate nitrate (1541 cm−1), confirming successful adsorption of NO on the catalyst surface [16]. As the reaction time increased, the intensities of these nitrate-related peaks continued to rise, while the peak at 1605 cm−1 (attributed to L-acid sites) gradually diminished and eventually disappeared after 30 min. The reaction occurring 5 min after NO + O2 introduction can be interpreted as the interaction between adsorbed NH3 and adsorbed NO species, indicating that a Langmuir–Hinshelwood (L-H) mechanism also operates on the catalyst surface. In summary, the AgCuSmTi catalyst exhibits both the E-Rand L-H mechanisms during the NH3-SCR process.
To further validate the above findings, a second set of experiments was conducted with a reversed reactant introduction order, where NH3 was introduced after NO + O2 had been pre-adsorbed on the catalyst. The results are shown in Figure 7d. Following 30 min of NO + O2 pre-adsorption, characteristic peaks corresponding to bidentate nitrate (1545 cm−1) and monodentate nitrate (1481 cm−1 and 1267 cm−1) were clearly observed. After introducing NH3 for 1 min, a new peak at 1169 cm−1 appeared, which is associated with L-acid centers. By 5 min, the intensities of the nitrate-related bands began to decrease, while signals corresponding to -NH2 species and Brønsted acid (B-acid) centers emerged. This evolution indicates that the adsorbed NO species interacted with the incoming NH3 via the L-H mechanism.
Figure 8 presents the proposed reaction pathway for the AgCuSmTi catalyst during the simultaneous removal of NO and CO. During the CO oxidation process, CO molecules adsorbed on the catalyst surface react with lattice oxygen to form either CO2 or carbonate species. Upon desorption of CO2, oxygen vacancies are generated, which are subsequently replenished by molecular oxygen, thereby facilitating continuous CO oxidation through lattice oxygen migration [63]. The detailed reaction mechanism is summarized as follows:
CO ( gas )     CO ( ads )
CO ads   +   O 2 latt     CO 2 g   +   v
CO ads   +   2 O 2 latt     CO 3 2 + v
CO 3 2     CO 2   +   O 2
O 2 g + v     2 O 2 latt
Here, “□v” represents an oxygen vacancy.
The NH3-SCR reaction mechanism over the AgCuSmTi catalyst is predominantly governed by the L-H mechanism, with the E-R mechanism playing a secondary role. In this process, gaseous NH3 is adsorbed and activated on the catalyst surface to form -NH2 intermediates (Equations (1)–(4)). Activated lattice oxygen facilitates the generation of -NH2 species and enhances the redox cycling during the reaction. Subsequently, the -NH2 intermediate reacts with either gaseous NO (Equation (5)) or adsorbed nitrate species (Equations (6) and (7)) to yield N2 and H2O. During the reaction, the -NH2 species can undergo further dehydrogenation to form -NH intermediates, which then react with NO species to produce N2O (Equations (8) and (9)). The formation of N2O reduces the overall N2 selectivity of the catalyst.
The detailed reaction mechanism is summarized as follows:
NH 3 ( g )     NH 3 ( ads )
O 2 ( g )     O 2 ( latt )
O 2 ( latt )     2 O 2
NH 3 ads + O 2     - NH 2 ( ads ) + - OH ( ads )
- NH 2 ( ads ) + NO ( g )     N 2 + H 2 O
NO ( g )     NO ( ads )
- NH 2 ( ads ) + NO ( ads )     NH 2 NO     N 2 + H 2 O
- NH 2 ( ads ) + O 2     - NH ( ads ) + - OH ( ads )
- NH ( ads ) + NO ( g )     N 2 O + H +

3. Experimental Section

3.1. Catalyst Preparation

All catalysts were synthesized using the sol–gel method. In a typical procedure, 25.2 mL of n-butanol and 2.8 mL of glacial acetic acid were first mixed uniformly, followed by the gradual addition of 2.4 mL of deionized water under continuous stirring for 30 min to obtain solution A. Subsequently, 1.88 g of Sm(NO3)3·6H2O, 0.68 g of Cu(NO3)2·3H2O, and a precisely weighed amount of AgNO3 were added to solution A. The mixture was stirred continuously until complete dissolution of all components. Then, 9.6 mL of titanium tetrabutyltitanate (C16H36O4Ti) was slowly introduced into the homogeneous solution, which was further stirred at room temperature for 3 h to obtain a viscous gel-like solution. The obtained sol was then aged at ambient temperature until gelation occurred. Afterward, the gel was transferred to an oven and dried at 110 °C overnight. Finally, the dried precursor was calcined in a muffle furnace at 500 °C for 4 h to obtain the AgxCuSmTi composite metal oxide catalysts with different Ag loadings (x = 0.5, 0.8, 1.0, 1.2, 1.5).
For comparative analysis, other catalytic systems were prepared following similar procedures. In one case, 9.6 mL of C16H36O4Ti was added directly to solution A without incorporating any metal nitrates. The mixture was then dried and calcined under the same conditions to yield pure TiO2. Additionally, three reference catalysts, SmTi, CuSmTi, and AgSmTi, were synthesized by sequentially adding the following combinations to solution A: (1) 1.88 g of Sm(NO3)3·6H2O, (2) 1.88 g of Sm(NO3)3·6H2O and 0.68 g of Cu(NO3)2·3H2O, and (3) 1.88 g of Sm(NO3)3·6H2O and a certain amount of AgNO3. Each mixture was processed through the same drying and calcination steps as described above.

3.2. Catalyst Activity Test

The catalytic performance was evaluated using a fixed-bed reactor system. Detailed descriptions of the experimental setup, operating procedures, test conditions, and conversion rate calculations are provided in the Supporting Information.

3.3. Reagents and Gases

The reagents used in the preparation of the catalysts and the gases employed in the evaluation of their catalytic performance are listed in the Supporting Information.

3.4. Catalyst Characterization

The physical properties of the catalyst were investigated using a range of characterization techniques, including X-ray diffraction (XRD), N2 adsorption, scanning electron microscopy (SEM), transmission electron microscopy (TEM), and Raman spectroscopy. Furthermore, the redox properties and acidic characteristics of the catalyst were systematically analyzed using X-ray photoelectron spectroscopy (XPS), H2 temperature-programmed reduction (H2-TPR), O2 temperature-programmed desorption (O2-TPD), NH3 temperature-programmed desorption (NH3-TPD), electron paramagnetic resonance (EPR), and in situ DRIFTS. Additional experimental details are provided in the Supporting Information.

4. Conclusions

This study presents an innovative approach in which Ag is introduced to modify the CuSmTi catalyst, resulting in the successful development of a dual-functional AgCuSmTi catalyst capable of effectively regulating oxygen structure. A combination of characterization techniques, including XRD, N2 adsorption, and XPS, revealed a pronounced synergistic interaction between Ag and Cu, which plays a critical role in enhancing the catalyst’s low-temperature activity. Firstly, the Ag-Cu synergy significantly improves the physical properties of the catalyst by optimizing the dispersion of active sites, thereby establishing a structural foundation for its superior performance. Secondly, this synergistic effect markedly enhances the mobility of lattice oxygen. The highly active lattice oxygen not only facilitates redox cycling during the reaction but also accelerates electron transfer between metal species. Furthermore, the Ag-Cu interaction effectively modulates the acidic properties of the catalyst surface, increasing the density of Lewis acid sites, which are particularly favorable for the NH3-SCR reaction. This enhancement promotes the adsorption and activation of reactant molecules, contributing to improved catalytic performance. These structural and chemical modifications significantly enhanced the performance of the AgCuSmTi catalyst, enabling complete conversion of NO and CO at temperatures as low as 150 °C and 175 °C, respectively, along with excellent N2 selectivity. To further elucidate the underlying reaction mechanisms, in situ DRIFTS studies were conducted for both the NH3-SCR and CO oxidation reactions. The results indicate that the NH3-SCR process predominantly follows the L-H mechanism, with the E-R mechanism also playing a secondary role. Meanwhile, the CO oxidation reaction proceeds via the MvK mechanism. Overall, the results presented herein offer both a fundamental understanding and a promising theoretical foundation for the simultaneous elimination of nitrogen oxides and carbon monoxide from low-temperature industrial flue gases.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal15070674/s1, Figure S1: NOx conversion, N2 selectivity, and CO conversion of catalysts with different proportions; Figure S2: TOF values for NO and CO conversions of the catalyst; Figure S3: Stability test (20 h) of AgCuSmTi catalyst at 150 °C and 175 °C; Figure S4: TEM spectrum of CuSmTi catalyst; Figure S5: N2 adsorption and desorption isotherms and pore size distribution curves of catalysts; Figure S6: In situ DRIFTS on CuSmTi catalyst: changes in the NH3 surface reaction with temperature when NH3 is continuously introduced; Table S1: Summary of preparation methods, operating reaction conditions, catalytic performance metrics, and reaction mechanism for Cu-based catalysts in NOx and CO abatement processes [11,15,16,17,21,22,23,24,25,26,38,64]; Table S2: Physical properties of catalysts; Table S3: Chemical state of Ag species on catalyst surface; Table S4: Distribution of desorbed oxygen species of catalyst.

Author Contributions

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

Funding

This work was supported by the Key Technologies Research and Development Program, China (No. 2022YFC3701900) and Applied Basic Research Project of Shanxi Province (Nos. 202203021211178, 202303021211032, 202303021211034, and 202403021211217).

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Li, C.; Han, Q.; Zhu, T.; Xu, W. Catalytic NO Reduction by CO over Ca–Fe Oxides in the Presence of O2 with Sintering Flue Gas Circulation. Ind. Eng. Chem. Res. 2020, 59, 20624–20629. [Google Scholar] [CrossRef]
  2. Gan, L.; Shi, W.; Li, K.; Chen, J.; Peng, Y.; Li, J. Synergistic Promotion Effect between NOx and Chlorobenzene Removal on MnOx–CeO2 Catalyst. ACS Appl. Mater. Interfaces 2018, 10, 30426–30432. [Google Scholar] [CrossRef]
  3. Qiao, B.; Wang, A.; Yang, X.; Allard, L.F.; Jiang, Z.; Cui, Y.; Liu, J.; Li, J.; Zhang, T. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat. Chem. 2011, 3, 634–641. [Google Scholar] [CrossRef]
  4. Jiang, L.; Xu, Y.; Jiang, W.; Pu, Y.; Yang, L.; Dai, Z.; Yao, L. The promotion of NH3-SCR performance by AC addition on Mn-Fe/Z catalyst. Sep. Purif. Technol. 2023, 322, 124274. [Google Scholar] [CrossRef]
  5. Zhu, M.; Lai, J.-K.; Tumuluri, U.; Wu, Z.; Wachs, I.E. Nature of Active Sites and Surface Intermediates during SCR of NO with NH3 by Supported V2O5–WO3/TiO2 Catalysts. J. Am. Chem. Soc. 2017, 139, 15624–15627. [Google Scholar] [CrossRef]
  6. Choi, J.-H.; Kim, J.-H.; Bak, Y.-C.; Amal, R.; Scott, J. Pt-V2O5-WO3/TiO2 catalysts supported on SiC filter for NO reduction at low temperature. Korean J. Chem. Eng. 2005, 22, 844–851. [Google Scholar] [CrossRef]
  7. Wang, Y.-Z.; Zhang, J.-L.; Liu, Z.-J.; Du, C.-B. Recent Advances and Research Status in Energy Conservation of Iron Ore Sintering in China. JOM 2017, 69, 2404–2411. [Google Scholar] [CrossRef]
  8. Dey, S.; Dhal, G.C.; Mohan, D.; Prasad, R. Advances in transition metal oxide catalysts for carbon monoxide oxidation: A review. Adv. Compos. Hybrid. Mater. 2019, 2, 626–656. [Google Scholar] [CrossRef]
  9. Li, X.; Ren, S.; Jiang, Y.; Chen, Z.; Wang, L.; Liu, M.; Chen, T. In situ IR spectroscopy study of NO removal over CuCe catalyst for CO-SCR reaction at different temperature. Catal. Today 2023, 418, 114082. [Google Scholar] [CrossRef]
  10. Hamada, H.; Haneda, M. A review of selective catalytic reduction of nitrogen oxides with hydrogen and carbon monoxide. Appl. Catal. A Gen. 2012, 421–422, 1–13. [Google Scholar] [CrossRef]
  11. Zeng, Y.; Rong, W.; Zhang, S.; Wang, Y.; Zhong, Q. Promoting NH3-SCR denitration via CO oxidation over CuO promoted V2O5-WO3/TiO2 catalysts under oxygen-rich conditions. Fuel 2022, 323, 124357. [Google Scholar] [CrossRef]
  12. Zhang, X.; Zhang, G.; Liu, J.; Li, G.; Zhao, Y.; Wang, Y.; Lv, Y. Effects of defective structure originating from N incorporation-evaporation of Co-based biomass carbon catalysts on methane dry reforming. Fuel 2024, 357, 129752. [Google Scholar] [CrossRef]
  13. Wang, X.; Sun, Y.; Han, F.; Zhao, Y. Effect of Fe addition on the structure and SCR reactivity of one-pot synthesized Cu-SSZ-13. J. Environ. Chem. Eng. 2022, 10, 107888. [Google Scholar] [CrossRef]
  14. Shen, Z.; Xing, X.; She, Y.; Wang, S.; Lv, M.; Li, J.; Li, H. Revealing of K and SO2 poisoning mechanism on CuCeOx catalyst for low-temperature CO oxidation. Chem. Eng. J. 2024, 485, 149373. [Google Scholar] [CrossRef]
  15. Liu, Q.; Mi, J.; Chen, X.; Wang, S.; Chen, J.; Li, J. Effects of phosphorus modification on the catalytic properties and performance of CuCeZr mixed metal catalyst for simultaneous removal of CO and NOx. Chem. Eng. J. 2021, 423, 130228. [Google Scholar] [CrossRef]
  16. Shen, Z.; Xing, X.; She, Y.; Guo, P.; Ren, S.; Niu, W.; Li, J.; Li, H.; Meng, H. Unveiling the promoting mechanism of Mo on the performance of CuCeOx catalyst for simultaneously NH3-SCR denitration and CO oxidation under oxygen-rich conditions. Sep. Purif. Technol. 2025, 355, 129561. [Google Scholar] [CrossRef]
  17. Lv, D.; Liu, J.; Zhang, G.; Wang, Y.; Zhao, Y.; Li, G. Novel insights for simultaneous NOx and CO Removal: Cu+-Sm3+-Ov-Ti4+ asymmetric active site promoting NH3-SCR coupled with CO oxidation reaction. Chem. Eng. J. 2024, 481, 148534. [Google Scholar] [CrossRef]
  18. Liu, Y.; Zhou, X.; Zhan, J.; Zhang, X.; Lian, H.-Y.; Zhou, H.; Yi, X.; Yang, H.-H.; Shan, J. Enhancing water resistance of α-MnO2 catalyst for efficient airborne benzene oxidation at mild temperatures via facile Ag incorporation. J. Environ. Chem. Eng. 2023, 11, 110725. [Google Scholar] [CrossRef]
  19. Rao, R.; Liang, H.; Hu, C.; Dong, H.; Dong, X.; Tang, Y.; Fang, S.; Ling, Q. A melamine-assisted pyrolytic synthesis of Ag-CeO2 nanoassemblys for CO oxidation: Activation of Ag-CeO2 interfacial lattice oxygen. Appl. Surf. Sci. 2022, 571, 151283. [Google Scholar] [CrossRef]
  20. Md Saad, S.K.; Ali Umar, A.; Ali Umar, M.I.; Tomitori, M.; Abd Rahman, M.Y.; Mat Salleh, M.; Oyama, M. Two-Dimensional, Hierarchical Ag-Doped TiO2 Nanocatalysts: Effect of the Metal Oxidation State on the Photocatalytic Properties. ACS Omega 2018, 3, 2579–2587. [Google Scholar] [CrossRef]
  21. Ren, S.; Li, X.; He, C.; Chen, L.; Wang, L.; Li, F. Surface tailoring on bifunctional CuOx/MnO2 catalyst to promote the selective catalytic reduction of NO with NH3 and oxidation of CO with O2. Sep. Purif. Technol. 2024, 346, 127471. [Google Scholar] [CrossRef]
  22. Li, X.; Lai, C.; Zhang, Y.; Wang, S.; Ding, S. Bifunctional catalysts V-Cu/TiO2 for selective catalytic reduction of NOx and CO oxidation under oxygen-rich conditions. Mol. Catal. 2024, 569, 114574. [Google Scholar] [CrossRef]
  23. Gui, R.; Xiao, J.; Gao, Y.; Li, Y.; Zhu, T.; Wang, Q. Simultaneously achieving selective catalytic reduction of NOx with NH3 and catalytic oxidation of CO with O2 over one finely optimized bifunctional catalyst Mn2Cu1Al1Ox at low temperatures. Appl. Catal. B Environ. 2022, 306, 121104. [Google Scholar] [CrossRef]
  24. Li, Z.; Cheng, H.; Zhang, X.; Ji, M.; Wang, S.; Wang, S. CuW/CeZr Catalysts: A Dual-Function Catalyst for Selective Catalytic Reduction of NO and CO Oxidation Under Oxygen-Rich Conditions. Catal. Lett. 2021, 151, 3361–3371. [Google Scholar] [CrossRef]
  25. Wang, L.; Wang, B.; Guo, Y.; Zheng, Y.; Zhu, T. Interactions between CO oxidation and selective catalytic reduction of NO with NH3 over Mn-based catalysts. Catal. Sci. Technol. 2022, 12, 4776–4788. [Google Scholar] [CrossRef]
  26. Wang, J.; Xing, Y.; Su, W.; Li, K.; Zhang, W. Bifunctional Mn-Cu-CeOx/γ-Al2O3 catalysts for low-temperature simultaneous removal of NOx and CO. Fuel 2022, 321, 124050. [Google Scholar] [CrossRef]
  27. Resitoglu, I.A.; Keskin, A. Hydrogen applications in selective catalytic reduction of NOx emissions from diesel engines. Int. J. Hydrog. Energy 2017, 42, 23389–23394. [Google Scholar] [CrossRef]
  28. Gao, C.; Shi, J.-W.; Fan, Z.; Wang, B.; Wang, Y.; He, C.; Wang, X.; Li, J.; Niu, C. “Fast SCR” reaction over Sm-modified MnOx-TiO2 for promoting reduction of NOx with NH3. Appl. Catal. A Gen. 2018, 564, 102–112. [Google Scholar] [CrossRef]
  29. Wang, B.; Wang, M.; Han, L.; Hou, Y.; Bao, W.; Zhang, C.; Feng, G.; Chang, L.; Huang, Z.; Wang, J. Improved Activity and SO2 Resistance by Sm-Modulated Redox of MnCeSmTiOx Mesoporous Amorphous Oxides for Low-Temperature NH3-SCR of NO. ACS Catal. 2020, 10, 9034–9045. [Google Scholar] [CrossRef]
  30. Sun, H.; Wang, H.; Qu, Z. Construction of CuO/CeO2 Catalysts via the Ceria Shape Effect for Selective Catalytic Oxidation of Ammonia. ACS Catal. 2023, 13, 1077–1088. [Google Scholar] [CrossRef]
  31. Djaoued, Y.; Badilescu, S.; Ashrit, P.V.; Bersani, D.; Lottici, P.P.; Robichaud, J. Study of Anatase to Rutile Phase Transition in Nanocrystalline Titania Films. J. Sol-Gel Sci. Technol. 2002, 24, 255–264. [Google Scholar] [CrossRef]
  32. Alhomoudi, I.A.; Newaz, G. Residual stresses and Raman shift relation in anatase TiO2 thin film. Thin Solid. Film. 2009, 517, 4372–4378. [Google Scholar] [CrossRef]
  33. Choudhury, B.; Dey, M.; Choudhury, A. Defect generation, d-d transition, and band gap reduction in Cu-doped TiO2 nanoparticles. Int. Nano Lett. 2013, 3, 25. [Google Scholar] [CrossRef]
  34. Liu, H.; Fan, Z.; Sun, C.; Yu, S.; Feng, S.; Chen, W.; Chen, D.; Tang, C.; Gao, F.; Dong, L. Improved activity and significant SO2 tolerance of samarium modified CeO2-TiO2 catalyst for NO selective catalytic reduction with NH3. Appl. Catal. B Environ. 2019, 244, 671–683. [Google Scholar] [CrossRef]
  35. Guo, X.; Lin, F.; Gao, M.; Dai, Q.; Zhan, W.; Wang, L.; Guo, Y.; Wang, A.; Guo, Y. Enhanced catalytic performance and N2 yield of Ag/CeO2 catalyst by Cu modification for NVOCs removal. Chem. Eng. J. 2024, 488, 151117. [Google Scholar] [CrossRef]
  36. Chen, C.; Cao, Y.; Liu, S.; Chen, J.; Jia, W. SCR catalyst doped with copper for synergistic removal of slip ammonia and elemental mercury. Fuel Process. Technol. 2018, 181, 268–278. [Google Scholar] [CrossRef]
  37. Wang, M.; Ren, S.; Jiang, Y.; Su, B.; Chen, Z.; Liu, W.; Yang, J.; Chen, L. Insights into co-doping effect of Sm and Fe on anti-Pb poisoning of Mn-Ce/AC catalyst for low-temperature SCR of NO with NH3. Fuel 2022, 319, 123763. [Google Scholar] [CrossRef]
  38. Zang, P.; Liu, J.; He, Y.; Zhang, G.; Li, G.; Wang, Y.; Lv, Y. LDH-derived preparation of CuMgFe layered double oxides for NH3-SCR and CO oxidation reactions: Performance study and synergistic mechanism. Chem. Eng. J. 2022, 446, 137414. [Google Scholar] [CrossRef]
  39. Liu, S.; Xue, W.; Ji, Y.; Xu, W.; Chen, W.; Jia, L.; Zhu, T.; Zhong, Z.; Xu, G.; Mei, D.; et al. Interfacial oxygen vacancies at Co3O4-CeO2 heterointerfaces boost the catalytic reduction of NO by CO in the presence of O2. Appl. Catal. B: Environ. 2023, 323, 122151. [Google Scholar] [CrossRef]
  40. Yao, X.; Zhang, L.; Li, L.; Liu, L.; Cao, Y.; Dong, X.; Gao, F.; Deng, Y.; Tang, C.; Chen, Z.; et al. Investigation of the structure, acidity, and catalytic performance of CuO/Ti0.95Ce0.05O2 catalyst for the selective catalytic reduction of NO by NH3 at low temperature. Appl. Catal. B Environ. 2014, 150–151, 315–329. [Google Scholar] [CrossRef]
  41. Qu, Z.; Wang, Z.; Quan, X.; Wang, H.; Shu, Y. Selective catalytic oxidation of ammonia to N2 over wire–mesh honeycomb catalyst in simulated synthetic ammonia stream. Chem. Eng. J. 2013, 233, 233–241. [Google Scholar] [CrossRef]
  42. Wang, X.; Qin, M.; Xu, Y.; Li, Q. Improvement of Cu-SAPO-34 hydrothermal stability by tuning P/Al ratio for selective catalytic reduction of NO by NH3. J. Colloid. Interface Sci. 2023, 638, 686–694. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, B.; Li, Z.; Huang, Q.; Chen, M.; Xu, L.; Shen, Y.; Zhu, S. Synergetic removal of elemental mercury and NO over TiCe0.25Sn0.25Ox catalysts from flue gas: Performance and mechanism study. Chem. Eng. J. 2019, 360, 990–1002. [Google Scholar] [CrossRef]
  44. Kang, R.; Wei, X.; Bin, F.; Wang, Z.; Hao, Q.; Dou, B. Reaction mechanism and kinetics of CO oxidation over a CuO/Ce0.75Zr0.25O2-δ catalyst. Appl. Catal. A Gen. 2018, 565, 46–58. [Google Scholar] [CrossRef]
  45. Qi, L.; Yu, Q.; Dai, Y.; Tang, C.; Liu, L.; Zhang, H.; Gao, F.; Dong, L.; Chen, Y. Influence of cerium precursors on the structure and reducibility of mesoporous CuO-CeO2 catalysts for CO oxidation. Appl. Catal. B Environ. 2012, 119–120, 308–320. [Google Scholar] [CrossRef]
  46. Davó-Quiñonero, A.; Bailón-García, E.; López-Rodríguez, S.; Juan-Juan, J.; Lozano-Castelló, D.; García-Melchor, M.; Herrera, F.C.; Pellegrin, E.; Escudero, C.; Bueno-López, A. Insights into the Oxygen Vacancy Filling Mechanism in CuO/CeO2 Catalysts: A Key Step Toward High Selectivity in Preferential CO Oxidation. ACS Catal. 2020, 10, 6532–6545. [Google Scholar] [CrossRef]
  47. Lu, J.; Wang, J.; Zou, Q.; He, D.; Zhang, L.; Xu, Z.; He, S.; Luo, Y. Unravelling the Nature of the Active Species as well as the Doping Effect over Cu/Ce-Based Catalyst for Carbon Monoxide Preferential Oxidation. ACS Catal. 2019, 9, 2177–2195. [Google Scholar] [CrossRef]
  48. Ma, Y.; Li, Y.; Liang, P.; Min, X.; Sun, T. SnO2-modified CuMnOx catalysts for humid CO oxidation at low temperature. J. Ind. Eng. Chem. 2024, 138, 300–310. [Google Scholar] [CrossRef]
  49. Sasmaz, E.; Wang, C.; Lance, M.J.; Lauterbach, J. In situ spectroscopic investigation of a Pd local structure over Pd/CeO2 and Pd/MnOx–CeO2 during CO oxidation. J. Mater. Chem. A 2017, 5, 12998–13008. [Google Scholar] [CrossRef]
  50. Dann, E.K.; Gibson, E.K.; Blackmore, R.H.; Catlow, C.R.A.; Collier, P.; Chutia, A.; Erden, T.E.; Hardacre, C.; Kroner, A.; Nachtegaal, M.; et al. Structural selectivity of supported Pd nanoparticles for catalytic NH3 oxidation resolved using combined operando spectroscopy. Nat. Catal. 2019, 2, 157–163. [Google Scholar] [CrossRef]
  51. Chen, Y.; Yan, H.; Teng, W.; Li, J.; Liu, W.; Ren, S.; Yang, J.; Liu, Q. Comparative study on N2O formation pathways over bulk MoO3 and MoO3-x nanosheets decorated Fe2O3-containing solid waste NH3-SCR catalysts. Fuel 2023, 337, 127210. [Google Scholar] [CrossRef]
  52. Zhang, Z.; Li, R.; Wang, M.; Li, Y.; Tong, Y.; Yang, P.; Zhu, Y. Two steps synthesis of CeTiOx oxides nanotube catalyst: Enhanced activity, resistance of SO2 and H2O for low temperature NH3-SCR of NOx. Appl. Catal. B Environ. 2021, 282, 119542. [Google Scholar] [CrossRef]
  53. Hao, Z.; Jiao, Y.; Shi, Q.; Zhang, H.; Zhan, S. Improvement of NH3-SCR performance and SO2 resistance over Sn modified CeMoOx electrospun fibers at low temperature. Catal. Today 2019, 327, 37–46. [Google Scholar] [CrossRef]
  54. Wang, X.; Peng, J.; Wang, Y.; Guo, N.; Li, T.; Li, H.; Ren, D.; Gui, K. Poisoning mechanism of different Cd precursors on Fe-Ce/TiO2 catalyst for selective catalytic reduction of NOx with NH3. J. Environ. Chem. Eng. 2023, 11, 109625. [Google Scholar] [CrossRef]
  55. Jia, B.; Liu, J.; Kang, J.; Zhang, G.; Lv, D.; Wang, Y. Investigating NH3-SCR coupled with CO oxidation reaction mechanisms on titanium nanotube-loaded CuMnFe composite metal catalysts. Appl. Surf. Sci. 2024, 652, 159299. [Google Scholar] [CrossRef]
  56. Liu, Y.; Gu, T.; Weng, X.; Wang, Y.; Wu, Z.; Wang, H. DRIFT Studies on the Selectivity Promotion Mechanism of Ca-Modified Ce-Mn/TiO2 Catalysts for Low-Temperature NO Reduction with NH3. J. Phys. Chem. C 2012, 116, 16582–16592. [Google Scholar] [CrossRef]
  57. Yao, W.; Wang, X.; Liu, Y.; Wu, Z. CeOP material supported CeO2 catalysts: A novel catalyst for selective catalytic reduction of NO with NH3 at low temperature. Appl. Surf. Sci. 2019, 467–468, 439–445. [Google Scholar] [CrossRef]
  58. Liu, F.; He, H.; Zhang, C.; Shan, W.; Shi, X. Mechanism of the selective catalytic reduction of NOx with NH3 over environmental-friendly iron titanate catalyst. Catal. Today 2011, 175, 18–25. [Google Scholar] [CrossRef]
  59. Wu, Z.; Jiang, B.; Liu, Y.; Wang, H.; Jin, R. DRIFT Study of Manganese/Titania-Based Catalysts for Low-Temperature Selective Catalytic Reduction of NO with NH3. Environ. Sci. Technol. 2007, 41, 5812–5817. [Google Scholar] [CrossRef]
  60. 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]
  61. Chen, Z.; Ren, S.; Wang, M.; Yang, J.; Chen, L.; Liu, W.; Liu, Q.; Su, B. Insights into samarium doping effects on catalytic activity and SO2 tolerance of MnFeOx catalyst for low-temperature NH3-SCR reaction. Fuel 2022, 321, 124113. [Google Scholar] [CrossRef]
  62. 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] [PubMed]
  63. Ye, Z.; Liu, Y.; Nikiforov, A.; Ji, J.; Zhao, B.; Wang, J. The research on CO oxidation over Ce–Mn oxides: The preparation method effects and oxidation mechanism. Chemosphere 2023, 336, 139130. [Google Scholar] [CrossRef] [PubMed]
  64. Li, X.; Ren, S.; Chen, Z.; Chen, L.; Wang, M.; Wang, L.; Wang, A. Promoting mechanism of CuO on catalytic performance of CeFe catalyst for simultaneous removal of NOx and CO. Fuel 2023, 347, 128435. [Google Scholar] [CrossRef]
Figure 1. NOx conversion (a), N2 selectivity (b), and CO conversion (c) of catalyst. Gas composition: GHSV = 60,000·h−1, [NOx] = [NH3] = 500 ppm, [CO] = 1000 ppm, O2 = 5 vol%, N2 as balance.
Figure 1. NOx conversion (a), N2 selectivity (b), and CO conversion (c) of catalyst. Gas composition: GHSV = 60,000·h−1, [NOx] = [NH3] = 500 ppm, [CO] = 1000 ppm, O2 = 5 vol%, N2 as balance.
Catalysts 15 00674 g001
Figure 2. SEM (ac) spectra of CuSmTi catalyst and SEM (df) and TEM (gi) spectra and EDX mapping (j) of AgCuSmTi catalyst.
Figure 2. SEM (ac) spectra of CuSmTi catalyst and SEM (df) and TEM (gi) spectra and EDX mapping (j) of AgCuSmTi catalyst.
Catalysts 15 00674 g002
Figure 3. XRD (a), Raman (b), and XPS spectrum of (c) Ag 3d, (d) Cu2p, and (e) Sm 3d for catalysts.
Figure 3. XRD (a), Raman (b), and XPS spectrum of (c) Ag 3d, (d) Cu2p, and (e) Sm 3d for catalysts.
Catalysts 15 00674 g003
Figure 4. XPS (a), O2-TPD (b), and EPR (c) spectrum of catalysts.
Figure 4. XPS (a), O2-TPD (b), and EPR (c) spectrum of catalysts.
Catalysts 15 00674 g004
Figure 5. H2-TPR (a) and NH3-TPD (b) spectrum of catalysts.
Figure 5. H2-TPR (a) and NH3-TPD (b) spectrum of catalysts.
Catalysts 15 00674 g005
Figure 6. In situ DRIFTS analysis of the AgCuSmTi catalyst under different gas atmospheres: CO + O2 adsorption across a temperature range of 50 to 300 °C (a) and CO adsorption alone versus co-adsorption with O2 at 50 °C (b).
Figure 6. In situ DRIFTS analysis of the AgCuSmTi catalyst under different gas atmospheres: CO + O2 adsorption across a temperature range of 50 to 300 °C (a) and CO adsorption alone versus co-adsorption with O2 at 50 °C (b).
Catalysts 15 00674 g006
Figure 7. In situ DRIFTS sectra of AgCuSmTi catalyst under various reaction conditions: NH3 adsorption (a) and NO + O2 adsorption at temperatures ranging from 50 to 300 °C (b). Reaction between pre-adsorbed NH3 species and NO + O2 at 150 °C for 30 min (c) and reaction of NH3 with pre-adsorbed NOx species at 150 °C for 30 min (d). Gas composition: [NH3] = [NO] = 500 ppm (when used), 5 vol% O2 (when used), and N2 as balance.
Figure 7. In situ DRIFTS sectra of AgCuSmTi catalyst under various reaction conditions: NH3 adsorption (a) and NO + O2 adsorption at temperatures ranging from 50 to 300 °C (b). Reaction between pre-adsorbed NH3 species and NO + O2 at 150 °C for 30 min (c) and reaction of NH3 with pre-adsorbed NOx species at 150 °C for 30 min (d). Gas composition: [NH3] = [NO] = 500 ppm (when used), 5 vol% O2 (when used), and N2 as balance.
Catalysts 15 00674 g007
Figure 8. Schematic illustration of the NH3-SCR and CO oxidation mechanisms occurring on the AgCuSmTi catalytic system.
Figure 8. Schematic illustration of the NH3-SCR and CO oxidation mechanisms occurring on the AgCuSmTi catalytic system.
Catalysts 15 00674 g008
Table 1. Surface elemental composition of the catalyst as determined by XPS analysis.
Table 1. Surface elemental composition of the catalyst as determined by XPS analysis.
SampleSurface Atomic Concentration (%)Relative Concentration Ratio (%)
AgCuSmTiOCu2+/
(Cu+ + Cu2+)
Sm3+/
(Sm2+ + Sm3+)
Oβ/
(Oα + Oβ)
AgCuSmTi8.792.615.5024.6358.4682.51 60.98 77.77
CuSmTi-2.904.9927.6764.4468.35 61.05 64.77
Table 2. Acidic site distribution of catalysis.
Table 2. Acidic site distribution of catalysis.
Sampleαβγ
AgCuSmTi11.25%76.86%11.89%
CuSmTi11.5%68.15%20.35%
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

Li, R.; Wei, J.; Jia, B.; Liu, J.; Liu, X.; Wang, Y.; Zhao, Y.; Li, G.; Zhang, G. Ag-Cu Synergism-Driven Oxygen Structure Modulation Promotes Low-Temperature NOx and CO Abatement. Catalysts 2025, 15, 674. https://doi.org/10.3390/catal15070674

AMA Style

Li R, Wei J, Jia B, Liu J, Liu X, Wang Y, Zhao Y, Li G, Zhang G. Ag-Cu Synergism-Driven Oxygen Structure Modulation Promotes Low-Temperature NOx and CO Abatement. Catalysts. 2025; 15(7):674. https://doi.org/10.3390/catal15070674

Chicago/Turabian Style

Li, Ruoxin, Jiuhong Wei, Bin Jia, Jun Liu, Xiaoqing Liu, Ying Wang, Yuqiong Zhao, Guoqiang Li, and Guojie Zhang. 2025. "Ag-Cu Synergism-Driven Oxygen Structure Modulation Promotes Low-Temperature NOx and CO Abatement" Catalysts 15, no. 7: 674. https://doi.org/10.3390/catal15070674

APA Style

Li, R., Wei, J., Jia, B., Liu, J., Liu, X., Wang, Y., Zhao, Y., Li, G., & Zhang, G. (2025). Ag-Cu Synergism-Driven Oxygen Structure Modulation Promotes Low-Temperature NOx and CO Abatement. Catalysts, 15(7), 674. https://doi.org/10.3390/catal15070674

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