DOI:
10.1039/D5TA03582A
(Paper)
J. Mater. Chem. A, 2025,
13, 29943-29955
Insights into the catalytic valorization of industrial high-concentration nitrous oxide for propylene synthesis
Received
6th May 2025
, Accepted 27th July 2025
First published on 28th July 2025
Abstract
Industrial nitrous oxide (N2O) emissions at high concentrations pose a significant challenge to climate change, while the oxidative dehydrogenation of propane (ODHP) with N2O presents a promising strategy for N2O valorization. However, the activity-selectivity trade-off and rapid deactivation by coke deposition present challenges for future applications. Herein, we select Pd/TiO2 as a model catalyst and design spatially separated redox centers on the rutile phase catalyst (Pd/R-TiO2), which enable parallel adsorption and independent activation of C3H8 on metallic Pd sites and N2O on oxygen vacancies. Furthermore, the catalyst suppresses product overoxidation by reducing reactive oxygen species concentration near C3 intermediates, enhancing C3H6 selectivity while inhibiting enolic intermediate and coke formation to ensure catalyst stability. Hydrogen spillover from Pd sites to the TiO2 surface connects the separated redox reactions, completing the catalytic cycle. This work presents a rational strategy for catalyst design in N2O-ODHP, contributing to the sustainable mitigation of non-CO2 greenhouse gases.
1. Introduction
For the first time, global warming exceeded 1.5 °C across an entire year above pre-industrial levels in 2024, emphasizing the urgent necessity for immediate and sustained efforts to combat climate change.1,2 Beyond CO2, attention should be given to non-CO2 greenhouse gases, including CH4, N2O, and fluorinated gases (F-gases). Among these, N2O is notably the third largest greenhouse gas for climate change, with a global warming potential 273 times higher than that of CO2.3 Human activities have accelerated N2O emissions, with an estimated increase rate of 2% per decade and a lifetime exceeding 110 years.4,5 Worse still, it is the single greatest stratospheric ozone-depleting substance of the 21st century, posing a severe environmental threat.6 Nevertheless, N2O has received less attention compared to CO2 and CH4. To achieve the goal of carbon neutrality, efficiently mitigating N2O emissions from both low- and high-concentration sources is crucial but challenging. N2O direct decomposition, selective catalytic reduction (SCR), and N2O revalorization as an oxidant are three major technologies for mitigating N2O emissions.7,8 For low-concentration scenarios, the first two technologies can achieve efficient abatement of N2O, but for high-concentration scenarios, where N2O can be considered for waste recovery, direct decomposition or SCR would undoubtedly result in resource waste. Specifically, industrial N2O emissions are expected to reach 1.4 Mt N2O–N per year by 2050,5 a challenge that will escalate in the coming decades. For example, N2O generated at high concentrations and in large quantities as a by-product in the production of adipic acid (30–40 vol%) or caprolactam (10–20 vol%) can serve as a low-cost and ideal N2O source.9,10 Valorizing N2O offers significant advantages for mitigating the environmental impact of its industrial emissions.11
On the other hand, the global demand for propylene is experiencing rapid growth, leading to a “propylene gap”,12 which traditional methods, such as steam cracking and fluid catalytic cracking, are struggling to fill.13–16 Direct propane dehydrogenation (PDH) has been studied and industrialized to address this issue,17,18 driven by a sudden increase in propane supply from the shale gas revolution.19 The harsh conditions required (typically ≥650 °C) due to its endothermic nature tend to cause excessive energy consumption. Specifically, it is estimated that the potential energy savings for moving to oxidative dehydrogenation of propane (ODHP) would be ∼45% of the energy consumption.13 Compared to the common oxidant O2, which tends to cause overoxidation, harnessing N2O as a milder oxidant can improve the selectivity of the reaction. Therefore, ODHP using N2O as a mild oxidant (N2O-ODHP: C3H8 + N2O → C3H6 + N2 + H2O) represents an emerging candidate technology for propylene production and N2O valorization.20–22 However, inadequate atomic-level understanding of the underlying reaction mechanisms has impeded the development of highly efficient catalysts in N2O-ODHP. Significant challenges remain in breaking the trade-off between activity and selectivity while addressing the tendency for rapid deactivation caused by coke deposition, with current strategies primarily relying on trial-and-error methods, resulting in high costs and limited success rates.23
Herein, a series of Pd/TiO2 catalysts were experimentally and theoretically studied as model catalysts in N2O-ODHP. Anatase and rutile titanium oxide supports are often chosen for mechanistic studies due to their well-understood properties and versatility.24–26 Palladium (Pd) was selected as the primary active metal because of its well-documented C–H cleavage ability and use in propane dehydrogenation.27–29 The results demonstrated that the Pd/R-TiO2 catalyst significantly boosted propylene production, although it had the same chemical composition as others. This begs critical questions: What is the structural feature of Pd/R-TiO2? Does the performance stem from the presence of redox center separation? Our findings highlight that redox center separation can break the activity-selectivity trade-off while ensuring high catalyst stability for N2O-ODHP. This study paves the way for the rational design of efficient catalysts, advancing future application of N2O-ODHP for sustainable N2O valorization and propylene production.
2. Methods
2.1. Chemicals and materials
All chemicals were purchased from Aladdin Chemical Reagent Company, while all gases utilized in this study were sourced from Hangzhou Jingong Special Gases Co., Ltd. All chemicals used in the experiments were reagent grade or higher and used as received without further purification. Experimental solutions were prepared using deionized (DI) water.
2.2. Synthesis of catalysts
Pd/A-TiO2 and Pd/R-TiO2 catalysts were prepared via sol immobilization.30 Only the TiO2 supports were varied, including anatase and rutile TiO2. Aqueous solutions of the metal precursors, Pd(NO3)2, were added to deionized water (140 mL) and stirred vigorously. PVA was then added to the solution, with a mass ratio of PVA to Pd of 1
:
1. Subsequently, a freshly prepared NaBH4 solution (0.15 M, with a NaBH4 to total metal ratio of 4
:
1, mole per mole) was added to the solution immediately, resulting in the formation of a sol. Following this, the support material (0.5 g) was added to the colloidal solution while stirring, facilitating the immobilization of the metal nanoparticles. After 30 min, the solid catalyst was recovered by filtration and washed with 500 mL of distilled water to remove Na+, NO3−, BH4−, and BO2− ions. Subsequently, the catalysts were dried overnight under vacuum at 80 °C. Finally, the as-obtained powders were calcined at 550 °C for 3 h in flowing Ar to prepare Pd/A-TiO2 and Pd/R-TiO2.
PdO/A-TiO2 and PdO/R-TiO2 catalysts were prepared using the impregnation method, with only the TiO2 supports being varied. 1.0 g of TiO2 support and a defined amount of Pd(NO3)2 were mixed in 140 mL of H2O under sonication for 30 min and stirring for 4 h. The catalysts were then dried overnight. The obtained powders were calcined at 550 °C for 3 h in static air to prepare PdO/R-TiO2 and PdO/A-TiO2 catalysts.
2.3. Characterizations
The actual metal loadings of the catalysts were determined using inductively coupled plasma-optical emission spectroscopy (ICP-OES) with a Thermo Fisher iCAP PRO instrument. Transmission electron microscopy (TEM) was conducted using a Tecnai G2 F20 S-TWIN instrument from FEI Company, USA, operating at an accelerating voltage of 200 kV. Powder X-ray diffraction (XRD) patterns were recorded using an X'Pert3 Powder analytical diffractometer (B.V., Netherlands) at a scan rate of 5 °C min−1 with monochromatized Cu Kα radiation. Ultraviolet-visible diffuse reflectance spectra (UV-vis DRS) were obtained using a UV-2600 Shimadzu spectrophotometer equipped with an integrating sphere assembly at wavelengths ranging from 200 to 800 nm, with a resolution of 0.5 nm. BaSO4 was used as the reflectance standard under ambient conditions. X-ray photoelectron spectroscopy (XPS) experiments were carried out on an ESCALAB 250 (Thermo, USA) using standard Al Kα X-ray radiation. All spectra were calibrated using the C 1s peak at 284.8 eV. Raman spectra were recorded on a HORIBA Scientific LabRAM HR Evolution spectrometer using a He/Cd excitation laser at 532 nm. Electron paramagnetic resonance (EPR) spectra were obtained using a Bruker EMXplus spectrometer at liquid nitrogen temperature (77 K).
Operando NAP-XPS spectra were recorded using a Thermo Scientific ESCALAB 250Xi instrument. Monochromatized Al Kα irradiation (1486.6 eV) with a pass energy of 20.0 eV was employed to collect all spectra. A pretreatment chamber was utilized to expose the catalysts to specific treatments. After treatment, the chamber was evacuated, and the sample was directly transferred to the vacuum analysis chamber without exposure to air.
H2-TPR experiments were conducted with a TP-5089 (Tianjin Xianquan Industry and Trade Development Co., China). The samples were pretreated prior to measurements at 200 °C for 1 h with pure He and then cooled to 50 °C. The baselines were allowed to sufficiently stabilize in 5% H2/Ar flow at 50 °C before commencing temperature programming, followed by heating to 800 °C in 5% H2/Ar flow at a heating rate of 5 °C min−1. The consumption of H2 was monitored using a thermal conductivity detector.
In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) experiments were conducted in the range of 4000–400 cm−1 with a spectral resolution of 4 cm−1 using a Bruker Tensor 27 spectrometer equipped with ZnSe windows. The spectrometer also featured a high-sensitivity MCT detector cooled by liquid nitrogen and a high-temperature reaction chamber under ambient pressure. All spectra were obtained by averaging 64 scans with a resolution of 4 cm−1. Before each experiment, the sample was pretreated in He at 400 °C for 30 min to degas (i.e. gaseous H2O and hydroxy species), followed by the collection of a background spectrum.
Thermogravimetric analysis (TGA) and derivative thermogravimetry (DTG) experiments were conducted with a NETZSCH STA 449F3.31 TGA and DTG measured the amount and rate of change in the weight of a material as a function of temperature or time in a controlled atmosphere. Here, the sample was heated from 30 to 600 °C in a platinum crucible with a heating rate of 10 °C min−1 in flowing air.
2.4. Catalytic evaluations
The N2O-ODHP reactions were conducted in a fixed-bed reactor constructed from stainless steel. In a typical experiment, 500 mg of sieved sample (40–60 mesh) was loaded between two layers of quartz wool, and the designed weight hourly space velocity (WHSV) was set at 12
000 mL gcat−1·h−1. The feed gas consisted of 16.7% C3H8, 16.7% N2O, balanced with He; 2% Ar in the flow was used as the inner standard. The quantitative N2O analysis of the feed and effluent gas was measured with an FTIR spectrometer (G210 analyser, 0.1–100%), purchased from Geotechnical Instruments (UK) Ltd. The products of the reaction were analyzed with an online gas chromatograph (GC, Agilent 7890A) equipped with a thermal conductivity detector (TCD) using He as the carrier gas with Porapak Q and TDX-01 packed columns and a flame ionization detector (FID) with a HP-PLOTQ capillary column. The probable products of the N2O-ODHP were C3H6, C2H6, C2H4, CH4, CO, and CO2. The carbon balance between the reactant (C3H8) and products (i.e., CO, CO2, CH4, C2H4, C2H6, and C3H6) reached ≥98%, as determined by GC.
2.5. Computational details
The geometrical optimization and transition state (TS) calculations were performed using the Vienna ab initio simulation package (VASP) code (version 5.4.4).32,33 Other details, including catalytic evaluations and DFT calculations, are shown in the SI.
3. Results and discussion
3.1. Synthesis and characterizations
Rutile (denoted as R) and anatase (denoted as A) TiO2 were used as the support. The home-made anatase displayed a uniform distribution of Ti and O atoms, as determined by scanning transmission electron microscopy (STEM) and elemental mapping (Fig. S1). Anatase and rutile had surface areas of 51.7 and 48.0 m2 g−1 (Fig. S2), respectively, providing comparable surfaces for 0.5 wt% Pd loadings. PdO/TiO2 catalysts were prepared for comparison using wet impregnation (Fig. 1a), labeled as PdO/R-TiO2 (Fig. 1b) and PdO/A-TiO2 (Fig. 1c). PdNP/TiO2 catalysts were synthesized by the sol immobilization method (Fig. 1d),30,34 denoted as Pd/R-TiO2 (Fig. 1e) and Pd/A-TiO2 (Fig. 1f). X-ray diffraction (XRD) showed comparable diffraction patterns with rutile TiO2 (JCPDS No. 21-1276) and anatase TiO2 (JCPDS No. 21-1272) for R-TiO2 and A-TiO2, respectively (Fig. S3). No reflections corresponding to crystalline Pd or PdO were observed, likely due to the low content and high dispersion of Pd species on the support. Notably, we chose NaBH4 instead of H2 as the reductant30 because chemical reduction at room temperature avoids high-temperature hydrogen-mediated thermal reduction that may cause a phase transformation from anatase to rutile, which would obscure the effects of the TiO2 crystalline phase on the Pd–TiO2 synergy. This is supported by the XRD profile of H2-reduced PdO/A-TiO2 at 550 °C (Fig. S4), showing a mixture of anatase and rutile crystalline regions.
 |
| Fig. 1 (a) Schematic of the impregnation method and (b and c) TEM images of PdO/R-TiO2 (b) and PdO/A-TiO2 (c) and (d) schematic of the sol immobilization method and (e and f) TEM images of Pd/R-TiO2 (e) and Pd/A-TiO2 (f); (g) XPS spectra of Pd 3d for all the Pd/TiO2 catalysts; (h) UV-vis DRS plots of TiO2 and Pd/TiO2 catalysts; (i) CO-DRIFTS spectra for all the Pd/TiO2 catalysts. Pd, mazarine; Ti, grey; O, red. | |
Transmission electron microscopy (TEM) images showed the exposed facets of TiO2 were anatase (101) and rutile (110), with lattice distances of 0.35 nm and 0.33 nm, respectively. The Pd species on both PdO/R-TiO2 and PdO/A-TiO2 were PdO, exhibiting an exposed facet of PdO(002) with a lattice distance of 0.27 nm. For Pd/A-TiO2 and Pd/R-TiO2, metallic Pd NPs with exposed Pd(111) surfaces with a lattice distance of 0.23 nm were observed, all with a similar size of ∼10 nm, as confirmed by line scanning (Fig. S5). Additionally, elemental mapping on Pd/R-TiO2 is also presented (Fig. S6), excluding the presence of TiOx overlayers due to strong metal-support interaction.18,35
In Pd 3d X-ray photoelectron spectroscopy (XPS, Fig. 1g), peaks at 334.4 and 339.7 eV, and at 335.2 and 340.5 eV, can be assigned to Pd0 and Pd2+ states, respectively.36 The results suggest that the Pd species on Pd/A-TiO2 and Pd/R-TiO2 predominantly exist as metallic Pd0 species, while the Pd species on both PdO/R-TiO2 and PdO/A-TiO2 are mainly in the oxidized Pd2+ form. Ultraviolet-visible diffuse reflectance spectra (UV-vis DRS, Fig. 1h) showed an absorption in the range of 200–400 nm that was associated with anatase and rutile TiO2, while a broader absorption peak at 450–550 nm was attributed to a d–d transition of PdO species for PdO/TiO2.37 Moreover, a significant absorption in the visible region above 600 nm was observed, which was attributed to a characteristic signal for Pd NPs for PdNP/TiO2.38 CO-adsorption diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) showed a broad band at 1900–2000 cm−1, attributed to Pd NPs,39 for PdNP/TiO2 (Fig. 1i). A Pd2+–CO band at 2185 cm−1 was also detected for PdO/TiO2. This suggested that the Pd species in PdNP/TiO2 were mainly Pd NPs, while those in PdO/TiO2 were primarily PdO species.
Raman spectra were also obtained (Fig. S7). All characteristic Raman shifts of anatase TiO2 were observed at 144 cm−1 (Eg), 198 cm−1 (Eg), 398 cm−1 (B1g), 520 cm−1 (A1g), and 638 cm−1 (Eg).25 The shifts to higher wavenumbers of the Eg peaks indicated the formation of the strongest Pd–TiO2 interactions and the most abundant oxygen vacancies. By contrast, four typical vibrational modes were observed for rutile TiO2 around 145 cm−1 (B1g), 445 cm−1 (Eg), 610 cm−1 (A1g), and 240 cm−1 for second-order effects.40 The redshifts to lower wavenumbers of the A1g and Eg vibration modes for Pd/R-TiO2 indicated the weakened bond strength of O–Ti–O. This implied facile formation of oxygen vacancies in TiO2, which could significantly facilitate the activation of N2O.
3.2. Catalytic performance and the identification of key intermediates
N2O-ODHP catalytic performance was evaluated in a fixed-bed reactor under a WHSV of 12
000 mL gcat−1 h−1, at a C3H8/N2O/He ratio of 1/1/4, where 16.7% N2O concentration is in the range of the tail gas emitted from caprolactam production (10–20 vol%).10 As shown in Fig. 2a–d, as the reaction temperature increased, both C3H8 and N2O conversions improved, but the selectivity for C3H6 declined. Furthermore, an increase in the CO ratio and the production of other C2 and C1 byproducts were observed. Therefore, an optimal reaction temperature was determined as 500 °C. Pd/R-TiO2 exhibited the highest C3H6 selectivity (67.0%) and C3H6 space-time yield (STY) of 74.2 kgC3H6 kgPd−1 h−1 with a C3H8 conversion of 14.2%. The catalytic performance of 9.5% yield outperforms those of previously reported benchmark catalysts (Table S1).20,41–43 PdO/R-TiO2 showed high C3H6 selectivity (33.1%) and STY (45.7 kgC3H6 kgPd−1 h−1) at higher C3H8 conversion (18.9%). In contrast, PdO/A-TiO2 demonstrated a high conversion (53.9%) but low C3H8 conversion (7.0%) and C3H6 selectivity (22.9%), resulting in a low STY (11.3 kgC3H6 kgPd−1 h−1). Pd/A-TiO2, while achieving the lowest N2O conversion (17.4%), showed slightly better C3H6 selectivity (24.5%) and STY (19.3 kgC3H6 kgPd−1 h−1). Overall, the C3H6 STY follows the order Pd/R-TiO2 > PdO/R-TiO2 > Pd/A-TiO2 > PdO/A-TiO2. Additionally, the O2-ODHP performance of Pd/R-TiO2 was examined (Fig. S8), showing C3H6 selectivities of less than 10% at different temperatures and a C3H8/O2 ratio of 1/1. The results showed lower C3H6 selectivity and overoxidation when using O2 as the oxidant, suggesting the significant role of N2O in increasing the selectivity.
 |
| Fig. 2 (a–d) Catalytic performance of PdO/A-TiO2 (a), Pd/A-TiO2 (b), PdO/R-TiO2 (c), and Pd/R-TiO2 in N2O-ODHP at 400–550 °C (d). Histogram showing the selectivity for different products; hollow circle, C3H8 conversion; solid circle, N2O conversion; hollow green star, C3H6 STY. (e and f) Catalytic stability of PdO/R-TiO2 (e) and Pd/R-TiO2 (f) for N2O-ODHP at 500 °C. Typical reaction conditions: 100 mL min−1 reaction feed of 16.7% C3H8, 16.7% N2O, He balance, Ar (2%) as the inner standard; 500 mg of catalyst; 12 000 mL gcat−1 h−1. (g and h) In situ DRIFT spectra of adsorbed species on the PdO/R-TiO2 (g) and Pd/R-TiO2 (h) catalysts as a function of time at 500 °C under a flow of 16.7% C3H8, 16.7% N2O, and He balance. | |
To assess potential structural or compositional changes during the N2O-ODHP reaction, the catalysts were subjected to the standard reaction conditions (500 °C, 100 mL min−1 feed of 16.7% C3H8, 16.7% N2O, balance He) for 6 h. After cooling to room temperature, the system was restarted under the same conditions,8 and the catalytic performance of the used catalysts was re-evaluated (Table S2). Compared to their fresh performance, the used PdO/A-TiO2 exhibited a slight decrease in N2O conversion (from 53.9% to 43.7%) and C3H8 conversion (from 7.0% to 6.7%), along with a drop in C3H6 selectivity (from 22.9% to 18.6%), resulting in a reduced propylene yield from 1.6% to 1.2%. Similarly, PdO/R-TiO2 showed a more significant decline in overall performance after the reaction, with C3H8 conversion decreasing from 18.9% to 16.0%, C3H6 selectivity from 33.1% to 21.7%, and propylene yield from 6.3% to 3.5%. In contrast, for Pd/A-TiO2, the catalytic performance remained largely stable, with a minor decrease in C3H8 conversion (from 9.9% to 9.6%) and a slight improvement in selectivity (from 24.5% to 27.1%), leading to a comparable propylene yield (2.4% fresh vs. 2.6% used). Notably, Pd/R-TiO2 maintained good catalytic stability, with nearly unchanged C3H8 conversion (14.2% fresh vs. 13.7% used), selectivity (67.0% vs. 67.2%), and C3H6 yield (9.5% vs. 9.2%), confirming its robustness and potential applicability under practical conditions.
We then recorded the time course of the STY of C3H8 for PdO/R-TiO2 and Pd/R-TiO2 to test the long-term continuous stability at 500 °C (Fig. 2e and f). Rapid deactivation occurred within 8 h, resulting in a decrease of C3H6 STY from 45.7 to 20.7 kgC3H6 kgPd−1 h−1. After 30 min of regeneration in N2O flow, the activity recovered to 40.7 kgC3H6 kgPd−1 h−1, lower than the initial value of 45.7 kgC3H6 kgPd−1 h−1. This suggested that the deactivation was primarily caused by coke formation. N2O flow effectively removed these coke species, thereby restoring the catalytic activity. However, rapid deactivation was observed in the second reaction cycle, with the STY dropping to 21.1 kgC3H6 kgPd−1 h−1 within 5 h, and an even faster deactivation occurred in the third reaction cycle within only 2 h. This indicated poor stability for PdO/R-TiO2. Furthermore, the activity of PdO/R-TiO2 decreased after regeneration, which might be attributed to either irreversible sintering or the accumulation of coke.12 To verify this, we conducted TEM analysis on the PdO/R-TiO2 sample after the reaction and regeneration (Fig. S9). The results showed the presence of apparent coke deposits surrounding the PdO species, suggesting that the decline in activity during the second cycle is mainly due to incomplete coke removal rather than sintering. Moreover, no significant increase in particle size was observed, further suggesting that sintering is not the dominant cause.
In contrast, Pd/R-TiO2 retained a high C3H6 STY (∼70 kgC3H6 kgPd−1 h−1) during the stability test, showing a reliable durability compared to that of PdO/R-TiO2. Thermogravimetric analysis (TGA) and derivative thermogravimetry (DTG) experiments showed PdO/R-TiO2 exhibited approximately 1.3 wt% weight loss (Fig. S10), which could be attributed to both water desorption (<100 °C) and coke formation (200–350 °C).44 In contrast, Pd/R-TiO2 showed almost no weight loss without coke formation under working conditions.
To evaluate the long-term durability, we extended the Pd/R-TiO2 catalytic test to 24 h (Fig. S11). A decrease in C3H6 STY was observed after 14 h, likely due to coke deposition. After a 1 hour regeneration in N2O flow, the catalyst's performance was fully restored and maintained for another 10 h. These findings confirm the catalyst's regenerability and demonstrate potential for industrial application. Nevertheless, the observed deactivation highlights the demand to further optimize catalyst design based on the redox center separation strategy in future catalyst development.
In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) was conducted under C3H8/N2O co-feed conditions with a ratio of 1/1 at 500 °C on PdO/R-TiO2 and Pd/R-TiO2 in Fig. 2g and h. Bands attributed to C3H8 (2980, 2967, 2959, 2901, 2887, 2796 cm−1),45 and N2O (2235, 2213, 1299, 1271 cm−1),46 were all observed for both samples. Additionally, peaks attributed to both C3H6 (1516, 1439 cm−1) and CO2 (2359 cm−1) were observed, consistent with the products detected in the catalytic evaluation. However, the peaks of acrylate species (1644, 1301 cm−1), acetates (1565, 1460 cm−1), and formate (1592, 1376 cm−1),47 were not observed, likely ruling out the formation of these potential intermediates. By contrast, a notable peak at 1620 cm−1 appeared over PdO/R-TiO2 (Fig. 2g), which was assigned to the characteristic peak of enolic species.48,49 This indicated C3H8 was partially oxidized to enolic species, leading to low C3H6 selectivity. In contrast, no enolic species signal was observed on Pd/R-TiO2 (Fig. 2h), indicating that Pd NPs inhibited N2O-ODHP from proceeding through the enolic path. Moreover, the oxygen-containing enolic intermediate was likely to serve as a precursor for coke formation by polymerization,50 causing poor stability.
3.3. Identifying the geometric and electronic structures of Pd/TiO2
Pd/TiO2 was selected as a model catalyst for density functional theory (DFT) calculations for N2O-ODHP (Fig. 3a). The charge density of PdO on PdO/TiO2 showed positive values near TiO2 and negative values near PdO, indicating electron transfer from PdO to TiO2 (Fig. S12). For Pd NPs on PdNP/TiO2, the charge transfer direction was reversed, with electron transfer occurring from TiO2 to Pd NPs (Fig. 3b); that is, a reversed charge transfer.51 This might be one of the reasons for the improved performance on Pd/R-TiO2.
 |
| Fig. 3 (a) Optimized atomic structures of different Pd/TiO2 models for PdO/A-TiO2, Pd/A-TiO2, PdO/R-TiO2, and Pd/R-TiO2. (b) Calculated charge density difference and planar-averaged charge density difference Δρ(z) of Pd/R-TiO2. Cyan indicates holes, and yellow indicates electrons. (c) Energy diagrams of Ov formation on different sites on Pd/TiO2. (d) Possible configurations for N2O adsorption on the Pd/R-TiO2 model. Pd, mazarine; Ti, grey; O, red; N, blue. | |
N2O reduction could readily occur at Ov sites. Hence, Ov formation energy was calculated (Fig. 3c). Notably, the Ov formation on PdO on PdO/A-TiO2 was even exothermic (−1.61 eV), indicating that PdO/A-TiO2 would exhibit highly reactive Ov sites and, consequently, the highest N2O conversion. Ov formation on PdO on PdO/R-TiO2 was only 0.34 eV, smaller than those on the TiO2 surface, suggesting that the step Ov + N2O → * + N2 on PdO occurred more readily than on the TiO2 surface, whether anatase or rutile. As evident from the DFT results, the most efficient Ov formation on the TiO2 surface was observed with Pd/R-TiO2, with the lowest energy of 0.83 eV. The order of Ov formation on the TiO2 surface was Pd/R-TiO2 > Pd/A-TiO2 (1.11 eV) > PdO/A-TiO2 (2.66 eV) > PdO/R-TiO2 (3.27 eV). However, given the facile Ov formation on PdO, the overall order of Ov formation tendency was PdO/A-TiO2 > PdO/R-TiO2 > Pd/R-TiO2 > Pd/A-TiO2, in good agreement with XPS O 1s results.
To assess the ability of Pd NPs for N2O activation, various initial configurations of N2O on Pd NPs were examined,52 including side-on and end-on modes (Fig. 3d). The N2O molecule energetically preferred a side-on coordination with an adsorption energy of −0.81 eV, which was relatively weak for facilitating further N–O bond cleavage. As a result, for PdO/TiO2, both C3H8 and N2O activation occurred on the PdO species. In contrast, for PdNP/TiO2, C3H8 was dissociated on the Pd NPs and N2O reduction occurred on the Ov sites on TiO2 with the recovery of Ov, achieving the separation of redox centers.
To further support the DFT results regarding reactive oxygen species, EPR spectra were recorded for fresh and used Pd/R-TiO2 samples (Fig. 4a). Evidently, a greater number of Ov sites were formed during the N2O-ODHP with more Ti3+ species.53 In the deconvoluted O 1s spectra, lattice oxygen (Olat) appears at ∼530.0 eV and surface-adsorbed oxygen species (Oads) at ∼531.7 eV.54 For fresh samples (Fig. 4b), the Oads ratios followed the order PdO/A-TiO2 (17.1%) > PdO/R-TiO2 (16.2%) > Pd/R-TiO2 (15.8%) > Pd/A-TiO2 (11.3%). The Oads species are considered reactive intermediates that can readily participate in surface reactions, leading to the formation of Ov sites.55 Thus, the Oads ratio serves as an indicator of the oxygen vacancy formation tendency on the TiO2 surface or PdO species. To further verify the dynamics of oxygen vacancy formation, operando near-ambient-pressure X-ray photoelectron spectroscopy (NAP-XPS) measurements were conducted at 500 °C under a C3H8/N2O (1
:
1) flow. When measured under N2O-ODHP conditions, the Oads ratios changed to 9.0%, 14.4%, 10.6%, and 12.9% for PdO/A-TiO2, Pd/A-TiO2, PdO/R-TiO2, and Pd/R-TiO2, respectively (Fig. 4c). The decrease in Oads content reflects the formation of oxygen vacancies via H* species generated during C3H8 dehydrogenation. The observed Oads reduction followed the order PdO/A-TiO2 (−8.1%), PdO/R-TiO2 (−5.6%), Pd/R-TiO2 (−2.9%), Pd/A-TiO2 (+3.1%), consistent with the DFT-predicted trend in vacancy formation energies. Notably, the slight increase in Oads observed for Pd/A-TiO2 might result from the generation and accumulation of O* species derived from activated N2O, which remain strongly adsorbed on the surface.
 |
| Fig. 4 (a) EPR spectra of fresh and used Pd/R-TiO2. (b) XPS spectra of O 1s. NAP-XPS spectra of O 1s (c), Ti 2p (d), and Pd 3d (e) recorded after catalysts were exposed to C3H8/N2O = 1/1 at 500 °C. (f) H2-TPR plots for PdO/A-TiO2, Pd/A-TiO2, PdO/R-TiO2, Pd/R-TiO2. | |
We also recorded the Ti 2p XPS spectra under reaction conditions. Prior to the reaction, all catalysts exhibited a Ti3+ component of less than 10% (Fig. S13), indicating a well-preserved crystalline structure of TiO2, consistent with the XRD results (Fig. S3). After the reaction, the Ti3+ proportion significantly increased, following the order Pd/R-TiO2 (40.5%) > Pd/A-TiO2 (38.3%) > PdO/A-TiO2 (35.1%) > PdO/R-TiO2 (24.2%) (Fig. 4d). This trend aligns well with the DFT-predicted tendency for Ov formation on TiO2 surfaces.
However, since the formation of Ov requires H* species generated from propane dehydrogenation, it suggests that PdO species may directly trap these H* species via their surface oxygen, thereby suppressing H* migration to the TiO2 support and inhibiting Ov formation. In contrast, metallic Pd facilitates hydrogen spillover, promoting vacancy generation on the TiO2 surface. To further validate this hypothesis, we collected NAP-XPS spectra of Pd 3d under reaction conditions (Fig. 4e). For both PdO/TiO2 catalysts, a portion of Pd0 species was observed, indicating partial reduction of PdO during the reaction, consistent with the loss of surface oxygen. In contrast, the Pd/A-TiO2 and Pd/R-TiO2 samples maintained a fully metallic Pd0 state throughout. Taken together, these findings support the DFT predictions regarding the oxygen vacancy formation tendencies and demonstrate that hydrogen spillover from Pd to the TiO2 surface plays a crucial role in Ov dynamics during the reaction.
Further, we conducted H2-temperature programmed reduction (H2-TPR) to investigate the dynamic behavior of oxygen species and the ability of the catalysts to dissociate H2 (Fig. 4f). For all samples, a slight but noticeable negative peak was observed below 100 °C, which can be attributed to H2 adsorption. Among them, Pd/R-TiO2 exhibited the strongest H2 adsorption capacity. For both PdO/TiO2 samples, a broad H2 consumption peak appeared below 400 °C, corresponding to the reduction of PdO species. Upon further increasing the temperature above 700 °C, an additional peak was observed, which can be attributed to the removal of lattice oxygen from the TiO2 support. Notably, PdO/A-TiO2 showed a surprisingly high H2 consumption above 500 °C, which is likely associated with the anatase-to-rutile phase transformation; a process known to involve substantial hydrogen participation. This observation is consistent with the XRD results (Fig. S4). In contrast, although Pd/A-TiO2 also exhibited a high-temperature signal above 700 °C, it was much less pronounced than that of PdO/A-TiO2. This suggests that the interaction between Pd nanoparticles and anatase TiO2 may help stabilize the crystal phase against transformation. Moreover, for both Pd/A-TiO2 and Pd/R-TiO2, a reduction peak appeared around 450 °C, which can be assigned to H2 dissociation on metallic Pd NPs. This indicates that at our reaction temperature of 500 °C, the catalysts are fully capable of facilitating H*-involved processes, which are essential for bridging propane dehydrogenation and N2O activation.56
3.4. Theoretical insights into the reaction pathway
We then calculated the reaction pathways (Fig. 5). The initial C–H bond activation in C3H8 is a crucial step in the ODHP for all models, where a hydrogen atom from methyl was eliminated and adsorbed onto the catalyst surface. The activation of C3H8 showed high energy barriers of 2.65 eV and 2.22 eV for PdO/A-TiO2 and Pd/A-TiO2, respectively, which strictly impeded subsequent dehydrogenation steps. In contrast, C3H8 activation was facile for both PdO/R-TiO2 and Pd/R-TiO2 with low barrier energies of 1.40 eV and 1.21 eV, respectively.
 |
| Fig. 5 (a and b) Energy diagrams of PdO/R-TiO2 (a) and Pd/R-TiO2 (b) for C3H8 dehydrogenation through the propylene and enolic pathways. (c) Summary of the energy barriers for H* spillover on the PdO/R-TiO2 and Pd/R-TiO2 models. (d and e) Energy diagrams of PdO/A-TiO2 (c), Pd/A-TiO2 (d) for C3H8 dehydrogenation through the propylene and enolic pathways. TS, transition states. Pd, mazarine; Ti, grey; O, red; C, dark grey; H, white; H*, green. (f) Plots of the activation energy for the first dehydrogenation step of propane against e−W. | |
For PdO/R-TiO2, as shown in Fig. 5a, the energy barriers for C–C scission and C–H cleavage of C3H7* species were determined to be 1.50 eV and 2.15 eV, respectively, facilitating the formation of enolic species via the enolic pathway. Subsequently, the C2H5* species underwent dehydrogenation to C2H4*, followed by oxidation by oxygen atoms on PdO to form enolic species, as observed by in situ DRIFTS. The formation of enolic species required overcoming an energy barrier of 1.73 eV, suggesting that it was a key step for overoxidation. In contrast, for Pd/R-TiO2 in Fig. 5b, the energy barriers for C–C scission and C–H cleavage of C3H7* species were 2.64 eV and 1.86 eV, respectively. These values indicated a stronger propensity to generate C3H6 rather than enolic species, leading to higher selectivity. Besides, the dissociated H* species could directly react with the PdO or TiO2 surface after migrating from the Pd species. However, the oxygen atom within PdO on PdO/R-TiO2 strictly hindered H* migration, with a higher barrier of 0.83 eV compared to that of 0.21 eV on Pd/R-TiO2 (Fig. 5c).
To shed light on the atomic-level origin of the higher propylene selectivity of Pd/R-TiO2 compared to Pd/A-TiO2, we performed comparative DFT calculations of the ODHP reaction pathway on PdO/A-TiO2 and Pd/A-TiO2, as shown in Fig. 5d and e, respectively. In the oxidative dehydrogenation reactions, product selectivity is generally governed by two key factors: (i) suppression of undesired C–C bond cleavage, and (ii) facile desorption of the target product, propylene.21 Our results reveal that anatase TiO2 does not significantly promote C–H bond activation, as reflected by the high C3H8 activation barriers (2.55 eV and 2.22 eV). Moreover, C3H6 binds weakly on anatase-supported Pd surfaces (1.26 eV and 0.96 eV), facilitating its desorption. However, despite the weak binding, anatase-based catalysts still exhibit low C3H6 selectivity, suggesting that propylene adsorption is not the determining factor for high yield. Instead, the competition between C–H bond scission and C–C bond cleavage of the surface intermediate C3H7* is more critical. For both PdO/A-TiO2 and Pd/A-TiO2, the C–C bond cleavage pathway shows a lower barrier (3.55 eV vs. 4.05 eV, and 2.74 eV vs. 3.12 eV, respectively) than the pathway toward C3H6 formation. Therefore, it is reasonable to propose that suppressing C–C bond cleavage is key to improving propylene selectivity.
To better understand the electronic factors governing catalytic activity, we investigated the relationship between the work function (W) and the activation barrier (Ea) for C3H8 activation across four Pd/TiO2 catalysts. The calculated W values were 4.43, 4.60, 5.26, and 5.41 eV for PdO/A-TiO2, Pd/A-TiO2, PdO/R-TiO2, and Pd/R-TiO2, respectively, consistent with previous reports.57 Catalysts with larger W values are electron-deficient, which enables them to readily accept electrons from the bonding orbital of the C–H bond in C3H8, suggesting that the Pd/R-TiO2 catalyst is particularly active for C3H8 activation.58–60 This trend has been recently reported, showing that e−W can effectively quantify the ability to activate the C–H bond, surpassing the predictive capability of the d-band center.61 We also plotted Ea against e−W, which exhibited a linear relationship (Fig. 5f). This not only supplements existing findings but, together with our data, further confirms that the work function can serve as a reliable descriptor for the development of future dehydrogenation catalysts. The plot shows that as W increases, the energy barrier decreases, indicating that a higher work function facilitates C–H bond dissociation during C3H8 activation.
The combined reaction mechanisms for N2O-ODHP are summarized in Fig. 6. The designed Pd/R-TiO2, with separated redox centers, enables efficient activation of both C3H8 and N2O while minimizing direct interaction between C3 intermediates and reactive oxygen species. Hydrogen spillover bridges the oxidation and reduction reactions, thereby completing the N2O-ODHP redox cycle and effectively disentangling the trade-off between activity and selectivity, leading to boosted propylene production. In contrast, PdO/R-TiO2 fails to achieve high propylene yield as the redox centers are not spatially separated, causing both C3H8 and N2O to be activated on PdO sites. In contrast, Pd/A-TiO2 and PdO/A-TiO2 exhibit limited C3H8 conversion due to their high energy barriers resulting from their low work functions.
 |
| Fig. 6 Schematic of the reaction mechanisms of Pd/TiO2 catalysts in N2O-ODHP. | |
4. Conclusion
Pd/TiO2 model catalysts were employed to elucidate the reaction mechanisms for N2O-ODHP, which is a promising technique for propylene production and N2O valorization. Comprehensive DFT calculations combined with experimental efforts revealed that the Pd/TiO2 catalyst with redox center separation can effectively avoid competition and direct contact between N2O and C3H8, where Ov sites on TiO2 dissociate N2O and the metallic Pd sites activate C3H8, thereby disentangling the trade-off between activity (C3H8 conversion) and C3H6 selectivity. The separated redox reactions are linked by the hydrogen spillover from the Pd sites to the TiO2 surface. This work clarifies the atomic-level mechanisms and catalyst design behind an innovative approach to valorizing high-concentration N2O emissions from industrial sources. Building on our current work, future efforts should prioritize the development of advanced catalysts for N2O valorization and the broader implementation of the redox center separation strategy.
Author contributions
Yunshuo Wu: conceptualization, methodology, formal analysis, writing – original draft. Yuxin Sun: methodology. Xuanhao Wu: formal analysis. Haiqiang Wang: conceptualization, writing – review & editing, project administration. Lianzhou Wang: project administration, writing – review & editing. Zhongbiao Wu: project administration, writing – review & editing.
Conflicts of interest
The authors declare no competing financial interest.
Data availability
The data supporting this article have been included as part of the SI.
The SI file contains detailed methods descriptions of catalysts preparation and theoretical calculations. It also includes TEM, STEM and EDX mapping images, N2 adsorption-desorption isotherms, XRD and Raman spectra, TGA and DTG curves, XPS data, DFT results, ODHP performance data, and a comparative table of N2O-ODHP performance against previously reported catalysts. See DOI: https://doi.org/10.1039/d5ta03582a.
Acknowledgements
The authors gratefully acknowledge the financial support of this work from the National Natural Science Foundation of China (NSFC-524B2140, NSFC-52370120), the Zhejiang Provincial “Lead Wild goose” Research and Development Project (No. 2024C03114), and the Program for Zhejiang Leading Team of S&T Innovation (Grant No. 2013TD07).
References
- M. T. McCulloch, A. Winter, C. E. Sherman and J. A. Trotter, 300 years of sclerosponge thermometry shows global warming has exceeded 1.5 °C, Nat. Clim. Change, 2024, 14(2), 171–177 CrossRef
.
- D. I. Armstrong McKay, A. Staal, J. F. Abrams, R. Winkelmann, B. Sakschewski, S. Loriani, I. Fetzer, S. E. Cornell, J. Rockström and T. M. Lenton, Exceeding 1.5°C global warming could trigger multiple climate tipping points, Science, 2022, 377(6611), eabn7950 CrossRef PubMed
.
- H. Tian, R. Xu, J. G. Canadell, R. L. Thompson, W. Winiwarter, P. Suntharalingam, E. A. Davidson, P. Ciais, R. B. Jackson, G. Janssens-Maenhout, M. J. Prather, P. Regnier, N. Pan, S. Pan, G. P. Peters, H. Shi, F. N. Tubiello, S. Zaehle, F. Zhou, A. Arneth, G. Battaglia, S. Berthet, L. Bopp, A. F. Bouwman, E. T. Buitenhuis, J. Chang, M. P. Chipperfield, S. R. S. Dangal, E. Dlugokencky, J. W. Elkins, B. D. Eyre, B. Fu, B. Hall, A. Ito, F. Joos, P. B. Krummel, A. Landolfi, G. G. Laruelle, R. Lauerwald, W. Li, S. Lienert, T. Maavara, M. MacLeod, D. B. Millet, S. Olin, P. K. Patra, R. G. Prinn, P. A. Raymond, D. J. Ruiz, G. R. van der Werf, N. Vuichard, J. Wang, R. F. Weiss, K. C. Wells, C. Wilson, J. Yang and Y. Yao, A comprehensive quantification of global nitrous oxide sources and sinks, Nature, 2020, 586(7828), 248–256 CrossRef PubMed
.
- X. Wu, J. Du, Y. Gao, H. Wang, C. Zhang, R. Zhang, H. He, G. Lu and Z. Wu, Progress and challenges in nitrous oxide decomposition and valorization, Chem. Soc. Rev., 2024, 53(16), 8379–8423 RSC
.
- M. Konsolakis, Recent Advances on Nitrous Oxide (N2O) Decomposition over Non-Noble-Metal Oxide Catalysts: Catalytic Performance, Mechanistic Considerations, and Surface Chemistry Aspects, ACS Catal., 2015, 5(11), 6397–6421 CrossRef
.
- A. R. Ravishankara, J. S. Daniel and R. W. Portmann, Nitrous Oxide (N2O): The Dominant Ozone-Depleting Substance Emitted in the 21st Century, Science, 2009, 326(5949), 123–125 CrossRef PubMed
.
- H. Liu, S. Yang, G. Wang, H. Liu, Y. Peng, C. Sun, J. Li and J. Chen, Strong Electronic Orbit Coupling between Cobalt and Single-Atom Praseodymium for Boosted Nitrous Oxide Decomposition on Co3O4 Catalyst, Environ. Sci. Technol., 2022, 56(22), 16325–16335 CrossRef PubMed
.
- Y. Wu, X. Wu, J. Fan, H. Wang and Z. Wu, Insights into the Roles of Different Iron Species on Zeolites for N2O Selective Catalytic Reduction by CO, Environ. Sci. Technol., 2024, 58(51), 22583–22593 CrossRef PubMed
.
- L. Li, J. Xu, J. Hu and J. Han, Reducing Nitrous Oxide Emissions to Mitigate Climate Change and Protect the Ozone Layer, Environ. Sci. Technol., 2014, 48(9), 5290–5297 CrossRef CAS PubMed
.
- E. A. Davidson and W. Winiwarter, Urgent abatement of industrial sources of nitrous oxide, Nat. Clim. Change, 2023, 13(7), 599–601 CrossRef
.
- F. Le Vaillant, A. Mateos Calbet, S. González-Pelayo, E. J. Reijerse, S. Ni, J. Busch and J. Cornella, Catalytic synthesis of phenols with nitrous oxide, Nature, 2022, 604(7907), 677–683 CrossRef CAS PubMed
.
- S. Chen, X. Chang, G. Sun, T. Zhang, Y. Xu, Y. Wang, C. Pei and J. Gong, Propane dehydrogenation: catalyst development, new chemistry, and emerging technologies, Chem. Soc. Rev., 2021, 50(5), 3315–3354 RSC
.
- J. T. Grant, C. A. Carrero, F. Goeltl, J. Venegas, P. Mueller, S. P. Burt, S. E. Specht, W. P. McDermott, A. Chieregato and I. Hermans, Selective oxidative dehydrogenation of propane to propene using boron nitride catalysts, Science, 2016, 354(6319), 1570–1573 CrossRef CAS
.
- D. Zhao, X. Tian, D. E. Doronkin, S. Han, V. A. Kondratenko, J.-D. Grunwaldt, A. Perechodjuk, T. H. Vuong, J. Rabeah, R. Eckelt, U. Rodemerck, D. Linke, G. Jiang, H. Jiao and E. V. Kondratenko, In situ formation of ZnOx species for efficient propane dehydrogenation, Nature, 2021, 599(7884), 234–238 CrossRef CAS PubMed
.
- W. Wang, S. Chen, C. Pei, R. Luo, J. Sun, H. Song, G. Sun, X. Wang, Z.-J. Zhao and J. Gong, Tandem propane dehydrogenation and surface oxidation catalysts for selective propylene synthesis, Science, 2023, 381(6660), 886–890 CrossRef CAS
.
- R. Almallahi, J. Wortman and S. Linic, Overcoming limitations in propane dehydrogenation by codesigning catalyst-membrane systems, Science, 2024, 383(6689), 1325–1331 CrossRef CAS
.
- H. Zhou, H. Li, L. Wang, S. Chu, L. Liu, L. Liu, J. Qi, Z. Ren, A. Cai, Y. Hui, Y. Qin, L. Song, X. Qin, J. Shi, J. Hou, Y. Ding, J. Ma, S. Xu, X. Tao, L. Li, Q. Yang, B. Hu, X. Liu, L. Chen, J. Xiao and F.-S. Xiao, Cobaltosilicate zeolite beyond platinum catalysts for propane dehydrogenation, Nat. Catal., 2025, 8, 357–367 CrossRef CAS
.
- S. Chen, Y. Xu, X. Chang, Y. Pan, G. Sun, X. Wang, D. Fu, C. Pei, Z.-J. Zhao, D. Su and J. Gong, Defective TiOx overlayers catalyze propane dehydrogenation promoted by base metals, Science, 2024, 385(6706), 295–300 CrossRef CAS PubMed
.
- Y. Wang, J. Wang, P. Zheng, C. Sun, J. Luo and X. Xie, Boosting selectivity and stability on Pt/BN catalysts for propane dehydrogenation via calcination & reduction-mediated strong metal-support interaction, J. Energy Chem., 2022, 67, 451–457 CrossRef CAS
.
- G. M. Beshara, I. Surin, M. Agrachev, H. Eliasson, T. Otroshchenko, F. Krumeich, R. Erni, E. V. Kondratenko and J. Pérez-Ramírez, Mechanochemically-derived iron atoms on defective boron nitride for stable propylene production, EES Catal., 2024, 2(6), 1263–1276 RSC
.
- X. Jiang, L. Sharma, V. Fung, S. J. Park, C. W. Jones, B. G. Sumpter, J. Baltrusaitis and Z. Wu, Oxidative Dehydrogenation of Propane to Propylene with Soft Oxidants via Heterogeneous Catalysis, ACS Catal., 2021, 11(4), 2182–2234 CrossRef CAS
.
- J. H. Carter, T. Bere, J. R. Pitchers, D. G. Hewes, B. D. Vandegehuchte, C. J. Kiely, S. H. Taylor and G. J. Hutchings, Direct and oxidative dehydrogenation of propane: from catalyst design to industrial application, Green Chem., 2021, 23(24), 9747–9799 RSC
.
- Y. Wu, X. Wu, H. Wang and Z. Wu, Advances and challenges in N2O valorization for alkane oxidative dehydrogenation to olefins, Green Chem. Eng., 2024 DOI:10.1016/j.gce.2024.1011.1005
.
- J. Zhou, Z. Gao, G. Xiang, T. Zhai, Z. Liu, W. Zhao, X. Liang and L. Wang, Interfacial compatibility critically controls Ru/TiO2 metal-support interaction modes in CO2 hydrogenation, Nat. Commun., 2022, 13(1), 327 CrossRef CAS
.
- Z. Tang, Y. Li, K. Zhang, X. Wang, S. Wang, Y. Sun, H. Zhang, S. Li, J. Wang, X. Gao, Z. Hou, L. Shi, Z. Yuan, K. Nie, J. Xie, Z. Yang and Y.-M. Yan, Interfacial Hydrogen Spillover on Pd-TiO2 with Oxygen Vacancies Promotes Formate Electrooxidation, ACS Energy Lett., 2023, 8(9), 3945–3954 CrossRef CAS
.
- H. Cheng and A. Selloni, Surface and subsurface oxygen vacancies in anatase TiO2 and differences with rutile, Phys. Rev. B:Condens. Matter Mater. Phys., 2009, 79(9), 092101 CrossRef
.
- J. Wang, H. Chen, Z. Hu, M. Yao and Y. Li, A Review on the Pd-Based Three-Way Catalyst, Catal. Rev., 2015, 57(1), 79–144 CrossRef CAS
.
- E. Araujo-Lopez, L. Joos, B. D. Vandegehuchte, D. I. Sharapa and F. Studt, Theoretical Investigations of (Oxidative) Dehydrogenation of Propane to Propylene over Palladium Surfaces, J. Phys. Chem. C, 2020, 124(5), 3171–3176 CrossRef CAS
.
- J. J. H. B. Sattler, J. Ruiz-Martinez, E. Santillan-Jimenez and B. M. Weckhuysen, Catalytic Dehydrogenation of Light Alkanes on Metals and Metal Oxides, Chem. Rev., 2014, 114(20), 10613–10653 CrossRef CAS PubMed
.
- X. Huang, O. Akdim, M. Douthwaite, K. Wang, L. Zhao, R. J. Lewis, S. Pattisson, I. T. Daniel, P. J. Miedziak, G. Shaw, D. J. Morgan, S. M. Althahban, T. E. Davies, Q. He, F. Wang, J. Fu, D. Bethell, S. McIntosh, C. J. Kiely and G. J. Hutchings, Au–Pd separation enhances bimetallic catalysis of alcohol oxidation, Nature, 2022, 603(7900), 271–275 CrossRef PubMed
.
- Y. Sun, Y. Wu, Y. Bai, X. Wu, H. Wang and Z. Wu, High performance iridium loaded on natural halloysite nanotubes for CO-SCR reaction, Fuel, 2024, 357, 129938 CrossRef
.
- G. Kresse and J. Furthmuller, Efficiency of ab initio total energy calculations for metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci., 1996, 6(1), 15–50 Search PubMed
.
- G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B:Condens. Matter Mater. Phys., 1996, 54(16), 11169–11186 Search PubMed
.
- J. Chen, S. Xiong, H. Liu, J. Shi, J. Mi, H. Liu, Z. Gong, L. Oliviero, F. Maugé and J. Li, Reverse oxygen spillover triggered by CO adsorption on Sn-doped Pt/TiO2 for low-temperature CO oxidation, Nat. Commun., 2023, 14(1), 3477 CrossRef
.
- J. Chen, Y. Su, Q. Meng, H. Qian, L. Shi, J. A. Darr, Z. Wu and X. Weng, Palladium Encapsulated by an Oxygen-Saturated TiO2 Overlayer for Low-Temperature SO2-Tolerant Catalysis during CO Oxidation, Angew. Chem., Int. Ed., 2023, 62(49), e202310191 CrossRef PubMed
.
- L. Kuai, Z. Chen, S. Liu, E. Kan, N. Yu, Y. Ren, C. Fang, X. Li, Y. Li and B. Geng, Titania supported synergistic palladium single atoms and nanoparticles for room temperature ketone and aldehydes hydrogenation, Nat. Commun., 2020, 11(1), 48 CrossRef PubMed
.
- Z. Sheng, Z. Wu, Y. Liu and H. Wang, Gas-phase photocatalytic oxidation of NO over palladium modified TiO2 catalysts, Catal. Commun., 2008, 9(9), 1941–1944 CrossRef
.
- C. Wu, Q. Liu, Y. Zhan, W. Tan, X. Wei, Q. Tong, H. Wan and L. Dong, Insights into the surface Structure-Sensitive photocatalytic oxidation of gaseous toluene on Pd/TiO2 catalysts, Chem. Eng. J., 2023, 475, 146294 CrossRef
.
- V. Muravev, G. Spezzati, Y.-Q. Su, A. Parastaev, F.-K. Chiang, A. Longo, C. Escudero, N. Kosinov and E. J. M. Hensen, Interface dynamics of Pd–CeO2 single-atom catalysts during CO oxidation, Nat. Catal., 2021, 4(6), 469–478 CrossRef
.
- Y. Zhang, C. X. Harris, P. Wallenmeyer, J. Murowchick and X. Chen, Asymmetric Lattice Vibrational Characteristics of Rutile TiO2 as Revealed by Laser Power Dependent Raman Spectroscopy, J. Phys. Chem. C, 2013, 117(45), 24015–24022 CrossRef CAS
.
- E. V. Kondratenko and M. Baerns, Catalytic oxidative dehydrogenation of propane in the presence of O2 and N2O—the role of vanadia distribution and oxidant activation, Appl. Catal., A, 2001, 222(1), 133–143 CrossRef CAS
.
- E. V. Kondratenko, M. Cherian, M. Baerns, D. Su, R. Schlögl, X. Wang and I. E. Wachs, Oxidative dehydrogenation of propane over V/MCM-41 catalysts: comparison of O2 and N2O as oxidants, J. Catal., 2005, 234(1), 131–142 CrossRef CAS
.
- A. Ates, C. Hardacre and A. Goguet, Oxidative dehydrogenation of propane with N2O over Fe-ZSM-5 and Fe–SiO2: Influence of the iron species and acid sites, Appl. Catal., A, 2012, 441–442, 30–41 CrossRef CAS
.
- B. Wang and G. Manos, A novel thermogravimetric method for coke precursor characterisation, J. Catal., 2007, 250(1), 121–127 CrossRef CAS
.
- Z. Ren, Z. Wu, W. Song, W. Xiao, Y. Guo, J. Ding, S. L. Suib and P.-X. Gao, Low temperature propane oxidation over Co3O4 based nano-array catalysts: Ni dopant effect, reaction mechanism and structural stability, Appl. Catal., B, 2016, 180, 150–160 CrossRef CAS
.
- M. L. Bols, B. E. R. Snyder, H. M. Rhoda, P. Cnudde, G. Fayad, R. A. Schoonheydt, V. Van Speybroeck, E. I. Solomon and B. F. Sels, Coordination and activation of nitrous oxide by iron zeolites, Nat. Catal., 2021, 4(4), 332–340 CrossRef CAS
.
- Z. Liu, Y. An, G. Xu, Y. Yu and H. He, Insight into the Promotion Effect of Trace Pd Doping on the Catalytic Performance of Ag/Al2O3 for C3H6-SCR of NOx, Environ. Sci. Technol., 2023, 57(39), 14760–14767 CrossRef CAS
.
- Q. Wu, H. He and Y. Yu, In situ DRIFTS study of the selective reduction of NOx with alcohols over Ag/Al2O3 catalyst: Role of surface enolic species, Appl. Catal., B, 2005, 61(1), 107–113 CrossRef CAS
.
- Y. Yu, H. He and Q. Feng, Novel Enolic Surface Species Formed during Partial Oxidation of CH3CHO, C2H5OH, and C3H6 on Ag/Al2O3: An in Situ DRIFTS Study, J. Phys. Chem. B, 2003, 107(47), 13090–13092 CrossRef CAS
.
- E. T. C. Vogt, D. Fu and B. M. Weckhuysen, Carbon Deposit Analysis in Catalyst Deactivation, Regeneration, and Rejuvenation, Angew. Chem., Int. Ed., 2023, 62(29), e202300319 CrossRef CAS
.
- Y. Wu, Y. Sun, X. Wu, H. Wang and Z. Wu, Reversed Charge Transfer Enables Dual Active Sites on Ir/hBN for Synergistic N2O Valorization and Propane Selective Oxidation, ACS Catal., 2024, 14(17), 13520–13530 CrossRef CAS
.
- D. Wu, K. Chen, P. Lv, Z. Ma, K. Chu and D. Ma, Direct Eight-Electron N2O Electroreduction to NH3 Enabled by an Fe Double-Atom Catalyst, Nano Lett., 2024, 24(28), 8502–8509 CrossRef CAS PubMed
.
- Y. Gao, H. Xu, K. Yin, Z. Li, J. Fan, X. Wu and Z. Wu, Modulating Charge Transfer for Photo-driven N2O Reduction via Electronegativity Differences Between Cu and Support, Appl. Catal., B, 2025, 125528 CrossRef CAS
.
- Y. Wu, H. Wang and Z. Wu, Synthesis of a Cu@ZnO1−x/Al2O3 catalyst with high-density ZnO1−x–Cu interfacial sites for enhanced CO2 hydrogenation to methanol, J. Mater. Chem. A, 2024, 12(44), 30692–30706 RSC
.
- H. Zhang, C. Han, C. Li, P. Wang, H. Huang, S. Wang and J. Li, Design of Cu/ZnO/Al2O3 catalysts with a rich Cu–ZnO interface for enhanced CO2 hydrogenation to methanol using zinc-malachite as the precursor, New J. Chem., 2023, 47(12), 5885–5893 RSC
.
- Y. Wu, H. Wang, F.-S. Xiao and Z. Wu, Gadolinium-Mediated Oxygen Affinity Induced Efficient Covalorization of CH4 and N2O in an Ir–Gd2O3 Single-Atom Catalyst, J. Am. Chem. Soc., 2025, 147(29), 25787–25798 CrossRef PubMed
.
- H. Lee, M. Shin, M. Lee and Y. J. Hwang, Photo-oxidation activities on Pd-doped TiO2 nanoparticles: critical PdO formation effect, Appl. Catal., B, 2015, 165, 20–26 CrossRef CAS
.
- M. T. Greiner, L. Chai, M. G. Helander, W.-M. Tang and Z.-H. Lu, Transition Metal Oxide Work Functions: The Influence of Cation Oxidation State and Oxygen Vacancies, Adv. Funct. Mater., 2012, 22(21), 4557–4568 CrossRef CAS
.
- A. Kahn, Fermi level, work function and vacuum level, Mater. Horiz., 2016, 3(1), 7–10 RSC
.
- Z. Wang, C. Wang, S. Mao, B. Lu, Y. Chen, X. Zhang, Z. Chen and Y. Wang, Decoupling the electronic and geometric effects of Pt catalysts in selective hydrogenation reaction, Nat. Commun., 2022, 13(1), 3561 CrossRef CAS
.
- X. Chang, Z. Lu, X. Wang, Z.-J. Zhao and J. Gong, Tracking C–H bond activation for propane dehydrogenation over transition metal catalysts: work function shines, Chem. Sci., 2023, 14(23), 6414–6419 RSC
.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.