325
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Modelling of NO Reduction on CeO2-Supported Pt and Pd Nanoclusters

, ORCID Icon, & ORCID Icon
Pages 3414-3425 | Received 07 May 2023, Accepted 21 May 2023, Published online: 27 Jul 2023

ABSTRACT

The reduction mechanism of NO to N2 on CeO2-supported Pt and Pd nanoclusters (Pt/CeO2 and Pd/CeO2) is analyzed using density functional theory. Different NO decomposition and N2 formation pathways are evaluated, and the thermodynamically preferable paths are identified. The energy barrier of NO decomposition on Pt/CeO2 indicates that the rate of reaction via metallic cluster sites and the ceria oxygen vacancies are comparable. In contrast, the cluster metallic sites of the Pd/CeO2 induce a lower energy barrier of NO decomposition relative to the oxygen CeO2 vacancy sites. The N2 formation on the Pt/CeO2 preferentially occurs via the N–N association, whereas the N2O deoxidation is energetically preferred on the Pd/CeO2.

Introduction

Nitric oxides (NOx) constitute a family of air pollutants captured by moisture to form acid rain. The NOx discharge is strictly regulated in many countries because of its toxicity and ecological impact (Desantes et al. Citation2020; Duncan et al. Citation2016). Emissions of NOx from combustion are primarily in the form of NO. Globally, power generation and automobile transportation contribute to most NO emissions (Bradley and Jones Citation2002). The purpose of the industrial detoxification of NOx is to generate harmless nitrogen (N2). Nowadays, two technologies are developed for NOx emission control, namely selective catalytic reduction (SCR), and three-way catalytic conversion (TWC) (Gómez-García, Pitchon, and Kiennemann Citation2005). The SCR is widely used in power plants and heavy diesel vehicles (Bradley and Jones Citation2002), which catalyzes the NOx reduction by introducing additional reductants (ammonia or urea) (Balland, Parmentier, and Schmitt Citation2014; Praveena and Martin Citation2018). The TWC is normally used in automobiles, which catalyzes the reaction of NOx with unburnt hydrocarbons or carbon monoxide generated from vehicle engines (Koltsakis, Kandylas, and Stamatelos Citation1998). Ceria (CeO2) based catalysts are generally used in both SCR (France et al. Citation2017; Hu et al. Citation2021; Yao et al. Citation2017) and TWC (Gu et al. Citation2019; Jin, Shen, and Zhu Citation2016; Kim Citation1982; Liu et al. Citation2011; Wang et al. Citation2008), and platinum group metals (PGMs) such as Pt and Pd (Bera et al. Citation2000; Cordatos and Gorte Citation1996; Li et al. Citation2023; Marei et al. Citation2022) are doped in the CeO2 to boost the catalytic activity.

PGMs exhibit remarkable catalytic activity for NOx reduction with hydrocarbons and CO. The reactivity of NO is higher on Pt, followed by Rh and Pd (Burch and Ramli Citation1998). The NO complete conversion temperature on Pt and Rh is lower than 350°C, with an N2 yield exceeding 80% (Burch and Ramli Citation1998). The formation of N2O and NH3 as by-products is observed when PGMs catalyze the NO reduction (Adams et al. Citation2015; Auvray and Olsson Citation2015; Khosravi et al. Citation2014; Nováková Citation2001).

A complete conversion of NO to N2 in the reaction with CO is observed at 400°C with CeO2, indicating a remarkable N2-selectivity (Liu et al. Citation2011). By doping the CeO2 with other transient metals like copper (Liu et al. Citation2011), nickel (Wang et al. Citation2008), and manganese (Kijlstra et al. Citation1997; Yao et al. Citation2017), the reduction temperature of NOx can be significantly decreased to below 300°C, with no loss of selectivity toward N2. The reaction temperature of NO with CO can be reduced to 150 ~ 200°C by doping 1%wt of Pt or Pd in CeO2 (Bera et al. Citation2000). Because of these properties, catalysts with PGM/CeO2 have been developed for NOx emission control (Harrison, Diwell, and Hallett Citation1988; Thomas et al. Citation2021; Törncrona et al. Citation1997; Wang et al. Citation2017; Zhang et al. Citation2021). These catalysts usually contain CeO2 as substrate, and PGMs are doped on the CeO2 surface as clusters.

This study aims to develop fundamental knowledge using density functional theory (DFT) calculations of the mechanism by which NO reduction occurs on the CeO2-supported Pt or Pd nanoclusters. The minimum energy profiles for the decomposition of NO on metallic sites of the nanoclusters and the oxygen vacancies at the cluster-ceria boundary are obtained. The N2 formation mechanism from the N–N association and the N2O formation-deoxidation is examined. This will illustrate the origin of the high catalytic activity and N2-selectivity of PGM/CeO2 catalysts.

Methodology

Electronic energies were obtained with density functional theory (DFT) calculations, using periodic boundary conditions with the Perdew–Burke–Ernzerhof (PBE) functional as implemented in the Vienna Ab-initio Simulation Package (Kresse and Furthmüller Citation1996; Perdew, Burke, and Ernzerhof Citation1996). Bulk models of CeO2, Pt, and Pd were minimized by fully relaxing the cell volume, shape, and atom positions. The convergence criteria were 1 × 10−5 eV/atom with linear-tetrahedron smearing with Blöchl correction, a k-spacing of 0.18 Å−1, and wave functions truncated with an energy cutoff of 400 eV. The effect of 4f electron correlation was accounted for the Ce by considering on-site Coulomb and exchange interactions using the rotationally invariant formulation of the on-site model (Dudarev et al. Citation1998), with an effective correction factor (Ueff) of 5.0 eV on the Ce-f orbital. The van der Waals correction for weakly adsorbed species was simulated with the DFT-D3 zero-damping method (Grimme et al. Citation2010).

Surface p(4×4) slabs of Pt(111) and Pd(111) were created from the relaxed bulk models. All slabs are periodically repeated with a vacuum region of 15 Å between the surfaces to obviate interactions between slabs. These slabs contain four atomic layers, and the topmost two atomic layers are allowed to relax. Adsorbates were placed on the surface and optimized with a k-spacing of ~0.16 per Å along the x and y axis, together with a 0.2 eV first-order Methfessel-Paxton smearing.

The (111) termination for CeO2 is the most thermodynamically stable (Branda et al. Citation2011; Nolan and Parker Citation2005; Vyas, Grimes, and Gay Citation1998; Wanklyn and Garrard Citation1984). Therefore, this termination is used to support nanoclusters. The CeO2(111) surface contains six atomic layers, and it is terminated on both sides, with oxygen having cerium in the adjacent layer. A 10-atom nanocluster is placed on the CeO2 surface to simulate the supported Pt or Pd cluster on the CeO2. The cluster contains two atomic layers, with seven atoms aligned as a hexagon in the bottom and three atoms aligned as a triangle in the top layer. The surface slabs were relaxed with a k-spacing of ~0.48 per Å in the x and y-axis, together with a 0.05 eV Gaussian smearing, and the spin was allowed to relax in all CeO2 slabs. Three atomic layers in the bottom of the slab are fixed during the relaxation.

The energy of gas-phase species (Egas) was obtained in a cubic cell of side length 15 Å with spin polarization corrections, using a Γ-point spacing and the Gaussian smearing with a width of 0.001 eV.

The binding energy (Eb) of adsorbates was calculated using Eq. 1, where the Eslab and Esys are, namely, the electronic energy of the bare surface slab and adsorbing system.

(1) Eb=Egas+EslabEsys(1)

Transition states were identified by the climbing image – nudge elastic band (CI-NEB) method (Henkelman, Uberuaga, and Jonsson Citation2000). The minimum energy path between the minimized initial and final structures was obtained using three or six intermediate images with a spring constant of 5 eV/Å. The image closest to the saddle point was allowed to climb up into the saddle point if the largest force on an atom was smaller than 1.0 eV/Å. The climbing process was terminated when the largest force was reduced to less than 0.05 eV/Å.

Results and discussion

The reduction of NO(g) to N2(g) on the metal clusters was analyzed by determining the reaction and the activation energies of the elementary steps as indicated in reactions R1 to R5. First, the binding characteristics of all species were evaluated on multiple binding sites, and the most stable configurations were selected. Second, the surface dissociation of NO to atomic nitrogen (N*) and oxygen (O*) and the recombination of N* to N2(g) were evaluated (R1 ~ R3). Other possible pathways of NO* to form N2O* and the subsequent formation of N2(g) were also analyzed (R4 and R5).

NO Adsorption NOg+NO(R1)

NO Decomposition NO+N+O(R2)

N–N Association N+NN2g(R3)

N2O Formation NO+NN2O(R4)

N2O Deoxidation N2ON2g+O(R5)

Binding characteristics of nitrogen species

shows the binding energies (Eb) of N*, NO*, N2*, and N2O* on the most stable sites of Pt/CeO2, Pd/CeO2, Pt(111), and Pd(111), and their corresponding molecular configurations. As can be seen, adsorbed product gas-phase species have lower binding energies (N2*, and N2O*) than intermediate species (NO* and N*).

Figure 1. (a) binding energies and (b) molecular configurations of N*, NO*, N2*, and N2O* on the Pt/CeO2, Pd/CeO2, Pt(111), and Pd(111) surfaces. Numbers in (b) are the bond lengths (in Å) of N-O or N-N.

Figure 1. (a) binding energies and (b) molecular configurations of N*, NO*, N2*, and N2O* on the Pt/CeO2, Pd/CeO2, Pt(111), and Pd(111) surfaces. Numbers in (b) are the bond lengths (in Å) of N-O or N-N.

The N* has the largest Eb of all species, which is reasonable given its free atomic radical nature. As seen in , there is a considerable structural deformation of the clusters upon the N binding on the Pd/CeO2 cluster, leading to lower EbN compared to the other surfaces. The NO binds N-down on the flat (111) surfaces and the clusters, with only slight deformation of the atomic cluster positions. The EbNO on the clusters are stronger than the flat surfaces. Both N2* and N2O* are only weakly bound to the flat surfaces, whereas the Eb of these two species on the clusters is marginally stronger.

NO decomposition

The NO decomposition proceeds through two pathways: On metal surfaces, the O species in the adsorbed NO* molecule is extracted by nearby metallic sites (NO* + * ⇌ N* + O*) (Gonzalez et al. Citation2018; Mavrikakis et al. Citation1999). On the CeO2 surface, oxygen vacancies ([Ov]) can extract the O from the NO molecule to fill the lattice (NO* + [Ov] ⇌ N* + [O]) (Wang et al. Citation2017). The CeO2-supported nanoclusters can provide both metallic sites and surface oxygen vacancies, so both pathways are evaluated and compared. show the energy profiles and transition state structures of these two pathways.

Figure 2. Energy profiles and transition state structures of NO decomposition with metallic sites on supported nanoclusters and flat metal surfaces. Orange numbers in the structure diagrams are the distance between the N and O atoms (dNO). Energies are relative to NO(g) and the clean surface.

Figure 2. Energy profiles and transition state structures of NO decomposition with metallic sites on supported nanoclusters and flat metal surfaces. Orange numbers in the structure diagrams are the distance between the N and O atoms (dN−O). Energies are relative to NO(g) and the clean surface.

Figure 3. Energy profiles and transition state structures of NO decomposition with metallic sites on supported nanoclusters and flat metal surfaces. Orange numbers in the structure diagrams are the distance between the N and O atoms (dNO). Energies are relative to NO(g) and the clean surface with an oxygen vacancy.

Figure 3. Energy profiles and transition state structures of NO decomposition with metallic sites on supported nanoclusters and flat metal surfaces. Orange numbers in the structure diagrams are the distance between the N and O atoms (dN−O). Energies are relative to NO(g) and the clean surface with an oxygen vacancy.

In , the step from the first to second stage refers to the binding energy of NO*; the energy difference between the third and second stages is the activation energy (Ea); the difference from second to fourth stage is the reaction energy referred to the co-adsorption state (ΔEr); and the energy difference between the fifth stage and the fourth stage is the co-adsorption energy of N* and O* on the same cluster (Ecoad).

As shown in , the NO decomposition is endothermic on Pt/CeO2 and Pd/CeO2. Either Ea or ΔE on the Pt/CeO2 is slightly lower than on the Pd/CeO2. As indicated in the figure, the dNO at the transition state on the Pt/CeO2 is marginally shorter than on the Pd/CeO2. The N atom at the transition state on the Pd/CeO2 causes a slight deformation to the cluster, which leads to the relatively low stability of this transition state. Also, the negative Ecoad at the final state indicates a stable co-adsorption structure of N* and O* on the clusters, which can benefit the NO decomposition.

The NO* decomposition on the Pt(111) and Pd(111) are evaluated, resulting in the Ea of 203 kJ·mol−1 on Pt and 226 kJ·mol−1 on Pd, which are consistent with literature data (Gonzalez et al. Citation2018; Mavrikakis et al. Citation1999). The reaction is slightly endothermic on both (111) surfaces, with a ΔE of 10 kJ·mol−1 on Pt and 59 kJ·mol−1 on Pd. Compared to metallic surfaces, the CeO2-supported clusters show remarkably low Ea for the NO decomposition.

The reaction of NO with the [Ov] is evaluated at the cluster boundary as shown in . A boundary lattice oxygen [O] is pre-removed at the initial state of the reaction, and the slab containing NO* and boundary [Ov] sites is re-optimized. The [Ov] only slightly reduces the binding energy of NO* compared to the non-defected surface. During the reaction, the O species moves in to the [Ov] and finally becomes lattice oxygen on the CeO2 surface, leaving N* at the cluster. The Ea of this reaction on the Pt/CeO2 is significantly lower than on the Pd/CeO2, and the reaction is more exothermic on the Pt/CeO2 as well. As can be seen in , the Pd cluster shows a considerable deformation at the transition state, while the Pt cluster is stable during the reaction.

In comparison, the NO decomposition on the Pt/CeO2 has similar Ea via the NO* + * and NO* + [Ov] pathways, while on the Pd/CeO2, the NO* + * pathway shows a significantly lower barrier. Both catalysts have better activity than the pure metal surfaces, and the Pt/CeO2 tends to be more active.

N2 formation

Two pathways of N2 formation are investigated, namely the association of two N* (R3), and the deoxidation of N2O (R4 and R5).

As indicated in , during the N–N association, two N atoms are first adsorbed on a supported cluster to form N2* molecule. The co-adsorption of 2 N* is exothermic on both Pt and Pd clusters. Therefore, the adsorption of multiple N atoms on one cluster is energetically preferable. The Ea of the N–N association is lower on Pt/CeO2 than on Pd/CeO2. The transition state structures indicate that the dNN on the Pt/CeO2 is shorter, which improves its stability. The reaction on both clusters is exothermic, where the Pd/CeO2 has a larger ΔE.

Figure 4. Energy profiles and transition state structures of N-N association on supported nanoclusters and flat metal surfaces. Green numbers in the structure diagrams are the distance between the N and N atoms (dNN). Energies are relative to N2(g) and the clean surface.

Figure 4. Energy profiles and transition state structures of N-N association on supported nanoclusters and flat metal surfaces. Green numbers in the structure diagrams are the distance between the N and N atoms (dN−N). Energies are relative to N2(g) and the clean surface.

Compared to the flat (111) surfaces, the supported clusters provide better catalytic activity for the N–N association. The Ecoad of 2 N* is 85 kJ·mol−1 on Pt(111) and 46 kJ·mol−1 on Pd(111). The Ea for the N–N association of 2 N* is 176 kJ·mol−1 on Pt(111) and 179 kJ·mol−1 on Pd(111), which are consistent with literature data (Gonzalez et al. Citation2018; Mavrikakis et al. Citation1999). Therefore, on the flat (111) surfaces, the N–N association requires higher Ea than the supported clusters.

The N2 can also be produced through the N2O formation-deoxidation pathway (R4 and R5). As compared in , the co-adsorption of N* + NO* on the Pt/CeO2 is marginally endothermic, while on the Pd/CeO2 is highly exothermic. The Ea from the co-adsorbed state on the Pt/CeO2 is lower than on the Pd/CeO2. The transition state structures in shows that the N2O formation only involves the interaction between the two N atoms, while the N-O bond is not affected. The N–N bond length (dNN) at the transition state on the Pt/CeO2 is close to the N2O(g) molecule (1.40 Å), resulting in a similar energy. On the contrary, the dNN at the transition state on the Pd/CeO2 is relatively longer, so the transition state is less stable and the Ea is also higher on the Pd cluster.

Figure 5. Energy profiles and transition state structures of N2O formation-deoxidation on supported nanoclusters and flat metal surfaces. Green numbers in the structure diagrams are the distance between N and N atoms (dNN), and oranges are between N and O atoms (dNO). Energies are relative to N2O(g) and the clean surfaces.

Figure 5. Energy profiles and transition state structures of N2O formation-deoxidation on supported nanoclusters and flat metal surfaces. Green numbers in the structure diagrams are the distance between N and N atoms (dN−N), and oranges are between N and O atoms (dN−O). Energies are relative to N2O(g) and the clean surfaces.

The Ea of the N2O deoxidation on the Pt/CeO2 is lower than that of Pd/CeO2. At the transition state shown in , the dNO on the Pt/CeO2 is longer than on the Pd/CeO2, while the dNN is shorter, resulting in a more stable N2* molecule on the Pt. The energy barriers of the N2O deoxidation are generally lower than the N2O formation in these two clusters, so the N2O formation is a rate-limiting step for N2 formation through this path. Furthermore, on both clusters, the Ea are lower than the EbN2O, which leads to the formation of N2 energetically preferable than the N2O desorption.

The co-adsorption of N* and NO* on Pt(111) and Pd(111) surfaces is also endothermic, with Ecoad of 43 kJ·mol−1 and 38 kJ·mol−1 on Pt and Pd, respectively. The Ea for N2O formation on the Pt(111) is 101 kJ·mol−1, which is considerably lower than the Pd(111) (198 kJ·mol−1). For the N2O deoxidation, the Ea is 137 kJ·mol−1 and 68 kJ·mol−1 on Pt(111) and Pd(111), respectively, and both are higher than the EbN2O on these surfaces. As a result, the N2O desorption tends to be more thermodynamically preferable on either Pt or Pd, which leads to the formation of N2O(g) as observed in previous experimental works (Novakova and Kubelkova Citation1997; Tanikawa and Egawa Citation2011). Therefore, the CeO2-supported nanocluster can provide a higher N2-selectivity and avoid the generation of unwanted product N2O.

Based on the differences in energy barriers and reaction energies, the N2 formation on the Pt/CeO2 prefers the N–N association pathway, whereas the N2O pathway is energetically preferable on the Pd/CeO2. Both supported clusters can improve the selectivity N2 during the NO reduction because of the stable adsorption of the by-product N2O.

Conclusion

The catalytic mechanism of NO reduction to N2 on the CeO2-supported Pt and Pd nanoclusters is investigated using DFT. The elementary pathways including NO decomposition, N–N association, N2O formation, and N2O deoxidation are assessed.

The supported clusters show remarkable activity for NO decomposition compared to the pure metal (111) surfaces. The activation energy is 90 kJ·mol−1 on Pt/CeO2 and 131 kJ·mol−1 on Pd/CeO2. Also, the NO decomposition with the oxygen vacancy at the cluster-ceria boundary is evaluated, which results in a barrier of 102 kJ·mol−1 on the Pt/CeO2 and 192 kJ·mol−1 on the Pd/CeO2. Therefore, the metal sites on both clusters are active for NO decomposition, but only the vacancy at Pt/CeO2 boundary shows considerable activity.

Two pathways of N2 formation are exanimated, namely the direct association of atomic nitrogen and the by-path through the formation and deoxidation of N2O. The N–N association barrier is 56 kJ·mol−1 on the Pt/CeO2, and 186 kJ·mol−1 on the Pd/CeO2. Both Ea are lower than the pure metal (111) surfaces.

The Pd/CeO2 shows better catalytic activity for N2 formation through the N2O pathway because of its lower barrier for the N2O formation compared to the Pt/CeO2. The barriers for the N2O deoxidation are generally lower than the N2O formation, so it would not limit the N2 formation rate through the N2O pathway. On the other hand, the relatively strong binding energy of N2O on the supported Pt and Pd clusters can avoid the desorption of N2O, which is one of the major advantages of such catalysts compared to the pure Pt and Pd.

Acknowledgements

The research reported in this work was funded by the Office of Sponsored Research at King Abdullah University of Science and Technology (KAUST) under the Competitive Research Grant OSR-CRG2018-3042.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

The work was supported by the King Abdullah University of Science and Technology [OSR-CRG2018-3042].

References

  • Adams, E. C., M. Skoglundh, M. Folic, E. C. Bendixen, P. Gabrielsson, and P.-A. Carlsson. 2015. Ammonia formation over supported platinum and palladium catalysts. Appl Catal B 165:10–19. doi:10.1016/j.apcatb.2014.09.064.
  • Auvray, X., and L. Olsson. 2015. Stability and activity of Pd-, Pt- and Pd–Pt catalysts supported on alumina for NO oxidation. Appl Catal B 168-169:342–52. doi:10.1016/j.apcatb.2014.12.035.
  • Balland, J., M. Parmentier, and J. Schmitt. 2014. Control of a combined SCR on filter and under-floor SCR system for low emission passenger cars. SAE Int. J. Engines 7 (3):1252–61. doi:10.4271/2014-01-1522.
  • Bera, P., K. Patil, V. Jayaram, G. Subbanna, and M. Hegde. 2000. Ionic dispersion of Pt and Pd on CeO2 by combustion method: Effect of metal–ceria interaction on catalytic activities for NO reduction and CO and hydrocarbon oxidation. J Catal 196:293–301. doi:10.1006/jcat.2000.3048.
  • Bradley, M. J., and B. M. Jones. 2002. Reducing global NOx emissions: Developing advanced energy and transportation technologies. Ambio: A J. Human Environ. 31:141–49. doi:10.1579/0044-7447-31.2.141.
  • Branda, M. M., R. M. Ferullo, M. Causa, and F. Illas. 2011. Relative stabilities of low index and stepped CeO2 surfaces from hybrid and GGA + U implementations of density functional theory. J. Phys. Chem. C 115 (9):3716–21. doi:10.1021/jp111427j.
  • Burch, R., and A. Ramli. 1998. A comparative investigation of the reduction of NO by CH4 on Pt, Pd, and Rh catalysts. Appl. Catal. B 15:49–62. doi:10.1016/S0926-3373(97)00036-2.
  • Cordatos, H., and R. Gorte. 1996. CO, NO, and H2 adsorption on ceria-supported Pd. J. Catal. 159:112–18. doi:10.1006/jcat.1996.0070.
  • Desantes, J., S. Molina, R. Novella, and M. Lopez-Juarez. 2020. Comparative global warming impact and NOX emissions of conventional and hydrogen automotive propulsion systems. Energy Convers. Manage. 221:113137. doi:10.1016/j.enconman.2020.113137.
  • Dudarev, S. L., G. A. Botton, S. Y. Savrasov, C. J. Humphreys, and A. P. Sutton. 1998. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study. Phys. Rev. B 57:1505. doi:10.1103/PhysRevB.57.1505.
  • Duncan, B. N., L. N. Lamsal, A. M. Thompson, Y. Yoshida, Z. Lu, D. G. Streets, M. M. Hurwitz, and K. E. Pickering. 2016. A space-based, high-resolution view of notable changes in urban NO x pollution around the world (2005-2014). J. Geophys. Res.: Atmos 121 (2):976–96. doi:10.1002/2015JD024121.
  • France, L. J., Q. Yang, W. Li, Z. Chen, J. Guang, D. Guo, L. Wang, and X. Li. 2017. Ceria modified FeMnOx—enhanced performance and sulphur resistance for low-temperature SCR of NOx. Appl Catal B 206:203–15. doi:10.1016/j.apcatb.2017.01.019.
  • Gómez-García, M., V. Pitchon, and A. Kiennemann. 2005. Pollution by nitrogen oxides: An approach to NOx abatement by using sorbing catalytic materials. Environ. Int. 31:445–67. doi:10.1016/j.envint.2004.09.006.
  • Gonzalez, J. D., K. Shojaee, B. S. Haynes, and A. Montoya. 2018. The effect of surface coverage on N2, NO and N2O formation over Pt (111). Phys. Chem. 20 (39):25314–23. doi:10.1039/C8CP04066D.
  • Grimme, S., J. Antony, S. Ehrlich, and H. Krieg. 2010. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132 (15):154104. doi:10.1063/1.3382344.
  • Gu, Q., L. Wang, Y. Wang, and X. Li. 2019. Effect of praseodymium substitution on La1-xPrxmno3 (x= 0–0.4) perovskites and catalytic activity for NO oxidation. J. Phys. Chem. Solids 133:52–58. doi:10.1016/j.jpcs.2019.05.001.
  • Harrison, B., A. Diwell, and C. Hallett. 1988. Promoting platinum metals by ceria. Platin Met. Rev. 32:73–83.
  • Henkelman, G., B. P. Uberuaga, and H. Jonsson. 2000. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 113:9901–04. doi:10.1063/1.1329672.
  • Hu, X., J. Chen, W. Qu, R. Liu, D. Xu, Z. Ma, and X. Tang. 2021. Sulfur-resistant ceria-based low-temperature SCR catalysts with the non-bulk electronic states of ceria. Environ. Sci. Technol. 55 (8):5435–41. doi:10.1021/acs.est.0c08736.
  • Jin, Q., Y. Shen, and S. Zhu. 2016. Praseodymium oxide modified CeO2/Al2O3 catalyst for selective catalytic reduction of NO by NH3. Chin. J. Chem. 34 (12):1283–90. doi:10.1002/cjoc.201600565.
  • Khosravi, M., C. Sola, A. Abedi, R. Hayes, W. Epling, and M. WOtsov. 2014. Oxidation and selective catalytic reduction of NO by propene over Pt and Pt: Pd diesel oxidation catalysts. Appl. Catal. B 147:264–74. doi:10.1016/j.apcatb.2013.08.034.
  • Kijlstra, W. S., D. S. Brands, H. I. Smit, E. K. Poels, and A. Bliek. 1997. Mechanism of the selective catalytic reduction of NO with NH3over MnOx/Al2O3. J. Catal. 171:219–30. doi:10.1006/jcat.1997.1789.
  • Kim, G. 1982. Ceria-promoted three-way catalysts for auto exhaust emission control. Ind. Eng. Chem. Prod. Res. Dev. 21 (2):267–74. doi:10.1021/i300006a014.
  • Koltsakis, G., I. Kandylas, and A. Stamatelos. 1998. Three-way catalytic converter modeling and applications. Chem. Eng. Commun. 164:153–89. doi:10.1080/00986449808912363.
  • Kresse, G., and J. Furthmüller. 1996. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54 (16):11169. doi:10.1103/PhysRevB.54.11169.
  • Li, Z., M. N. Marei, A. Farooq, A. R. Masri, and A. Montoya. 2023. Micro-kinetics of ethylene and methane oxidation on platinum. The Chem. Engg. J. 464:142608. doi:10.1016/j.cej.2023.142608.
  • Liu, L., Y. Cao, W. Sun, Z. Yao, B. Liu, F. Gao, and L. Dong. 2011. Morphology and nanosize effects of ceria from different precursors on the activity for NO reduction. Catal. Today 175:48–54. doi:10.1016/j.cattod.2011.04.018.
  • Liu, L., Z. Yao, Y. Deng, F. Gao, B. Liu, and L. Dong. 2011. Morphology and crystal‐plane effects of nanoscale ceria on the activity of CuO/CeO2 for NO reduction by CO. Chem. Cat. 3:978–89. doi:10.1002/cctc.201000320.
  • Marei, M. N., H. A. Khan, J. A. Badra, A. Montoya, A. A. M. Farooq, and A.R. Masri. 2022. Comparative study of the catalytic oxidation of hydrocarbons on platinum and palladium wires and nanoparticles. Energy & Fuels 36 (4):2044–57. doi:10.1021/acs.energyfuels.1c04136.
  • Mavrikakis, M., L. B. Hansen, J. J. Mortensen, B. Hammer, and J. K. Nørskov. 1999. Dissociation of N 2 NO, and CO on transition metal surfaces. In Transition state modeling for catalysis, 245–58. American Chemical Society. doi:10.1021/bk-1999-0721.ch019.
  • Nolan, M., S. C. A. W. Parker. 2005. The electronic structure of oxygen vacancy defects at the low index surfaces of ceria. Surf. Sci. 595 (1–3):223–32. doi:10.1016/j.susc.2005.08.015.
  • Nováková, J. 2001. Reduction of NO by hydrogen versus reduction by CO over Pt, Pd and Rh clusters in NaX zeolite. Appl. Catal. B 30 (3–4):445–57. doi:10.1016/S0926-3373(00)00260-5.
  • Novakova, J., and L. Kubelkova. 1997. Contribution to the mechanism of NO reduction by CO over Pt/NaX zeolite. Appl. Catal. B 14:273–86. doi:10.1016/S0926-3373(97)00029-5.
  • Perdew, J. P., K. Burke, and M. Ernzerhof. 1996. Generalized gradient approximation made simple. Phys. Rev. Lett. 77 (18):3865. doi:10.1103/PhysRevLett.77.3865.
  • Praveena, V., and M. L. J. Martin. 2018. A review on various after treatment techniques to reduce NOx emissions in a CI engine. J. Energy Inst. 91 (5):704–20. doi:10.1016/j.joei.2017.05.010.
  • Tanikawa, K., and C. Egwa. 2011. Effect of barium addition over palladium catalyst for CO–NO–O2 reaction. J. Mol. Catal. A Chem. 349 (1–2):94–99. doi:10.1016/j.molcata.2011.08.025.
  • Thomas, C. R., J. A. Pihl, V. Y. Prikhodko, M. K. Kidder, J. A. Lauterbach, and T. J. Joops. 2021. The effects of ceria loading on three-way catalysts for passive SCR operation. Catal. Commun. 156:106308. doi:10.1016/j.catcom.2021.106308.
  • Törncrona, A., M. Skoglundh, P. Thormählen, E. Fridell, and E. Jobson. 1997. Low temperature catalytic activity of cobalt oxide and ceria promoted Pt and Pd: -influence of pretreatment and gas composition. Appl. Catal. B 14:131–45. doi:10.1016/S0926-3373(97)00018-0.
  • Vyas, S., R. W. Grimes, D. H. A. R. Gay. 1998. Structure, stability and morphology of stoichiometric ceria crystallites. J. Chem. Soc. Faraday Trans. 94 (3):427–34. doi:10.1039/a707052g.
  • Wang, Y., R. Oord, D. Van Den Berg, B. M. Weckhuysen, and M. Makee. 2017. Oxygen vacancies in reduced Rh/ and Pt/ceria for highly selective and reactive reduction of NO into N 2 in excess of O 2. Chem. Cat. 9 (15):2935–38. doi:10.1002/cctc.201700578.
  • Wang, Y., A. Zhu, Y. Zhang, C. Au, X. Yang, and C. Shi. 2008. Catalytic reduction of NO by CO over NiO/CeO2 catalyst in stoichiometric NO/CO and NO/CO/O2 reaction. Appl. Catal. B 81:141–49. doi:10.1016/j.apcatb.2007.12.005.
  • Wanklyn, B., and B. Garrard. 1984. The flux growth of large thoria and ceria crystals. J. Cryst. Growth 66:346–50. doi:10.1016/0022-0248(84)90218-5.
  • Yao, X., T. Kong, L. Chen, S. Ding, F. Yang, and L. Dong. 2017. Enhanced low-temperature NH3-SCR performance of MnOx/CeO2 catalysts by optimal solvent effect. Appl. Surf. Sci. 420:407–15. doi:10.1016/j.apsusc.2017.05.156.
  • Zhang, L., G. Spezzati, V. Muravev, M. A. Verheijen, B. Zijlstra, I. A. W. Filot, Y. Q. Su, M. W. Chang, and E. J. M. Hensen. 2021. Improved Pd/CeO2 Catalysts for low-temperature NO reduction: Activation of CeO2 lattice oxygen by Fe doping. ACS Catal. 11 (9):5614–27. doi:10.1021/acscatal.1c00564.