Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2023 Feb 14;120(8):e2216584120. doi: 10.1073/pnas.2216584120

In situ formation of cocatalytic sites boosts single-atom catalysts for nitrogen oxide reduction

Pengfei Wang a,b,1, Guoquan Liu a,1, Zhifei Hao a,1, He Zhang a, Yi Li c, Wenming Sun d, Lirong Zheng e, Sihui Zhan a,2
PMCID: PMC9974487  PMID: 36787366

Significance

The use of oxide-supported isolated Pt single atoms as catalytic active sites is of interest due to their maximum atom utilization efficiency and unique catalytic properties for the removal of NOx. However, relationships between the structure of active center, the dynamic response to environments, and catalytic functionality have proved difficult to experimentally establish. This research reveals that the in situ formation of adjacent cocatalytic sites boosts single-atom catalysts for the conversion of NOx during the catalytic reaction of NOx reduction with H2. The presence of cocatalytic sites tunes and optimizes the catalytic performance at the atomic scale that can benefit the environment and human health by removing NOx efficiently.

Keywords: single-atom catalysts, cocatalytic sites, NOx reduction, H2 activation, air pollution

Abstract

Nitrogen oxide (NOx) pollution presents a severe threat to the environment and human health. Catalytic reduction of NOx with H2 using single-atom catalysts poses considerable potential in the remediation of air pollution; however, the unfavorable process of H2 dissociation limits its practical application. Herein, we report that the in situ formation of PtTi cocatalytic sites (which are stabilized by Pt–Ti bonds) over Pt1/TiO2 significantly increases NOx conversion by reducing the energy barrier of H2 activation. We demonstrate that two H atoms of H2 molecule are absorbed by adjacent Pt atoms in Pt–O and Pt–Ti, respectively, which can promote the cleave of H–H bonds. Besides, PtTi sites facilitate the adsorption of NO molecules and further lower the activation barrier of the whole de-NOx reaction. Extending the concept to Pt1/Nb2O5 and Pd1/TiO2 systems also sees enhanced catalytic activities, demonstrating that engineering the cocatalytic sites can be a general strategy for the design of high-efficiency catalysts that can benefit environmental sustainability.


Nitrogen oxide (NOx) from stationary sources and vehicle exhaust is the primary source of acid rain, greenhouse effect, photochemical smog, and fine particulate matter (13), which has a substantial impact on the global environment and human health. A promising modern technology for the removal of NOx is selective catalytic reduction with typical reductants (H2, NH3, and HC), with the use of H2 as the reducing agent, in particular, being able to attract considerable attention due to its many distinct advantages, including sustainability, cleanliness, and superior lower temperature activity (4). It is assumed in previous studies of NOx reduction by H2 that the reaction mechanism begins with the activation of H2, following that the generated active hydrogen species cleave the N–O bond in NOx (5, 6). Therefore, the efficient activation of H2 is the critical step in the removal of NOx (7); however, it is still a challenge.

Recently, Pt-based catalysts, especially Pt single-atom catalysts (Pt1/Al2O3, Pt1/WO3, and Pt1/NC, etc.) with the maximum atom utilization efficiency and unique catalytic properties (811), show remarkable performance due to the superior activity of H2 activation in catalytic hydrogenation reaction and reverse water–gas shift reaction. Despite the increasing consensus that isolated Pt sites are the dominant active sites (12), the nature of single Pt active sites is still under debate, and it still needs to optimize their ability of H2 activation for real applications. Generally, Pt single atoms possess various chemical coordination environments (Pt–Ox) on oxide supports; in this case, some unstable Pt–O bonds will be interrupted in the high-temperature reducing atmosphere, accompanied by the in situ formation with the new coordination (monatomic metal sites and metal nanoparticle sites, etc.) (11, 13). Moreover, it has been reported that the catalytic performance will be promoted by constructing the cocatalytic sites next to single-atom sites in oxygen reduction reaction (14, 15). The reason is that the adjacent cocatalytic sites can regulate the physical/chemical property of isolated active sites to boost the activation of gas molecule (1416). For example, adjacent Fe nanoparticles of cocatalytic sites can significantly optimize the electronic structure of isolated Fe–N–C sites, thus assisting the O2 activation on the Fe–N–C sites and improving the performance of single-atom catalysts (17). These findings suggest that the in situ formation of Pt species with new coordination evolved from unstable Pt–O coordination will facilitate the activation of H2 molecules in the reduction of NO with H2; however, it is rarely discussed.

In this work, using the Pt1/TiO2 catalyzed reduction of NO with H2 as an example, we propose that adjacent PtTi cocatalytic sites (which are stabilized by Pt–Ti bonds) can significantly boost single-atom catalysts for H2 activation, as well as reduce the energy barrier for N–O activation to accelerate the catalytic reaction of NO reduction. The formation of PtTi sites in Pt1/TiO2 is directly identified by in situ high-angle annular dark-field-scanning transmission electron microscopy (HAADF-STEM) and X-ray absorption fine structure (XAFS) under the real reaction condition. Meanwhile, theoretical studies reveal the surface reaction mechanism of NO reduction boosted by PtTi sites, which promote the efficient activation of NO by active H species. Furthermore, the superior de-NOx performance (ratios differ by 20%) through the formation of cocatalytic sites is also confirmed in Pt1/Nb2O5 and Pd1/TiO2 catalysts, which provides the possibility to promote environmental sustainability and protect human health.

Results

Characterization of Pt1/TiO2 Single-Atom Catalyst.

Pt single atom (0.22 wt%) supported on TiO2 nanosheets (denoted as fresh Pt1/TiO2) was prepared by immersion method as the model catalyst for the reduction of NO by H2 (SI Appendix, Table S1) (18, 19). The fresh Pt1/TiO2 was first characterized by X-ray diffraction and high-resolution transmission electron microscopy (SI Appendix, Figs. S1–S3). The results showed that the characteristic peaks mainly indexed to anatase TiO2 phase, and no discernible peak of platinum was detected. The HAADF-STEM clearly showed the dispersion of individual Pt atoms (marked by the yellow circles) on TiO2, and energy-dispersive X-ray spectroscopy (EDS) mapping images revealed that the Pt species were uniformly dispersed on the surface of TiO2 nanosheet (Fig. 1A and SI Appendix, Fig. S4). Notably, there were some adjacent Pt atoms in fresh Pt1/TiO2 (marked by the orange circles), which might induce the dynamic migration of Pt atoms (11, 13). As expected, the position of Pt atoms was found to change significantly after reaction (denoted as used Pt1/TiO2), but Pt atoms (marked by the orange circles) were confirmed by HAADF-STEM to distribute in the form of single atoms (Fig. 1 E and IL and SI Appendix, Fig. S5). All Pt atoms in fresh Pt1/TiO2 were anchored in bright lattice fringe (Ti atomic columns) of TiO2 (101), while those in used Pt1/TiO2 were trapped in both bright lattice fringes and dark lattice fringes (O atomic columns). Additionally, line-scan intensity analysis (Fig. 1 M and N) also demonstrated the position evolution of Pt atoms, i.e., the Pt atoms originally anchored only on Ti atomic column and evolved into atomic columns of both O and Ti after reaction (taken along the lines 1 and 2 in Fig. 1 A and E, respectively).

Fig. 1.

Fig. 1.

HAADF-STEM characterization of Pt1/TiO2. (A and E) HAADF-STEM images of fresh Pt1/TiO2 (A) and used Pt1/TiO2 (E). (B–D) In situ HAADF-STEM images and relevant intensity profiles of Pt1/TiO2 without reaction in Ar atmosphere at 125 °C (1 bar). (F–H) In situ HAADF-STEM images and relevant intensity profiles of Pt1/TiO2 under the reaction at 125 °C (NO:H2 ~ 1:1, 1 bar). (I–L) EDS mapping of used Pt1/TiO2. (M and N) The intensity profiles obtained in A and E, respectively.

To further unravel the dynamic structural evolution of Pt single atom under real reaction condition, in situ HAADF-STEM was used to monitor the reaction process, in which small probe currents and minimal acquisition times were used in observation to minimize the electron beam effects (20). For fresh Pt1/TiO2, the positions of Pta and Ptb atoms (marked by white circle) remained unchanged in 600 s (125 °C, under Ar atmosphere), which was also confirmed by the results of line-scan intensity analysis that the distance of Ptb atom and adjacent Ti atomic column was changeless (~3.5 Å, displayed in Fig. 1 B–DInset), free from the interference of temperature. Then, the gas mixture of NO and H2 (volume ratio NO:H2 ~ 1:1, 1 bar) was introduced, and PtA and PtB atoms were found to be trapped in bright lattice fringe of TiO2 (Fig. 1F), corresponding to the initial state of Pt single atoms in fresh catalyst. After 60 s of exposure to the gas mixture, we found that the PtB atom moved close to PtA atom and was anchored in dark lattice fringe (Fig. 1G), accompanied by the distance of PtB atom and the adjacent Ti atom column decreased from 3.6 Å to 2.1 Å (Fig. 1 F and G, Inset). With increasing exposure time, the PtB atom was finally anchored at the position of O atom column after 600 s gas exposure as shown in Fig. 1H, during which the position of PtA atom remained stable. The observation was consistent with the result of ex situ HAADF-STEM (Fig. 1E), which visually confirmed the phenomenon of dynamic structural evolution.

Chemical Nature of Evolution over Pt Single Atoms.

To identify the chemical nature of the evolution over Pt single atoms, we employed XAFS, density functional theory (DFT) calculations, in situ Raman spectroscopy, in situ CO adsorption diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS), and quasi in situ X-ray photoelectron spectroscopy (XPS) measurements. According to the results of X-ray absorption near-edge structure (XANES) curves of Pt L3-edge (Fig. 2A), the white-line intensities of both fresh Pt1/TiO2 and used Pt1/TiO2 were between Pt foil and PtO2, indicating that the valences of Pt atoms were 2.6 (fresh Pt1/TiO2) and 1.3 (used Pt1/TiO2) (Fig. 2 A, Inset), which proved that Pt single atoms were in the lower valence states after reaction. Moreover, for fresh Pt1/TiO2, the extended X-ray absorption fine structure (EXAFS) displayed a major peak around 1.5 Å, attributed to Pt–O coordination (Fig. 2B), suggesting that Pt in fresh Pt1/TiO2 catalyst was largely present as isolated single atoms on TiO2. However, besides the dominant peak of Pt–O around 1.5 Å, a new peak appeared at 2.2 Å on used Pt1/TiO2 catalyst (Fig. 2B). Notably, the wavelet transform (WT) EXAFS (SI Appendix, Fig. S6) showed that the intensity between nanophase resolved 2 and 3 Å was localized in a different k range from Pt–Pt coordination (5 to 7 Å−1 compared to 8 to 12 Å−1), implying that the EXAFS peak around 2.2 Å was attributed to Pt–Ti coordination (11, 21). Combined with the observation of HAADF-STEM, it also confirmed that more than one Pt species were present in used Pt1/TiO2, which further revealed the structural evolution of Pt atoms in fresh Pt1/TiO2 during the reaction.

Fig. 2.

Fig. 2.

Chemical nature of evolution over Pt single atoms. (A) Pt L3-edge XANES spectra of Pt foil, PtO2, fresh Pt1/TiO2, and used Pt1/TiO2. (B) FT-EXAFS spectra of Pt foil, PtO2, fresh Pt1/TiO2, and used Pt1/TiO2 at the Pt K-edge. (C) Top views of TiO2 (101) surface model adopted in this work. And possible Pt–O coordination was labeled. (D and E) The optimized structure models of fresh Pt1/TiO2 (D) and used Pt1/TiO2 (E). (F) The formation mechanism of Pt–Ti bonds calculated by DFT. (G) EXAFS R space-fitting curve and the original curve over fresh Pt1/TiO2. (H) EXAFS R space-fitting curve and the original curve over used Pt1/TiO2. (I) In situ Raman spectra of Pt1/TiO2 under the reaction at 125 °C (volume ratio NO:H2 ~ 1:1).

To further validate the existence form of Pt single atoms, we first screened various Pt–O structures stabilized on the (101) crystal plane of TiO2 over fresh Pt1/TiO2. Notably, TiO2 (101) surface only had five-coordinated Ti5c and two-coordinated O2c on the outermost surface (Fig. 2C), thus the possible O-coordinated active sites of Pt5c–O and Pt2c–O were finally constructed (22, 23), i.e., 1) Pt single atom substituted one Ti5c on the surface and bonded with five O atoms and 2) Pt single atom adsorbed on the surface of TiO2 and bonded with two two-coordinated surface O2c, respectively. Based on EXAFS analysis, it implied that Pt2c–O (coordination number = 2) and Pt5c–O (coordination number = 5) structures were coexisted on fresh Pt1/TiO2. Pt species were easily reduced in hydrogen-involved reactions (13), and the symmetry of the highest occupied molecular orbital in Pt5c sites matched antibonding of H2 well based on DFT calculations (SI Appendix, Figs. S7–S9). Notably, the electrons from 5d orbitals should be donated to the antibonding orbital of H2 molecules in the process of H2 activation, which indicated that symmetric matching of orbitals was necessary in the coupling process (SI Appendix, Fig. S7). Therefore, the symmetric matching of orbitals in the coupling process suggested that Pt5c sites were the active centers for H2 dissociation (24), which could accelerate the migration of O atoms and provide a reasonable site for the migration of lower coordinated Pt species (Pt2c) to form Pt–Ti bonds with subsurface Ti atoms. Therefore, the model with the adjacent coexistence structure of Pt5c and Pt2c sites was established to analyze the interaction between Pt5c and Pt2c sites (Fig. 2D).

To determine the formation mechanism of Pt–Ti bonds, the process of structural evolution was conducted by DFT calculations (Fig. 2F). First, the H2 molecule was adsorbed on Pt5c sites (−2.64 kcal/mol) and would dissociate into two adsorbed H* with an activation barrier of 15.41 kcal/mol (TS1). Then, one adsorbed H* would be transferred to the Pt5c–O1 to form H*–Pt5c–O1–H*, with a Pt2c–O1 bond distance of 2.16 Å (the original bond distance of Pt2c site with two bridge O atoms was 2.03 Å) (stage III). Next, the other H* was transferred to O1 atom to form H*–O1–H* with an activation barrier of 31.94 kcal/mol (TS2). In this case, the distance of Pt2c–O1 bonds was stretched to 2.34 Å (stage IV). As a result, the Pt2c atom could move and occupy O1 position by removing H2O molecule (H*–O1–H*), leading to that the Pt2c atom only bonded with Ti atoms instead of O atoms (donated as PtTi), and the optimization of used Pt1/TiO2 structural model is shown in Fig. 2E. The above calculations well explained the cleavage of Pt2c–O bonds and the formation of Pt–Ti bonds, leading to that the Pt2c site in fresh Pt1/TiO2 was evolved to the PtTi site in used Pt1/TiO2. More importantly, the formation of Pt–Ti bonds played an important role in suppressing the aggregation of Pt single atoms; the corresponding calculated structural parameters of transition states are shown in Fig. 2F and SI Appendix, Figs. S10–S12. The ab initio molecular dynamics simulation was further performed to prove that the structural model could serve as single-atom catalysts with high stability (SI Appendix, Figs. S13 and S14). Moreover, the Hirshfeld charges of Pt atoms were also calculated (SI Appendix, Figs. S15 and S16) (25). Because the Pt species were reduced, the charge of Pt2c and Pt5c sites, respectively, decreased from 0.204 and 0.611 eV to 0.117 and 0.508 eV, which was consistent with the XANES results and confirmed the formation of low-valence Pt single atoms.

Furthermore, quantitative least-squares EXAFS best-fitting analysis (Fig. 2 G and H and SI Appendix, Figs. S17 and S18 and Tables S2 and S3) was performed to verify the rationality of DFT calculations (26). We should always be aware that the structure information obtained by XAFS was an average result of the detected Pt element due to the coexistence of Pt5c and Pt2c (or PtTi) sites in our work (27), that was, Pt atoms only bonded with oxygen atoms with an average coordination number of ~4.7 over fresh Pt1/TiO2. Obviously for used Pt1/TiO2, the coordination number of Pt–O path decreased from 4.7 to 4.0 after reaction, which was caused by the evolution from Pt2c–O to Pt–Ti coordination. These results revealed that Pt5c atoms coordinated strongly with oxygen through occupying the position of Ti atoms, while Pt2c atoms initially coordinated with the surface oxygen of TiO2 and evolved to Pt–Ti coordination (PtTi sites) during the reaction.

Then, in situ Raman spectroscopy was used to further reveal the formation of PtTi sites during the reaction on Pt1/TiO2. As shown in Fig. 2I and SI Appendix, Fig. S19, we found that the vibrational characteristic peak (141 cm−1) of O–Ti–O species shifted to a higher wavenumber (149 cm−1) after introducing reactant gas (volume ratio NO:H2 ~ 1:1) (28). Besides, the quantum chemical calculation of vibrational Raman spectra was performed on fresh Pt1/TiO2 and used Pt1/TiO2; the results showed that the blue shift from 141 to 149 cm−1 (Fig. 2I) could be attributed to the formation of O–Ti–Pt structure, which disrupted the symmetry of the O–Ti–O network (SI Appendix, Figs. S20 and S21) (29). Therefore, both experimental and DFT results provided strong evidence for the formation of PtTi sites.

Besides, CO was used as the probe molecule in DRIFTS to characterize the Pt-based catalysts, which presented three bands centered at 2,090, 2,062, and 2,055 cm−1 (Fig. 3A). Among them, the band at 2,090 cm−1 was assigned to CO adsorption on Pt single atoms, and that at 2,055 cm−1 was assigned to the linear CO adsorption on Pt nanoparticles (30). Besides, the CO adsorption band was preserved on PtTi sites at 2,062 cm−1 over used Pt1/TiO2 (31), further showing the formation of Pt–Ti bonds after reaction. Meanwhile, the evolution of chemical environment of Pt atoms and the surface of TiO2 during the reaction was measured with quasi in situ XPS. After we introduced reactant gases (volume ratio NO:H2 ~ 1:1), the binding energy of Pt 4f on Pt1/TiO2 decreased rapidly (76.0 and 72.8 eV vs. 75.4 and 71.9 eV) (Fig. 3B), while there was no shift of the characteristic peak of Ti 2p (Fig. 3C). The O 1s XPS spectra were divided into two peaks centered at 531.4 and 529.7 eV, which were assigned to oxygen vacancies (denoted Oads) and lattice oxygen (denoted Olatt), respectively. Notably, the ratio of lattice oxygen and oxygen vacancy did not change in the hydrogen-involved reaction, as well as the results of in situ electron paramagnetic resonance (Fig. 3D and SI Appendix, Figs. S22–S25 and Table S4). These quasi in situ studies of XPS clearly showed that the singly dispersed Pt species were reduced in the mild reducing atmosphere, which could be related to the structural evolution of Pt single atoms based on DFT and EXAFS results. Additional study using hydrogen temperature-programmed reduction had revealed an outstanding low-temperature catalytic activity of Pt-based catalysts in the reduction of NO with H2 (SI Appendix, Fig. S26). In summary, all the DFT calculations and characterization data discussed above systematically confirmed the structural evolution of Pt single atoms during the reaction, leading to the formation of PtTi sites near Pt5c single atom sites.

Fig. 3.

Fig. 3.

Identification of Pt single atoms and quasi in situ studies of surface chemistry. (A) In situ CO adsorption DRIFTS of TiO2, fresh Pt1/TiO2, used Pt1/TiO2, and PtNP/TiO2. (B–D) Quasi in situ XPS spectra of Pt 4f (B), Ti 2p (C), and O 1s (D) of Pt1/TiO2 under different conditions.

Correlation of Cocatalytic Site and Catalytic Activity.

The catalytic activities of NO reduction with H2 over fresh Pt1/TiO2, used Pt1/TiO2, and PtNP/TiO2 were evaluated on a fix-bed reactor. More importantly, cyclic experiment was used to explore the correlation of cocatalytic site and catalytic activity. As shown in Fig. 4A and SI Appendix, Fig. S27, used Pt1/TiO2 exhibited better catalytic activity (ratios differ by 20%) and N2 selectivity than those of fresh Pt1/TiO2 in the temperature range of 75 to 150 °C due to the formation of PtTi sites, and the performance of used Pt1/TiO2 tended to stabilize in the third cycle. Then, in situ DRIFTS studies were further performed on fresh Pt1/TiO2 and used Pt1/TiO2 to reveal the effect of PtTi sites on the reaction (SI Appendix, Fig. S28), and the intensity of NO adsorption on used Pt1/TiO2 was clearly observed to be stronger than that on fresh Pt1/TiO2. Notably, the active intermediates of adsorbed NO species would be consumed to generate –NH2 amide species more quickly on used Pt1/TiO2, and the latter could further react with NO molecule to form N2 in the reduction of NO with H2. The above results confirmed the critical role of cocatalytic sites in promoting the process of NO conversion. Meanwhile, used Pt1/TiO2 showed a superior catalytic performance and Pt species were still isolated and uniformly dispersed in the presence of oxygen (SI Appendix, Figs. S29 and S30). Moreover, the apparent activation energy of reaction catalyzed by used Pt1/TiO2 (Ea = 50.6 kJ/mol) was also smaller than that of fresh Pt1/TiO2 (Ea = 63.6 kJ/mol) (Fig. 4B). Besides, for comparison, Pt nanoparticles (0.23 wt%) supported on TiO2 nanosheets (denoted as PtNP/TiO2) were also synthesized, which showed inferior catalytic activity and turnover frequency compared with single-atom catalysts (SI Appendix, Figs. S31–S34 and Table S5) (3234).

Fig. 4.

Fig. 4.

Catalytic activity of catalysts and DFT calculations. (A) The cyclic experiment over Pt1/TiO2 catalysts. The reaction condition is: [NO] = 1,000 ppm, [H2] = 1,000 ppm, balanced in Ar. GHSV = 60,000 mL g−1 h−1. (B) Arrhenius plot of NO + H2 reaction on fresh Pt1/TiO2 and used Pt1/TiO2. (C) DFT calculations of relative energy on Pt5c–O over fresh Pt1/TiO2 for H2 activation. (D) DFT calculations of relative energy on Pt5c–O and Pt5c–PtTi over used Pt1/TiO2 for H2 activation. (E) DFT-calculated potential energy diagrams and the barrier energy for the reduction of NO with H2 over used Pt1/TiO2.

Insights into the NO Reduction Boosted by Cocatalytic Sites.

The DFT calculations were conducted to further evaluate the role of PtTi cocatalytic sites for boosting catalytic performance. Considering that the adsorption and dissociation of reactant gas molecule was the initial and critical step of the reaction (35), the adsorption and activation of H2 was first examined over fresh Pt1/TiO2 and used Pt1/TiO2 (Fig. 4 C and D). As shown in Fig. 4C, the adsorption energy and dissociation barrier of H2 on the Pt5c site of fresh Pt1/TiO2 were −2.64 kcal/mol and 15.41 kcal/mol, respectively. After the formation of PtTi sites, the adsorption of H2 (−6.77 kcal/mol) was significantly enhanced on Pt5c site (Fig. 4D). Notably, there were two pathways (Pt5c–O path and Pt5c–PtTi path) for H2 dissociation over used Pt1/TiO2, in which energy barriers were only 14.23 kcal/mol and 3.20 kcal/mol, respectively. The results suggested that the adjacent PtTi sites assisted the H2 activation on the Pt5c single-atom site (Fig. 4 C and D and SI Appendix, Fig. S35). Meanwhile, it was also found that the barrier of H2 dissociation along Pt5c–O path got decreased over used Pt1/TiO2 (14.23 kcal/mol) compared to fresh Pt1/TiO2 (15.41 kcal/mol), which was caused by the formation of low-valence Pt single atoms.

Besides, the comprehensive DFT calculations were carried out to reveal the surface reaction mechanism of the reduction of NO (H2 + NO → H2O + N2) and the promotion mechanism of PtTi sites over fresh Pt1/TiO2 (SI Appendix, Fig. S37), used Pt1/TiO2 (Fig. 4E), and Pt (111) surface (SI Appendix, Fig. S38), which was initiated by the adsorption of H2 and NO and completed by the desorption of H2O and N2. The details on the reaction process of each step and the corresponding reaction path expression and reaction energy barrier are presented and discussed in Fig. 4E and SI Appendix, Figs. S36–S39 and Tables S6–S8. For used Pt1/TiO2, the catalytic cycle was initiated by the adsorption of NO on the PtTi site and the adsorption of H2 on the Pt5c site. Then, one adsorbed H* could be transferred to the adsorbed NO* to form HNO (NO* + H* → HNO) with an activation barrier of 2.08 kcal/mol (TS1), and the other H* would be transferred to the PtTi site to form HNOH with the energy barrier of 5.55 kcal/mol (TS2). The HHNO was generated from the HNOH (TS3, ~31.65 kcal/mol). In the next step, the second H2 molecule was introduced and activated, forming HNOH with the energy barrier of 31.18 kcal/mol. Thereafter, the third active H* would transfer to the N atom to form NH–HOH (TS5, ~32.57 kcal/mol), which promoted the removal of H2O and the formed NH2 structure. Subsequently, the second NO was introduced and adsorbed on the N site, leading to the formation of N–N (NNOHH → NN–HOH → N–N). Finally, the produced H2O and N2 were desorbed from the surface and the catalytic cycle was completed. By comparing the energy barriers of each transition state, it could be found that the energy barrier of TS5 (32.57 kcal/mol) was the highest, indicating that this step was the rate-determining step of the whole reaction. Notably, the DFT results suggested that the energy barrier of N–O activation on used Pt1/TiO2 (32.57 kcal/mol) was lower than that on fresh Pt1/TiO2 (36.50 kcal/mol) and Pt (111) nanoparticle surface (38.58 kcal/mol), revealing that the formation of PtTi sites could significantly boost single-atom catalysts for the reduction of NO.

Discussion

To further expand and verify the universal significance of cocatalytic sites, the catalytic activity was also confirmed to be promoted through the formation of cocatalytic sites over a series of single-atom catalysts (Pt1/Nb2O5, Pd1/TiO2, and Rh1/TiO2). HAADF-STEM and in situ CO adsorption DRIFTS revealed that the single atoms (Pt, Pd, and Rh) remained stable before and after reaction (SI Appendix, Figs. S40–S42). Notably, we found that cocatalytic sites were only formed in Pt1/Nb2O5 and Pd1/TiO2 after reaction (SI Appendix, Fig. S42), accompanied by the formation of low-valence metal active center (Fig. 5 A and B and SI Appendix, Figs. S43 and S44). Conversely, the cocatalytic sites were not identified in used Rh1/TiO2 (Fig. 5C and SI Appendix, Figs. S42 and S45). Correspondingly, the formation of cocatalytic sites boosted single-atom catalysts for NOx reduction due to the superior catalytic performance in cyclic experiment (Fig. 5 D–F).

Fig. 5.

Fig. 5.

Correlation of metal–metal bonds and catalytic activity. (A) Pt 4f XPS spectra of fresh Pt1/Nb2O5 and used Pt1/Nb2O5. (B) Pd 3d XPS spectra of fresh Pd1/TiO2 and used Pd1/TiO2. (C) Rh 3d XPS spectra of fresh Rh1/TiO2 and used Rh1/TiO2. (D–F) The cyclic experiment over Pt1/Nb2O5 (D), Pd1/TiO2 (E), and Rh1/TiO2 (F) catalysts. The reaction condition of D–F is: [NO] = 1,000 ppm, [H2] = 1,000 ppm, balanced in Ar. GHSV = 60,000 mL g−1 h−1.

Advances in highly active single-atom catalysts for addressing global atmosphere challenge may be realized by designing cocatalytic sites for efficient NO removal with H2. Taking Pt1/TiO2 as an example, in situ HAADF-STEM was used to monitor the reaction process and unravel the dynamic evolution of Pt single atom under real reaction condition. Combined with the EXAFS results, DFT calculations were performed to confirm the existence form of Pt single atoms and clarify the mechanism of structural evolution. The results revealed that Pt5c atoms coordinated strongly with oxygen through occupying the position of Ti atoms, while unstable Pt2c atoms initially coordinated with the surface oxygen of TiO2 and evolved to Pt–Ti coordination with subsurface Ti atoms during the reaction. Besides, DFT results definitely confirmed that the adjacent PtTi cocatalytic sites could significantly enhance H2 activation on Pt5c single-atom sites, resulting that the barrier of H2 dissociation along Pt5c–PtTi path got decreased to 3.20 kcal/mol. Compared with fresh Pt1/TiO2 (36.50 kcal/mol), the presence of PtTi sites also accelerated the whole de-NOx process by reducing the energy barrier of the NO activation, with the energy barrier of 32.57 kcal/mol. It was consistent with the results of catalytic activity that used Pt1/TiO2 exhibited better catalytic activity than fresh Pt1/TiO2 (ratios differ by 20%). Furthermore, the formation of cocatalytic sites was also confirmed on other single-atom catalysts (Pt1/Nb2O5 and Pd1/TiO2), and the role of cocatalytic sites for enhancing the catalytic activity was revealed. Our findings provided an understanding to rationally design highly active single-atom catalysts by introducing adjacent cocatalytic sites.

Materials and Methods

Synthesis of Fresh Pt1/TiO2 Catalyst.

First, 0.5 g TiO2 was dispersed in 50 mL deionized water, and 2.5 mL (1 mg mL−1) chloroplatinic acid (H2PtCl6, Aladdin) solution was dropped into the above suspension under vigorous stirring at 70 °C for 4 h. The mixture solution was then washed with deionized water by centrifugation to remove excess Pt atoms which were not stabilized by the TiO2 support. Finally, the catalysts were obtained by drying at 80 °C overnight and calcining at 300 °C for 4 h in flowing air after ramping up the temperature at a rate of 1 °C min−1.

Synthesis of Fresh Pt1/Nb2O5, Fresh Pd1/TiO2, and Fresh Rh1/TiO2 Catalysts.

The preparation of fresh Pt1/Nb2O5 catalyst is same as that of fresh Pt1/TiO2 catalyst except that Nb2O5 (Aladdin, ≥99% purity) replaced TiO2 as support. And the preparation of fresh Pd1/TiO2 and fresh Rh1/TiO2 catalysts is same as that of fresh Pt1/TiO2 catalyst except that palladium chloride (PdCl2, Aladdin) and rhodium chloride (RhCl3, Aladdin) solution replaced H2PtCl6 as precursor.

In Situ HAADF-STEM.

The HAADF-STEM images and elemental mapping results were obtained by a Titan 80-300 scanning transmission electron microscopy (STEM) operated at 100 kV and 1 × 10−8 ~ 1 × 10−9 Pa with double-aberration correctors and four EDS detectors. For in situ HAADF-STEM, the sample was encapsulated into a chip from the DENS solutions company before the experiments, whose specified range of error for temperature was ≤5%. The holder was then inserted into Titan 80-300 STEM and the air tightness of the system was checked. Argon and reaction gas (H2:NO ~ 1:1) were introduced to a system of direct-current plasma-enhanced chemical vapor deposition at 125 °C. The chamber pressure was 1 bar.

X-Ray Absorption Spectra Collection and Data Processing.

The X-ray absorption find structure spectra (XAFS) experiment was conducted on the beamline 1W1B station in Beijing Synchrotron Radiation Facility. The EXAFS data were analyzed by ATHENA module implemented in the IFEFFIT software packages (36), and relevant quantitative structural parameters were obtained by ARTEMIS module of IFEFFIT software packages (37).

Method and Model for DFT Calculations.

First-principles calculations were executed by DFT from cambridge sequential total energy package (CASTEP) package with the plane-wave pseudopotential method (38). The interaction of ion core and valence electron was proved by Vanderbilt-type ultrasoft pseudopotential (39). The exchange–correlation interactions were treated by the generalized-gradient approximation (GGA) with spin-polarized Perdew–Burke–Ernzerhof scheme (40). The energy cutoff for the plane-wave basis was set to be 400 eV. The convergence thresholds between optimization cycles for energy change and maximum force were set as 10−5 eV/atom and 0.03 eV/Å, respectively. According to the surface energy values, the (101) surface was the most abundant one with high thermodynamic stability, followed by (100) and (001) surfaces (41). The corresponding p(2 × 2) TiO2 (101) surface was constructed using the same method of our previous work (22, 24, 42). The slab contained three Ti2O4 layers, i.e., 48 O atoms and 24 Ti atoms. A 15-Å vacuum layer between the images in the direction of the surface normal was utilized. The models of Pt-loaded catalyst were constructed based on the experimental results (especially the results of XAFS) together with the results of our previous theoretical studies (22). A previous study had determined that GGA was reliable for transition state calculations in anatase TiO2 (101) surface when compared with GGA+U calculations (22, 43, 44); hence, only GGA was adopted in this part (SI Appendix, Fig. S46 and Table S9). The Hirshfeld charge population analysis was used to determine the atomic charge. The adsorb energies of adsorbed species were defined as:

ΔEads = Emol/sub - Esub - Emol,

where Eads was the total energy of the species adsorbing on the substrate. Emol/sub was the total energy of the complex consisting of substrate and adsorbed species. Esub and Emol were the energy of isolated substrate and adsorbed species, respectively.

Theoretical vibrational Raman spectra were performed within density-functional perturbation theory as implemented in Quantum-ESPRESSO code, which was an integrated suite of computer codes for electronic-structure calculations and materials modeling, and the acronym ESPRESSO stands for opEn Source Package for Research in Electronic Structure, Simulation, and Optimization. (45). The local density approximation to the exchange–correlation function with norm-conserving pseudopotential was employed for the calculation of phonon frequencies (46); 2, 12, and 18 electrons were considered as valence electrons in O, Ti, and Pt elements, respectively. To achieve the converged results, a plane-wave kinetic energy cutoff of 70 Ry (952 eV) was used for the wave functions, and the convergence threshold was set to 10−10 eV and 10−12 eV in the electron and phonon self-consistent calculation, respectively. A Monkhorst–Pack k-point mesh of 1 × 1 × 1 was used to sample the Brillouin Zones for slab systems in Raman spectra relative calculations.

Supplementary Material

Appendix 01 (PDF)

Acknowledgments

We gratefully acknowledge the financial support by the Natural Science Foundation of China as general projects (grant Nos. 22225604, 22076082, 21874099, 22176140, 22006029, 42277059, and 22206090), the Tianjin Commission of Science and Technology as key technologies R&D projects (grant Nos. 19YFZCSF00740, 20YFZCSN01070, and 21YFSNSN00250), and the Frontiers Science Center for New Organic Matter (grant No. 63181206). We acknowledge Dr. W. Xi at Tianjin University of Technology for the in situ high-angle annular dark-field scanning transmission electron microscopy characterization and helpful discussions, and Haihe Laboratory of Sustainable Chemical Transformations.

Author contributions

P.W., G.L., and S.Z. designed research; G.L. and Z.H. performed research; L.Z. contributed new reagents/analytic tools; G.L., H.Z., Y.L., and W.S. analyzed data; and P.W., G.L., Z.H., and S.Z. wrote the paper.

Competing interests

The authors declare no competing interest.

Footnotes

This article is a PNAS Direct Submission.

Data, Materials, and Software Availability

The shared data (including information, code analyses, sequences, etc.) are not included for the data involved in our manuscript, and all the necessary study data are included in the article and/or SI Appendix.

Supporting Information

References

  • 1.Oberschelp C., Pfister S., Raptis C. E., Hellweg S., Global emission hotspots of coal power generation. Nat. Sustain. 2, 113–121 (2019). [Google Scholar]
  • 2.Li K., et al. , A two-pollutant strategy for improving ozone and particulate air quality in China. Nat. Geosci. 12, 906–910 (2019). [Google Scholar]
  • 3.Johnston F. H., et al. , Unprecedented health costs of smoke-related PM2.5 from the 2019–20 Australian megafires. Nat. Sustain. 4, 42–47 (2020). [Google Scholar]
  • 4.Liu Z., Li J., Woo S. I., Recent advances in the selective catalytic reduction of NOx by hydrogen in the presence of oxygen. Energy Environ. Sci. 5, 8799–8814 (2012). [Google Scholar]
  • 5.Hibbitts D. D., et al. , Catalytic NO activation and NO-H2 reaction pathways. J. Catal. 319, 95–109 (2014). [Google Scholar]
  • 6.Hecker W. C., Bell A. T., Reduction of NO by H2 over silica-supported rhodium: infrared and kinetic studies. J. Catal. 92, 247–259 (1985). [Google Scholar]
  • 7.Farberow C. A., Dumesic J. A., Mavrikakis M., Density functional theory calculations and analysis of reaction pathways for reduction of nitric oxide by hydrogen on Pt(111). ACS Catal. 4, 3307–3319 (2014). [Google Scholar]
  • 8.Zhang Z., et al. , Thermally stable single atom Pt/m-Al2O3 for selective hydrogenation and CO oxidation. Nat. Commun. 8, 16100 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Wang L., et al. , Boosting activity and stability of metal single-atom catalysts via regulation of coordination number and local composition. J. Am. Chem. Soc. 143, 18854–18858 (2021). [DOI] [PubMed] [Google Scholar]
  • 10.Zhou M., et al. , On the mechanism of H2 activation over single-atom catalyst: an understanding of Pt1/WOx in the hydrogenolysis reaction. Chinese J. Catal. 41, 524–532 (2020). [Google Scholar]
  • 11.Chen L., et al. , Unlocking the catalytic potential of TiO2-supported Pt single atoms for the reverse water-gas shift reaction by altering their chemical environment. J. Am. Chem. Soc. 1, 977–986 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Nguyen L., et al. , Reduction of nitric oxide with hydrogen on catalysts of singly dispersed bimetallic sites Pt1Com and Pd1Con. ACS Catal. 6, 840–850 (2016). [Google Scholar]
  • 13.Piccolo L., et al. , Operando X-ray absorption spectroscopy investigation of photocatalytic hydrogen evolution over ultradispersed Pt/TiO2 catalysts. ACS Catal. 10, 12696–12705 (2020). [Google Scholar]
  • 14.Han A., et al. , An adjacent atomic platinum site enables single-atom iron with high oxygen reduction reaction performance. Angew. Chem. Int. Ed. 133, 19411–19420 (2021). [DOI] [PubMed] [Google Scholar]
  • 15.Yin S., et al. , Construction of highly active metal-containing nanoparticles and FeCo-N4 composite sites for the acidic oxygen reduction reaction. Angew. Chem. Int. Ed. 59, 21976–21979 (2020). [DOI] [PubMed] [Google Scholar]
  • 16.Zhao Z., et al. , Boosting nitrogen activation via bimetallic organic frameworks for photocatalytic ammonia synthesis. J. Am. Chem. Soc. 11, 9986–9995 (2021). [Google Scholar]
  • 17.Zhao S., et al. , Electronically and geometrically modified single-atom Fe sites by adjacent Fe nanoparticles for enhanced oxygen reduction. Adv. Mater. 34, 2107291 (2022). [DOI] [PubMed] [Google Scholar]
  • 18.Macino M., et al. , Tuning of catalytic sites in Pt/TiO2 catalysts for the chemoselective hydrogenation of 3-nitrostyrene. Nat. Catal. 2, 873–881 (2019). [Google Scholar]
  • 19.Spezzati G., et al. , Atomically dispersed Pd-O species on CeO2(111) as highly active sites for low-temperature CO oxidation. ACS Catal. 7, 6887–6891 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Xi W., et al. , Dynamic co-catalysis of Au single atoms and nanoporous Au for methane pyrolysis. Nat. Commun. 11, 1919 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Wang L., et al. , Catalysis and in situ studies of Rh1/Co3O4 nanorods in reduction of NO with H2. ACS Catal. 3, 1011–1019 (2013). [Google Scholar]
  • 22.Wang X., Zhang L., Bu Y., Sun W., Interplay between invasive single atom Pt and native oxygen vacancy in anatase TiO2(101) surface: A theoretical study. Appl. Surf. Sci. 540, 148357 (2021). [Google Scholar]
  • 23.Wang X., Zhang L., Bu Y., Sun W., Interplay between invasive single atom Pt and native oxygen vacancy in rutile TiO2(110) surface: A theoretical study. Nano Res. 15, 669–676 (2022). [Google Scholar]
  • 24.Chen Y., et al. , Engineering the atomic interface with single platinum atoms for enhanced photocatalytic hydrogen production. Angew. Chem. Int. Ed. 59, 1295–1301 (2020). [DOI] [PubMed] [Google Scholar]
  • 25.Wang P., et al. , Unraveling the interfacial charge migration pathway at the atomic level in a highly efficient Z-scheme photocatalyst. Angew. Chem. Int. Ed. 58, 11329–11334 (2019). [DOI] [PubMed] [Google Scholar]
  • 26.Wang P., et al. , Atomic insights for optimum and excess doping in photocatalysis: A case study of few-layer Cu-ZnIn2S4. Adv. Funct. Mater. 29, 1807013 (2019). [Google Scholar]
  • 27.Qin R., Liu K., Wu Q., Zheng N., Surface coordination chemistry of atomically dispersed metal catalysts. Chem. Rev. 120, 11810–11899 (2020). [DOI] [PubMed] [Google Scholar]
  • 28.Tian F., Zhang Y., Zhang J., Pan C., Raman spectroscopy: A new approach to measure the percentage of anatase TiO2 exposed (001) facets. J. Phys. Chem. C 116, 7515–7519 (2012). [Google Scholar]
  • 29.Liu G., et al. , Efficient promotion of anatase TiO2 photocatalysis via bifunctional surface-terminating Ti-O-B-N structures. J. Phys. Chem. C 113, 12317–12324 (2009). [Google Scholar]
  • 30.Han B., et al. , Strong metal-support interactions between Pt single atoms and TiO2. Angew. Chem. Int. Ed. 59, 11824–11829 (2020). [DOI] [PubMed] [Google Scholar]
  • 31.Li J., et al. , In situ formation of isolated bimetallic PtCe sites of single-dispersed Pt on CeO2 for low-temperature CO oxidation. ACS Appl. Mater. Interfaces 10, 38134–38140 (2018). [DOI] [PubMed] [Google Scholar]
  • 32.Xu J., et al. , Organic wastewater treatment by a single-atom catalyst and electrolytically produced H2O2. Nat. Sustain. 4, 233–241 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Nie L., et al. , Activation of surface lattice oxygen in single-atom Pt/CeO2 for low-temperature CO oxidation. Science 358, 1419–1423 (2017). [DOI] [PubMed] [Google Scholar]
  • 34.Abdel-Mageed A. M., et al. , Highly active and stable single-atom Cu catalysts supported by a metal-organic framework. J. Am. Chem. Soc. 141, 5201–5210 (2019). [DOI] [PubMed] [Google Scholar]
  • 35.Li Z., et al. , Oxidation of reduced ceria by incorporation of hydrogen. Angew. Chem. Int. Ed. 58, 14686–14693 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Ravel B., Newville M., ATHENA, ARTEMIS, HEPHAESTUS: Data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 12, 537–541 (2005). [DOI] [PubMed] [Google Scholar]
  • 37.Rehr J. J., Albers R. C., Theoretical approaches to X-ray absorption fine structure. Rev. Mod. Phys. 72, 621–654 (2000). [Google Scholar]
  • 38.Clark S., et al. , First principles methods using CASTEP. Z. Kristallogr Cryst. Mater. 220, 567–570 (2005). [Google Scholar]
  • 39.Vanderbilt D., Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B Condens. Matter 41, 7892–7895 (1990). [DOI] [PubMed] [Google Scholar]
  • 40.Perdew J., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
  • 41.Diebold U., The surface science of titanium dioxide. Surf. Sci. Rep. 48, 53–229 (2003). [Google Scholar]
  • 42.Chen Y., et al. , Discovering partially charged single-atom Pt for enhanced anti-markovnikov alkene hydrosilylation. J. Am. Chem. Soc. 140, 7407–7410 (2018). [DOI] [PubMed] [Google Scholar]
  • 43.Li Y., Gao Y., Interplay between water and TiO2 anatase (101) surface with subsurface oxygen vacancy. Phys. Rev. Lett. 112, 206101 (2014). [DOI] [PubMed] [Google Scholar]
  • 44.Zhu L., Hu Q., Yang R., The effect of electron localization on the electronic structure and migration barrier of oxygen vacancies in rutile. J. Phys. Condens. Matter 26, 055602 (2014). [DOI] [PubMed] [Google Scholar]
  • 45.Giannozzi P., et al. , QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 21, 395502 (2009). [DOI] [PubMed] [Google Scholar]
  • 46.Perdew J. P., Zunger A., Self-interaction correction to density-functional approximations for many-electron systems. Phys. Rev. B 23, 5048–5079 (1981). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Appendix 01 (PDF)

Data Availability Statement

The shared data (including information, code analyses, sequences, etc.) are not included for the data involved in our manuscript, and all the necessary study data are included in the article and/or SI Appendix.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES