Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Tuning the oxidative dehydrogenation of propane mechanism by Pd–B/Al2O3 bifunctional catalysis through suppression of gas-phase radicals and enhancement of surface-mediated pathways

Chunyan Maabc, Cheng Chenac, Zhenhao Houac, Zilin Yanac, Fengbang Wangac, Lei Biac, Maoyong Song*abc and Guibin Jiangbc
aKey Laboratory of Environmental Nanotechnology and Health Effects, Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China. E-mail: smsong@rcees.ac.cn; Fax: +86 10 62918178; Tel: +86 10 62918178
bState Key Laboratory of Environmental Chemistry and Ecotoxicology, Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China
cUniversity of Chinese Academy of Sciences, Beijing 100049, China

Received 14th August 2025 , Accepted 9th October 2025

First published on 10th October 2025


Abstract

Boron-based catalysts typically promote the oxidative dehydrogenation of propane through a gas-phase radical mechanism, achieving high propylene selectivity but limited propane conversion. In this study, palladium is incorporated into B/Al2O3 to construct a bifunctional Pd–B/Al2O3 catalyst that shifts the reaction pathway from a radical-dominated route to a surface-catalyzed Langmuir–Hinshelwood mechanism. This cooperative effect increases propane conversion while maintaining the propylene selectivity. Quantitative analysis of both gaseous and surface H2O2 indicates a likely shift in the reaction mechanism from a gas-phase radical pathway to a surface-catalyzed process. Combined in situ EPR, DRIFT, and XPS analyses, along with DFT calculations, reveal that Pd sites promote propane adsorption and substantially lower the dehydrogenation barrier of C3H7, while adjacent BOx(OH)3−x species selectively oxidize hydrogen to H2O, mitigating over-oxidation. The presence of surface B–OH groups further improves performance, increasing propylene selectivity by approximately 5% under humidified conditions. These findings highlight a new strategy for designing efficient ODHP catalysts by harnessing bifunctional active sites to promote surface-mediated reaction pathways.


Introduction

Boron-based catalysts including boron nitride (BN),1–9 boron carbide (B4C),10 boron oxide,11,12 and supported boron materials13–15 have attracted considerable attention for propane oxidative dehydrogenation (ODHP) due to their exceptional propylene selectivity. This performance originates from surface oxygen-functionalized boron species that initiate the reaction by abstracting hydrogen atoms from propane,1–16 leading to two distinct mechanistic pathways. One is the well-established gas-phase radical pathway,1–9 and the other is a surface-catalyzed pathway.13,14

In the radical mechanism, gas-phase C3H7· radicals undergo dehydrogenation to form propylene and simultaneously generate gaseous H2O2 as a characteristic byproduct due to radical reactions.6,7,12,16 In contrast, the surface-catalyzed pathway involves chemisorbed intermediates in which surface-bound C3H7 species are dehydrogenated through interactions with lattice oxygen. In this process, hydrogen is transferred to surface OH or O sites, resulting in the formation of H2O as the definitive byproduct.13,14 Since direct detection of C3H7· radicals remains experimentally challenging, the concentration of gaseous H2O2 serves as a key indicator for distinguishing between the two pathways.

Although boron-based catalysts offer high propylene selectivity through radical pathway mechanisms, in which surface boron sites suppress alkoxyl radicals that lead to deep-oxidation products, they show inherently limited efficiency for propane dehydrogenation.1–16 Surface-catalyzed ODHP mechanisms are less prevalent than radical pathways in pure boron systems due to the absence of cooperative interactions between distinct active sites. Successful ODHP requires both dehydrogenation sites for activating propane and nearby oxidation sites that convert hydrogen selectively into water. This cooperation helps drive the reaction toward propylene formation by thermodynamically removing hydrogen from the equilibrium.17,18 Noble metals such as palladium and platinum demonstrate excellent dehydrogenation performance.19–22 However, when used alone, they often favor unselective oxidation pathways. This drawback arises from the lack of specific sites for hydrogen oxidation, which leads to over-oxidation and combustion reactions instead of the desired selective dehydrogenation.

To address these limitations, we incorporated both palladium and boron onto an Al2O3 support to construct dual functional sites. In this configuration, palladium centers are responsible for propane dehydrogenation, while neighboring BOx(OH)3−x species selectively oxidize hydrogen to H2O. This structure satisfies the fundamental requirement for cooperative site functionality in ODHP. Catalytic performance studies revealed a significant increase in propane conversion by 22% without any loss in propylene selectivity. Colorimetric analysis of H2O2 and in situ spectroscopy further confirmed a mechanistic shift from gas-phase radical reactions to surface-mediated catalysis. This study presents a new strategy for the rational design of selective alkane dehydrogenation catalysts based on boron systems.

Experimental

Catalyst synthesis

The supported boron on aluminum oxide (B/Al2O3) was prepared using an impregnation method. Boric acid (H3BO3, AR, ≥99.5%, Macklin) was introduced to Al2O3 (AEROXIDE), followed by the addition of ultrapure water at a ratio of 50 ml H2O per gram of Al2O3. The mixture was stirred vigorously at 50 °C until the solution was completely evaporated. The resulting product was then dried at 70 °C and calcined at 650 °C for 4 hours. These products were designated as YB/Al2O3, with Y representing the weight percentage of boron.

The supported Pd–B/Al2O3 catalysts were prepared using a wetness impregnation method. An aqueous solution of palladium nitrate (Pd(NO3)2·2H2O, Macklin) was introduced to B/Al2O3, followed by drying at 80 °C overnight and subsequent calcination at 225 °C for 3 hours. The catalysts were designated as XPd–YB/Al2O3, where X and Y represent the weight percentages of palladium and boron, respectively. 3Au–1B/Al2O3 and 3Pt–1B/Al2O3 were synthesized using the same procedure as 3Pd–1B/Al2O3, with HAuCl4·3H2O and H2PtCl6·6H2O as the respective metal precursors.

Catalyst characterization

Before undergoing XRD, STEM, XPS, NMR, EPR, Raman and DRIFTS characterization, the sample was pretreated using a four-step reduction process. The process involved the following steps: step 1: reduction at 150 °C for 30 min with a temperature ramping rate of 4 °C min−1. Step 2: reduction at 200 °C for 30 min with a temperature ramping rate of 1 °C min−1. Step 3: reduction at 250 °C for 30 min with a temperature ramping rate of 1 °C min−1. Step 4: reduction at 490 °C for 30 min with a temperature ramping rate of 5 °C min−1 in a 10% H2/N2 balance. Prior to XRD, STEM, XPS and NMR characterization studies, the reduction catalysts were kept under O2-free atmosphere.

X-ray diffraction (XRD) was performed using a PANalytical X'Pert3 Powder diffractometer with Cu Kα radiation (k = 0.15406 nm). The catalyst powder was placed on a silicon slice sample holder for testing. Samples for high-resolution transmission electron microscopy (HRTEM) were prepared by suspending catalyst powder in ethanol, followed by sonication and deposition on a Cu grid with a carbon film. HRTEM images were acquired using a JEM 2100F electron microscope operating at an acceleration voltage of 200 kV, equipped with Oxford energy-dispersive X-ray spectroscopy (EDS). X-ray photoelectron spectroscopy (XPS) and in situ XPS analysis of ODHP over catalysts were performed using an ESCALAB 250Xi (ThermoFisher) equipped with an Al Kα source. Detailed acquisition of individual elemental regions was performed at 40 eV. H2-temperature-programmed reduction (TPR) was performed using a BSD-Chem C200 Automatic Chemisorption Analyzer. Prior to analysis, a 0.20 g sample was pretreated under He flow at 120 °C for 60 min. Subsequently, the system was cooled to 50 °C and subjected to reduction in 10% H2/Ar with a temperature ramp of 10 °C min−1 to 900 °C, monitored by thermal conductivity detection (TCD). 11B MAS nuclear magnetic resonance (NMR) performed on a JEOL JNM-ECZ600R spectrometer equipped with a 3.2 mm MAS probe operating at 192 MHz. The chemical shift was referenced to boric acid aqueous solution (1 mol L−1), which has a chemical shift of 19.6 ppm. The 11B direct polarization MAS NMR was conducted with a spinning rate of 12 kHz, 1210 scans, and a pulse delay of 3 s. 1H MAS NMR conducted at the same spectrometer (JEOL JNM-ECZ600R) with a 3.2 mm MAS probe operating at 600 MHz. Hexafluoroisopropanol was used as the standard sample for 1H MAS NMR quantification. The 1H direct polarization MAS NMR was carried out with a spinning rate of 12 kHz, 88 scans, and a pulse delay of 5 s. Electron paramagnetic resonance (EPR) experiments using 5,5-dimethyl-1-pyrroline N-oxide (DMPO) as a spin-trapping were performed on a JEOL-FA200 spectrometer at 123 K. Samples of Pd/Al2O3, B/Al2O3 or Pd–B/Al2O3 (30 mg) were collected after oxidative dehydrogenation of propane at 490 °C for 10 minutes. The samples were quickly placed in a quartz cell, followed by the addition of 0.2 mL of DMPO (50 mg mL−1). These samples were thawed and subjected to EPR measurements. Raman spectra were collected using a Renishaw inVia Raman microscope with an excitation wavelength of 532 nm, covering the range of 200–4000 cm−1 with a 20 s integration time. Catalyst powder was placed in an in situ cell reactor and pretreated using the aforementioned four-step reduction process. After the pretreatment, Raman spectra were recorded. Then a gas mixture of 6.2% C3H8/3.1% O2 with N2 balance was introduced, and Raman spectra were recorded after 5 hours of reaction. All in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) spectra were acquired using a Bruker INVENIO R instrument equipped with a DRIFTS cell and a liquid nitrogen-cooled LN-MCT detector. The spectra were collected by accumulating 256 scans over the range of 600 to 4000 cm−1 at a resolution of 4 cm−1. The catalyst underwent a four-step reduction process before DRIFTS measurements. After the pre-treatment, the temperature was maintained at 490 °C under a 10% H2/N2 balance, and a baseline spectrum was recorded. After recording the baseline, the gas composition was switched to either 6.2% C3H8/3.1% O2 with N2 balance, or the same mixture with N2 balance carrying saturated H2O or D2O vapor at 25 °C. Spectra were continuously recorded for 20 minutes to capture the ODHP spectrum. FT-IR spectra of adsorbed CO were obtained with the same Bruker INVENIO R instrument and liquid nitrogen-cooled LN-MCT detector. The catalyst was pre-treated using the same four-step reduction process. After pre-treatment, the catalyst was cooled down to 25 °C and exposed to nitrogen for 2 hours to remove physical adsorbed oxygen, followed by recording a baseline spectrum. Afterward, a pulse of CO was introduced, and the spectra were collected by accumulating 256 scans over the range of 600 to 4000 cm−1 at a resolution of 4 cm−1.

Catalytic testing

The reaction was conducted in a quartz-tube reactor with an internal diameter of 10 mm. The catalyst (100 mg) was diluted with quartz sand (100 mg) and then loaded into the reactor. The catalysts were subjected to a four-step reduction process prior to the reaction. The temperature was maintained at 490 °C, and the catalyst was exposed to the reaction mixture. The total flow rate of the reaction mixture was 70 mL min−1, comprising 4.4 mL of C3H8 and 2.2 mL of O2 with either dry N2 balance or N2 balance carried saturated water vapor at 25 °C. The inverse weight hour-space velocity (WHSV−1) was 0.29 gC3H8−1 gcatalyst h, and the weight hour-space velocity (WHSV) was 1800 LC3H8 kgcatalyst−1 h−1. To assess the effects of H2O, the catalyst was physically mixed with a desiccant (macroporous silica gel). The desiccant activity was evaluated using quartz sand and a desiccant that was physically mixed with quartz sand. The CO and CO2 products were analyzed using a gas chromatograph (Agilent 7890 B) equipped with a nickel furnace and an FID detector. C3H8, C3H6, C2H6, C2H4, and CH4 were analyzed using an Agilent 6890N equipped with an FID detector. Propane conversion and propylene selectivity were determined by the equations below:
image file: d5cy00990a-t1.tif

image file: d5cy00990a-t2.tif
Here, Yi represents the mole fraction of product i in the outlet feed gas.

Quantitative characterization of H2O2

Preparation of H2O2 standards. The color reaction of potassium titanium oxalate (PTO dihydrate, 99.99%, Sigma-Aldrich, USA) with H2O2 was used to quantitatively characterize the concentration of H2O2. H2O2 solutions were prepared with low H2O2 concentrations (0, 0.02, 0.05 and 0.07 mg L−1) and high concentrations (0.1, 0.5, 1, and 2 mg L−1) using an H2O2 standard solution. Each H2O2 solution was mixed with a 0.1 M PTO solution in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 volume ratio. The mixed solutions (100 μL per well) were analyzed using a Microplate Reader at 400 nm to determine absorbance. A standard curve was created for low and high concentrations of H2O2 based on absorbance measurements.
H2O2 measurement from reaction gas. The catalyst (100 mg) was placed in a quartz-tube reactor with an internal diameter of 10 mm. The catalysts were pretreated using a four-step reduction process. A reaction mixture of 3% C3H8/1.5% O2/N2 was passed through the reactor at a total flow rate of 70 mL min−1 and 490 °C for 1 h. After the reaction, 1 mL of ultrapure water (18.2 MΩ cm, Rephile) was used to collect and dissolve liquid drops containing gaseous H2O2 from the condenser's quartz tube. The resulting solution was mixed with PTO and analyzed as described above.
H2O2 measurement on the catalyst surface. The catalyst (100 mg) was removed from the quartz tube reactor and dispersed in 1 mL of ultrapure water. The suspension was then centrifuged at 10[thin space (1/6-em)]000 rpm for 10 min. The upper layer was collected for H2O2 quantification using the PTO method as described previously.

DFT calculation details

Density functional theory (DFT) calculations were performed using the projector-augmented plane-wave method in the Vienna ab initio simulation package. A generalized gradient approximation using the Perdew–Burke–Ernzerhof functional was used. The cutoff energy for the plane wave was set to 480 eV. The energy criterion was set to 10−4 eV in the iterative solution of the Kohn–Sham equation. All structures were relaxed until the residual forces on the atoms were less than 0.05 eV Å−1. A distance of 20 Å was used in the vertical direction to prevent interactions between the periodic units. A Monkhorst–Pack scheme with a 2 × 3 × 1 k-point mesh was used. The nudge elastic band (CI-NEB) method was used to explore the diffusion barrier of adsorbed hydrogen at different adsorption sites.

Results and discussion

The catalytic performances of ODHP

The catalytic performances of the 1B/Al2O3, 3Pd/Al2O3, XPd–1B/Al2O3, and 3Pd–YB/Al2O3 catalysts (where X and Y represent the weight percentages of Pd and B, respectively) were compared after 1 h of reaction to assess their efficiency in propane conversion and propylene selectivity. As shown in Fig. 1a, the 3Pd/Al2O3 catalyst exhibit the highest propane conversion of 26.7%, but only 5.9% selectivity towards propylene. Most propane was completely oxidized to CO2 over 3Pd/Al2O3. In contrast, 1B/Al2O3 achieved a propane conversion of 15.9% with a significantly higher propylene selectivity of 54.6%. Compared to 1B/Al2O3, Au–B/Al2O3 displayed inferior ODHP activity, while Pt–B/Al2O3 initially showed higher activity but suffered from poor stability, with approximately 54% loss in propane conversion after 300 min of reaction (Fig. S1 and S2). In contrast, the Pd–B/Al2O3 catalysts exhibited similar propylene selectivity, but notably increased propane conversion, suggesting that palladium doping enhanced propane conversion while maintaining high propylene selectivity. The performance of the Pd–B/Al2O3 catalysts was influenced by the atomic ratio of Pd to B. Among the XPd–1B/Al2O3 catalysts, 3Pd–1B/Al2O3 exhibit the highest propane conversion of 19.3% and propylene selectivity of 54.6% after 1 h (Fig. S3 and S4). In the 3Pd–YB/Al2O3 series, 3Pd–0.5B/Al2O3 exhibit a higher propane conversion of 22.1%, but a slightly lower propylene selectivity of 48.3% than that of 3Pd–1B/Al2O3 (Fig. S5 and S6). Consequently, 3Pd–1B/Al2O3 exhibited particularly high effectiveness for ODHP, attributed to its excellent propylene selectivity. As reported, H–BN achieved 79% selectivity to propylene at 14% propane conversion,1 while the borosilicate MFI framework (BS-1) demonstrated approximately 10% propane conversion with ∼55% propylene selectivity at 490 °C,14 highlighting the strong catalytic performance of 3Pd–1B/Al2O3. A comparison of ODHP performance over various boron-based catalysts (Table S1) revealed notable differences. Due to the much lower WHSV−1 of 0.29 gC3H8−1 gcatalyst h, the 3Pd–1B/Al2O3 catalyst exhibited remarkable ODHP activity even at short contact times. In contrast, B2O3/SBA-15 showed superior ODHP performance at a higher WHSV−1 of 9.4 gC3H8−1 gcatalyst h, while SiB6 demonstrated better activity at an elevated temperature of 545 °C. Notably, the 3Pd–1B/Al2O3 catalyst maintained stable ODHP performance over 150 hours, with only a slight decline in propane conversion (Fig. 1b).
image file: d5cy00990a-f1.tif
Fig. 1 (a) Catalytic performance of ODHP over 3Pd/Al2O3, 1B/Al2O3 and 3Pd–YB/Al2O3 catalysts after 1 hour. (b) Propane conversion and propylene selectivity over time on stream for 3Pd–1B/Al2O3. Conditions: 490 °C, WHSV−1 = 0.29 gC3H8−1 gcatalyst hours, WHSV = 1800 LC3H8 kgcatalyst−1 h−1, PC3H8[thin space (1/6-em)]:[thin space (1/6-em)]PO2[thin space (1/6-em)]:[thin space (1/6-em)]PN2 = 4.4[thin space (1/6-em)]:[thin space (1/6-em)]2.2[thin space (1/6-em)]:[thin space (1/6-em)]93.4 kPa. (c) Rates of propane consumption using 3Pd–1B/Al2O3 as a function of PC3H8 (PO2 constant at 0.22 atm) at 460 °C. (d) Rates of propane consumption using 3Pd–1B/Al2O3 as a function of PO2 (PC3H8 constant at 0.44 atm) at 460 °C. (e and f) Corresponding propane consumption rates at 490 °C. (g and h) Corresponding propane consumption rates at 520 °C.

Further kinetic experiments were performed to study the influence of reactant concentrations [partial pressures of O2 and C3H8 (PO2 and PC3H8, respectively)] on the rate of propane consumption over the 3Pd–1B/Al2O3 catalyst. These experiments were aimed at identifying the active sites involved in the reaction. As shown in Fig. 1c–h, the rate of propane consumption at 460, 490 and 520 °C using 3Pd–1B/Al2O3 exhibits a first-order dependence on PC3H8 and PO2, which follows the Langmuir–Hinshelwood (L–H) mechanism (eqn (1)). According to the L–H mechanism, two distinct active sites are involved in ODHP, one for the activation of oxygen and the other for the activation of propane, facilitating the reaction cycles. In comparison, BN catalysts typically follow the Eley–Rideal (E–R) mechanism in ODHP, where the oxygen activation of the boron species leads to hydrogen abstraction from the secondary carbon of propane.1,8 However, in the BS-1 catalyst, the L–H mechanism was observed, with O2 and C3H8 co-adsorbing on the dihydroxyl group of boron.14 For the 3Pd–1B/Al2O3 catalyst, the mechanism shifted toward the L–H pathway owing to the addition of palladium, leading to enhanced propane activation and improved catalytic performance.

 
image file: d5cy00990a-t3.tif(1)

The characterization of structural and physicochemical properties

Structural and physicochemical properties were investigated to understand the nature of the active sites. XRD patterns confirm that the crystalline structure of γ-Al2O3 and θ-Al2O3 remains stable after impregnation with Pd and B (Fig. S7). EDS elemental composition and mapping reveal that Pd forms bright nanorods in the 3Pd–1B/Al2O3 catalyst (Fig. 2a–d). HRTEM images further demonstrate the presence of an interfacial heterostructure between the Pd nanorods and Al2O3, as indicated by the white dashed lines (Fig. 2e and f). This interfacial structure contributes to the interaction between Pd and Al2O3.
image file: d5cy00990a-f2.tif
Fig. 2 (a) HRTEM image of 3Pd–1B/Al2O3, and (b–d) EDS elements composition and mapping of the illustrated region. (e) STEM image of 3Pd–1B/Al2O3, and (f) magnified view of the region marked in (e). (h) B 1s XPS spectra of 3Pd/Al2O3 and 3Pd–1B/Al2O3. (i) TPR spectra of 3Pd/Al2O3, 1B/Al2O3 and 3Pd–1B/Al2O3. (j) 11B MAS NMR spectra of 1B/Al2O3 and 3Pd–1B/Al2O3 measured at 14.1 T. (k) 1H MAS NMR spectra of 1B/Al2O3 and 3Pd–1B/Al2O3 measured at 14.1 T. (l) FTIR spectra of 3Pd/Al2O3, 1B/Al2O3 and 3Pd–1B/Al2O3.

Decomposition of the Pd 3d XPS spectra provides insight into the chemical state of palladium, as shown in Fig. 2g and S8. Both 3Pd/Al2O3 and Pd–B/Al2O3 catalysts exhibited two deconvoluted signals. The 3Pd/Al2O3 catalyst displayed Pd0 species with a Pd 3d5/2 binding energy of 334.9 eV, along with PdOx species at 336.1 eV.21 For 3Pd–1B/Al2O3, the Pd0 species appeared at a binding energy of 335.4 eV, while PdOx/Pd–O–B species were also observed at 336.1 eV.23,24 The Pd0 binding energy in 3Pd–1B/Al2O3 shows a positive shift compared with that of 3Pd/Al2O3, indicating electron transfer from Pd to the adjacent electron-deficient B species. Additionally, the possible presence of PdO species with a Pd 3d5/2 binding energy near 337.0 eV could be excluded in either the 3Pd/Al2O3 or 3Pd–1B/Al2O3 catalysts.25 The B 1s XPS spectrum exhibit a signal for 1B/Al2O3 at a binding energy of 192.9 eV, which is attributed to the B–O bond of boron oxide (Fig. 2h).26 This signal shifts to a lower binding energy of 192.6 eV for 3Pd–0.5B/Al2O3 and 3Pd–1B/Al2O3 (Fig. 2h and S9), suggesting that the palladium atoms reduced boron via the Pd–O–B structure.27

H2-TPR analysis was conducted to assess the redox properties of the catalysts. As shown in Fig. 2i, 3Pd/Al2O3 exhibited two distinct reduction peaks: a prominent negative peak at 70 °C, corresponding to hydrogen evolution from the decomposition of β-Pd–Hx, and a broad peak at 848 °C, attributed to the reduction of PdOx species strongly interacting with the Al2O3 lattice.28,29 For 1B/Al2O3, two reduction peaks were observed at 398 °C and 750 °C, assigned to the reduction of B2O3 and B–O species bonded to the Al2O3 lattice, respectively. Notably, 3Pd–1B/Al2O3 exhibited a distinct reduction peak at 244 °C, indicating a lower reduction temperature attributed to Pd-facilitated reduction of B2O3, likely arising from the formation of Pd–O–B species as revealed by XPS analysis.

The Raman spectrum exhibits a band at 807 cm−1, indicating the presence of the B2O3 phase in 3Pd–1B/Al2O3 (Fig. S10). The molecular structure of B is investigated using 14.1 T 11B NMR spectroscopy (Fig. 2j). The 11B NMR signals at 15.7 and 13.2 ppm were assigned to the tricoordinated boron species, whereas the signal at 1.2 ppm was assigned to the tetracoordinated boron species.30,31 The signal at 15.7 ppm was attributed to the boroxol-ring species associated with the terminal boron hydroxyl groups (BOx(OH)3−x, where x = 1 or 2), and the signal at 13.2 ppm was ascribed to the non-ring species related to the framework structure as B(OAl)3.30,31 Compared to 1B/Al2O3, a substantial decrease in the B(OAl)3 signal in 3Pd–1B/Al2O3 indicated an alteration in the boron framework structure by palladium. 1H NMR spectra of 1B/Al2O3 and 3Pd–1B/Al2O3 reveal split peaks at 1.2–1.8 ppm, corresponding to terminal hydroxyl groups on the Al2O3 surface and B sites, along with a signal at 4.9 ppm assigned to adsorbed H2O (Fig. 2k).14 The shift of the terminal hydroxyl group signals to a lower chemical shift in 3Pd–1B/Al2O3 suggests that the interaction of palladium with Al2O3 affects the electronegativity of these hydroxyl groups. The infrared transmission spectra of 3Pd/Al2O3, 1B/Al2O3 and 3Pd–1B/Al2O3 were analyzed (Fig. 2l). Compared to 3Pd/Al2O3, characteristic peaks corresponding to B2O3 were observed in the range of 1200–1600 cm−1 for 1B/Al2O3 and 3Pd–1B/Al2O3. In addition, peaks in the range of 2800–3000 cm−1 for 1B/Al2O3 and 3Pd–1B/Al2O3 were attributed to H3BO3, indicating the presence of BOH groups.32

Based on the L–H mechanism revealed by kinetic studies, two types of active sites are involved in the ODHP reaction over the 3Pd–1B/Al2O3 catalyst. Because the tricoordinated B species are more active than the tetracoordinated ones,3,5,6,12 BOx(OH)3−x and B(OAl)3 are likely responsible for oxygen activation. Furthermore, in comparison to 1B/Al2O3, BOx(OH)3−x species (but not the B(OAl)3 species) are retained in 3Pd–1B/Al2O3, while the propylene selectivity remains similar between the two catalysts (Fig. 1a). This suggests that BOx(OH)3−x species are the key active sites for oxygen activation. In addition, the increased propane conversion observed upon Pd doping indicates that Pd sites are likely the primary active sites for propane dehydrogenation.

The investigation of reaction pathways

To reveal the synergistic effect of the two active sites in enhancing ODHP performance, the reaction pathways over the 1B/Al2O3, 3Pd/Al2O3, and 3Pd–1B/Al2O3 catalysts were investigated. The production of H2O2 was measured both in the gas phase and on the catalyst's surface during ODHP. As shown in Fig. 3a, a gas mixture containing 6.2% C3H8, 3.1% O2, and N2 balance was passed through a heated quartz tube reactor (490 °C) containing 100 mg of the catalyst at a flow rate of 70 mL min−1. The effluent gas was then passed through a cooled quartz-tube (approximately 20 °C) to condense gaseous H2O2, which was collected over 1 h of ODHP and diluted in 1 mL of ultrapure water. Following the reaction, H2O2 on the catalyst surface was washed with 1 mL of ultrapure water. H2O2 is immediately quantified using the potassium titanium oxalate (PTO) color reaction, and the concentration is calculated using standard calibration curves (Fig. S11 and S12). As shown in Fig. 3b, the concentrations of gaseous and catalytic surface H2O2 are 0.159 vs. 0.011, 0.015 vs. 0.046, and 0.004 vs. 0.027 mg L−1 for 1B/Al2O3, 3Pd/Al2O3, and 3Pd–1B/Al2O3, respectively. These results demonstrate that boron sites primarily contribute to the formation of gaseous H2O2, which is typically generated via a radical route over boron-based catalysts.6,7,12,16 In contrast, Pd sites predominantly generate catalyst surface H2O2 owing of their efficient direct H2O2 synthesis capabilities.33,34 Furthermore, doping palladium onto B/Al2O3 significantly decreased gaseous H2O2 formation, likely suppressing the radical pathway.
image file: d5cy00990a-f3.tif
Fig. 3 (a) Schematic image of the experimental setup for measuring gaseous and catalyst surface H2O2 concentrations in a quartz tube reactor after ODHP reaction for 1 hour. (b) Gaseous and catalyst surface H2O2 concentration after 1 hour of ODHP reaction over 3Pd/Al2O3, 1B/Al2O3 and 3Pd–1B/Al2O3 catalysts. (c) EPR spectra of radicals after 10 minutes of ODHP reaction over 3Pd/Al2O3, 1B/Al2O3 and 3Pd–1B/Al2O3 catalysts. (d) DRIFT spectra after 20 minutes of ODHP reaction over 3Pd/Al2O3, 1B/Al2O3 and 3Pd–1B/Al2O3 catalysts.

Next, the radical species present on the catalyst surface after 10 min of ODHP were investigated using EPR. As shown in Fig. 3c, hydroxyl radicals (·OH) are observed on the 1B/Al2O3 catalyst, which likely originates from decomposition of gaseous H2O2. In contrast, oxygen vacancies (Ov) were detected in the 3Pd/Al2O3 catalyst, which could be attributed to the consumption of lattice oxygen during the complete oxidation of propane, following the Mars–van Krevelen mechanism.35 Notably, neither ·OH nor Ov were detected on the 3Pd–1B/Al2O3 catalyst, suggesting that 3Pd–1B/Al2O3 inhibited the complete oxidation of propane and the formation of gaseous H2O2, which typically formed through the radical pathway. Subsequently, the ODHP reaction was investigated using DRIFTS. As shown in Fig. 3d, a band at 3735 cm−1 is observed for the 1B/Al2O3 and 3Pd–1B/Al2O3 catalysts, assignable to B–OH species. However, an additional band at 3695 cm−1 appeared in the spectrum of 3Pd–1B/Al2O3, corresponding to the bridged OH species,36 likely BOH⋯C3H7 (or C3H6). This bridged OH structure indicated that the dehydrogenated C3H8 species (C3H7 or C3H6) were confined to the catalyst surface and interacted with the surface hydroxyl groups. These results confirmed that the 3Pd–1B/Al2O3 catalyst followed a surface-catalyzed reaction pathway. In addition, a broad band in the range of 3280–3550 cm−1, attributed to hydrogen-bonded OH species, was observed for the 3Pd–1B/Al2O3 catalysts. This band is indicative of H2O generation.37 H2O has previously been shown to occur in the surface-catalyzed pathway in ODHP.13,14

We employed quasi in situ XPS to investigate whether C3H8 and O2 influence the chemical states of Pd and B sites on 1B/Al2O3, 3Pd/Al2O3, and 3Pd–1B/Al2O3 catalysts during ODHP. After 1 hour of reaction, notable changes were observed in the Pd 3d and B 1s spectra compared to the fresh catalysts (Fig. 4a to d). For 3Pd/Al2O3 (Fig. 4a), the Pd 3d peaks at 335.1 eV and 336.5 eV, corresponding to Pd0 and PdOx respectively, showed a compositional shift, with the proportion of Pd0 increasing from 55.9% to 77.7% and PdOx decreasing from 44.1% to 22.3%. This indicates that PdOx was reduced during ODHP. In the case of 3Pd–1B/Al2O3 (Fig. 4b), the signal attributed to PdOx/Pd–O–B species at 336.9 eV decreased from 26.9% before the reaction to nearly undetectable levels after ODHP, suggesting a transformation in the interfacial Pd–O–B bonding. These results for both 3Pd/Al2O3 and 3Pd–1B/Al2O3 indicate that oxidized Pd species were reduced to metallic Pd0 by hydrocarbons or hydrogen during ODHP, which can be attributed to the inherent dehydrogenation ability of Pd. In the B 1s spectra (Fig. 4c and d), both 1B/Al2O3 and 3Pd–1B/Al2O3 exhibited a downward shift of approximately 0.3 eV after ODHP, indicating the reduction of B–O bonds, likely due to the formation of BOH groups during the reaction, as observed in DRIFT of ODHP (Fig. 3d).


image file: d5cy00990a-f4.tif
Fig. 4 Quasi in situ XPS spectra of the Pd 3d region for 3Pd/Al2O3 (a) and 3Pd–1B/Al2O3 (b), and the B 1s region for 1B/Al2O3 (c) and 3Pd–1B/Al2O3 (d), recorded before and after ODHP reaction for 1 h.

DFT calculations are conducted to further elucidate the reaction pathway of ODHP on B/Al2O3 (Fig. S13) and Pd–B/Al2O3 catalysts (Fig. 5). Based on structural characterization, the γ-Al2O3(110) surface featuring BO2(OH) sites was selected as the representative model. The γ-Al2O3(110) surface was chosen because it represents the most thermodynamically stable and dominant crystal facet, as confirmed by both experimental XRD reflections (Fig. S7) and previous theoretical studies. Other facets of γ-Al2O3 and θ-Al2O3, as well as BO(OH)2 sites, were not studied. The adsorption energy of C3H8 on the BOH site differed significantly between the two catalysts: −0.196 eV for B/Al2O3 and −4.610 eV for Pd–B/Al2O3, indicating that C3H8 adsorbs much more readily on the Pd–B/Al2O3 surface. Because of the weak adsorption of propyl species on the BOH sites in B/Al2O3, C3H7 and C3H6 were more likely to exist as radicals than in the adsorbed states. The energy barrier for the formation of OOH from O2 gas was 0.817 eV, whereas that for the H atom abstraction from the C3H7 radical to form image file: d5cy00990a-t4.tif was significantly higher reaching 3.564 eV. Consequently, image file: d5cy00990a-t5.tif desorption from the B/Al2O3 surface was more favorable, facilitating gaseous H2O2 formation in the radical pathway. In the radical pathway, dehydrogenation of the C3H7 radical (3.564 eV) represents the most energy-intensive step. In contrast, for the Pd–B/Al2O3 catalyst, the energy barrier for the dehydrogenation of C3H7 over a Pd site to form a Pd–H state is much lower, at just 0.334 eV (Fig. 5). The subsequent migration of the H atom from Pd to the B–O site requires slightly more energy; however, this H spillover mechanism is critical for the ODHP reaction on Pd–B/Al2O3. The primary products were C3H6 and H2O obtained in this surface-catalyzed pathway aligned with the experimental results. When H2O desorbs from the catalyst surface, it creates oxygen vacancies that are rapidly replenished by O2, restoring the BO2(OH) sites. Therefore, the boron sites activate and provide oxygen for the oxidation of the migrated hydrogen to promote hydrogen oxidation, while inhibiting the overoxidation of hydrocarbons to COx. Moreover, the Pd sites facilitated propane dehydrogenation. This synergistic interaction between the two active sites in the Pd–B/Al2O3 catalyst enhanced propylene selectivity and propane conversion during the ODHP process. Raman analysis of the spent catalyst reveals characteristic bands at 1350 and 1573 cm−1 (Fig. S14), corresponding to graphitic carbon species,38 indicating that carbon deposition is the main deactivation mechanism of 3Pd–1B/Al2O3 during long-term ODHP testing (Fig. 1b).


image file: d5cy00990a-f5.tif
Fig. 5 DFT calculated energy profile for ODHP over Pd–B/Al2O3 following a surface-catalyzed pathway. The diagram presents the energy profile along the reaction coordinate, highlighting the initial state, transition state, and final state from left to right. The models (a–h) show side views of the geometries for the various states. The atoms are color-coded as follows: Pd in grey, B in green, Al in light blue, O in red, C in dark brown, and H in pink.

The positive effect of surface OH on ODHP

H2O is frequently employed to enhance surface OH groups on catalysts, a strategy well supported by existing literature.39,40 To elucidate the critical role of surface OH in ODHP, we conducted experiments using 3Pd–1B/Al2O3 catalysts with systematically controlled surface OH concentrations. In these experiments, humidified feed gas was used to increase surface OH, while physically mixing the catalyst with a desiccant aimed to reduce it. Control studies confirmed that the desiccant alone had no noticeable impact on propane conversion or propylene selectivity (Fig. S15). As shown in Fig. 6a, under humid-feed conditions, propane conversion increased by approximately 0.5% and propylene selectivity by approximately 5% compared to dry-feed conditions. In contrast, introducing a desiccant to the 3Pd–1B/Al2O3 catalyst significantly suppressed catalytic activity, decreasing propane conversion from 20% to 13% and propylene selectivity from 51% to 9% (Fig. 6b). These results indicate a direct correlation between surface OH concentration and ODHP performance.
image file: d5cy00990a-f6.tif
Fig. 6 (a) Propane conversion and propylene selectivity over time on stream for 3Pd–1B/Al2O3 in a dry feed gas and in a humid feed gas containing saturated water vapor at 25 °C. (b) Propane conversion and propylene selectivity over time on stream for 3Pd–1B/Al2O3 compared to 3Pd–1B/Al2O3 physically mixed with a desiccant. Conditions: 490 °C, WHSV−1 = 0.29 gC3H8−1 gcatalyst h, WHSV = 1800 LC3H8 kgcatalyst−1 h−1, PC3H8[thin space (1/6-em)]:[thin space (1/6-em)]PO2[thin space (1/6-em)]:[thin space (1/6-em)]PN2 = 4.4[thin space (1/6-em)]:[thin space (1/6-em)]2.2[thin space (1/6-em)]:[thin space (1/6-em)]93.4 kPa. (c) In situ DRIFT of ODHP with saturated H2O gas over 3Pd–1B/Al2O3. (d) In situ DRIFT of ODHP with saturated D2O gas over 3Pd–1B/Al2O3. (e) The reaction mechanism of ODHP over 3Pd–1B/Al2O3.

DRIFT spectroscopy further supported the beneficial role of surface OH groups. During ODHP with H2O-saturated feed, signals corresponding to B–OH and bridged OH species increased over time (Fig. 6c), whereas B–OD signals increased markedly under D2O-saturated conditions (Fig. 6d). These findings suggest that both H2O and D2O dissociate to form OH (or OD) groups on the 3Pd–1B/Al2O3 surface, enriching B–OH sites during ODHP. Based on these observations, we propose an ODHP reaction mechanism over 3Pd–1B/Al2O3 (Fig. 6e) that begins with propane dehydrogenation at Pd sites, followed by hydrogen transfer to BOx(OH)3−x sites, where oxidation occurs. The observed increase in B–OH concentration is positively correlated with enhanced catalytic performance in ODHP.

Conclusions

In summary, this study demonstrates that Pd doping effectively alters the reaction mechanism of ODHP on B/Al2O3 catalysts. The introduction of Pd sites shifts the dominant pathway from a gas-phase radical mechanism, which is characterized by the formation of gaseous H2O2, to a surface-catalyzed Langmuir–Hinshelwood mechanism. Quantitative analysis of H2O2 shows that Pd–B/Al2O3 produces significantly less gaseous H2O2 compared to B/Al2O3, indicating suppression of the radical pathway. The presence of synergistic active sites enables this transition, where Pd sites promote propane adsorption and lower the activation barrier for C3H7 dehydrogenation, while adjacent BOx(OH)3−x species selectively oxidize hydrogen to H2O, thereby preventing over-oxidation to COx. The combined effect of these bifunctional sites enhances propane conversion by 22% (from 15.9% to 19.3%) while maintaining a propylene selectivity of 55%. Surface B–OH further contribute to catalytic performance, as supported by humid-feed experiments showing an approximately 5% increase in propylene selectivity. This work offers a promising strategy for designing efficient ODHP catalysts by integrating noble metals like Pd with boron-based materials to harness surface-mediated pathways and overcome limitations associated with radical mechanisms. This work offers a promising strategy for designing efficient ODHP catalysts by integrating noble metals like Pd with boron-based materials to harness surface-mediated pathways and overcome limitations associated with radical mechanisms. Future efforts to enhance the catalytic efficiency of Pd–B/Al2O3 will focus on tuning the Pd–B interaction through controlled dispersion and electronic modulation, as well as optimizing the support properties to achieve a better balance between propane activation and propylene selectivity.

Author contributions

C. M. performed most experiments and wrote the initial draft. C. C. contributed to discussion. Z. H. carried out catalyst synthesis. Z. Y., F. W. and L. B. contributed to a part of characterization. M. S. was responsible for supervision, review and draft revision. G. J. was responsible for review.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the supplementary information (SI).

Supplementary information is available. See DOI: https://doi.org/10.1039/d5cy00990a.

Acknowledgements

Financial support by the National Natural Science Foundation of China (grants 22125606, 22376205, 21976198).

References

  1. 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, 1570–1573 CrossRef CAS PubMed.
  2. H. Li, J. Zhang, P. Wu, S. Xun, W. Jiang, M. Zhang, W. Zhu and H. Li, O2 Activation and oxidative dehydrogenation of propane on hexagonal boron nitride: Mechanism revisited, J. Phys. Chem. C, 2019, 123, 2256–2266 CrossRef CAS.
  3. M. C. Cendejas, O. A. Paredes Mellone, U. Kurumbail, Z. Zhang, J. H. Jansen, F. Ibrahim, S. Dong, J. Vinson, A. N. Alexandrova, D. Sokaras, S. R. Bare and I. Hermans, Tracking active phase behavior on boron nitride during the oxidative dehydrogenation of propane using operando X-ray raman spectroscopy, J. Am. Chem. Soc., 2023, 145, 25686–25694 CrossRef CAS PubMed.
  4. Z. Zhang, J. Tian, X. Wu, I. Surin, J. Perez-Ramirez, P. Hemberger and A. Bodi, Unraveling radical and oxygenate routes in the oxidative dehydrogenation of propane over boron nitride, J. Am. Chem. Soc., 2023, 145, 7910–7917 CrossRef CAS PubMed.
  5. Z. Zhang, I. Hermans and A. N. Alexandrova, Off-stoichiometric restructuring and sliding dynamics of hexagonal boron nitride edges in conditions of oxidative dehydrogenation of propane, J. Am. Chem. Soc., 2023, 145, 17265–17273 CrossRef CAS PubMed.
  6. X. Zhang, R. You, Z. Wei, X. Jiang, J. Yang, Y. Pan, P. Wu, Q. Jia, Z. Bao, L. Bai, M. Jin, B. Sumpter, V. Fung, W. Huang and Z. Wu, Radical chemistry and reaction mechanisms of propane oxidative dehydrogenation over hexagonal boron nitride catalysts, Angew. Chem., Int. Ed., 2020, 59, 8042–8046 CrossRef CAS PubMed.
  7. J. M. Venegas, Z. Zhang, T. O. Agbi, W. P. McDermott, A. Alexandrova and I. Hermans, Why boron nitride is such a selective catalyst for the oxidative dehydrogenation of propane, Angew. Chem., Int. Ed., 2020, 59, 16527–16535 CrossRef CAS PubMed.
  8. J. Tian, J. Tan, M. Xu, Z. Zhang, S. Wan, S. Wang, J. Lin and Y. Wang, Propane oxidative dehydrogenation over highly selective hexagonal boron nitride catalysts: The role of oxidative coupling of methyl, Sci. Adv., 2019, 5, eaav8063 CrossRef CAS PubMed.
  9. L. Shi, D. Wang, W. Song, D. Shao, W. P. Zhang and A. H. Lu, Edge-hydroxylated Boron Nitride for Oxidative Dehydrogenation of Propane to Propylene, ChemCatChem, 2017, 9, 1788–1793 CrossRef CAS.
  10. L. Zhu, Z. Liu, Q. Zhou, W.-D. Lu and D. Wang, Propane oxidative dehydrogenation on nanosized boron carbide: Effect of boron content and its oxidation implicated by DFT calculations, J. Phys. Chem. C, 2023, 127, 6280–6293 CrossRef CAS.
  11. J. Tian, G. Collinge, S. F. Yuk, J. Lin, V.-A. Glezakou, M.-S. Lee, Y. Wang and R. Rousseau, Dynamically formed active sites on liquid boron oxide for selective oxidative dehydrogenation of propane, ACS Catal., 2023, 13, 8219–8236 CrossRef CAS.
  12. W.-D. Lu, D. Wang, Z. Zhao, W. Song, W.-C. Li and A.-H. Lu, Supported boron oxide catalysts for selective and low-temperature oxidative dehydrogenation of propane, ACS Catal., 2019, 9, 8263–8270 CrossRef CAS.
  13. H. Yan, S. Alayoglu, W. Wu, Y. Zhang, E. Weitz, P. C. Stair and J. M. Notestein, Identifying boron active sites for the oxidative dehydrogenation of propane, ACS Catal., 2021, 11, 9370–9376 CrossRef CAS.
  14. H. Zhou, X. Yi, Y. Hui, L. Wang, W. Chen, Y. Qin, M. Wang, J. Ma, X. Chu, Y. Wang, X. Hong, Z. Chen, X. Meng, H. Wang, Q. Zhu, L. Song, A. Zheng and F.-S. Xiao, Isolated boron in zeolite for oxidative dehydrogenation of propane, Science, 2021, 372, 76–80 CrossRef CAS PubMed.
  15. R. Belgamwar, A. G. M. Rankin, A. Maity, A. K. Mishra, J. S. Gómez, J. Trébosc, C. P. Vinod, O. Lafon and V. Polshettiwar, Boron nitride and oxide supported on dendritic fibrous nanosilica for catalytic oxidative dehydrogenation of propane, ACS Sustainable Chem. Eng., 2020, 8, 16124–16135 CrossRef CAS.
  16. X. Gao, M. Liu, Y. Huang, W. Xu, X. Zhou and S. Yao, Dimensional understanding of boron-based catalysts for oxidative propane dehydrogenation: Structure and mechanism, ACS Catal., 2023, 13, 9667–9687 CrossRef CAS.
  17. H. Wang, H. Li, J. Duan, L. Wang and F. S. Xiao, Adjustment of molecular sorption equilibrium on catalyst surface for boosting catalysis, Acc. Chem. Res., 2025, 58, 440–451 CrossRef CAS PubMed.
  18. X. Li, G. Hai, D. H. C. Wan, Y. Liao, Z. Yao, F. Zhao, L. Huang, J. Zhou, G. Li, G. F. Chen, F. R. Wang, M. K. H. Leung and H. Wang, Favoring the originally unfavored oxygen for enhancing nitrogen-to-nitrate electroconversion, J. Am. Chem. Soc., 2025, 147, 8587–8596 CrossRef CAS PubMed.
  19. M. Peng, Z. Jia, Z. Gao, M. Xu, D. Cheng, M. Wang, C. Li, L. Wang, X. Cai, Z. Jiang, H. Jiang, N. Wang, D. Xiao, H. Liu and D. Ma, Antisintering Pd1 catalyst for propane direct dehydrogenation with in situ active sites regeneration ability, ACS Catal., 2022, 12, 2244–2252 CrossRef CAS.
  20. F. Jiang, L. Zeng, S. Li, G. Liu, S. Wang and J. Gong, Propane dehydrogenation over Pt/TiO2–Al2O3 catalysts, ACS Catal., 2014, 5, 438–447 CrossRef.
  21. P. Wang, X. Zhang, R. Shi, J. Zhao, G. I. N. Waterhouse, J. Tang and T. Zhang, Photocatalytic ethylene production by oxidative dehydrogenation of ethane with dioxygen on ZnO-supported PdZn intermetallic nanoparticles, Nat. Commun., 2024, 15, 789 CrossRef CAS PubMed.
  22. C. Duan, J. Liu, Z. Li, R. Shi, J. Zhao, G. I. N. Waterhouse, X. Wen, L. Zhang, L. Wu and T. Zhang, Efficient photocatalytic propane direct dehydrogenation to propylene over PtO2 clusters, Adv. Mater., 2025, 37, 2411648 CrossRef CAS PubMed.
  23. S. Yue, C. S. Praveen, A. Klyushin, A. Fedorov, M. Hashimoto, Q. Li, T. Jones, P. Liu, W. Yu, M. G. Willinger and X. Huang, Redox dynamics and surface structures of an active palladium catalyst during methane oxidation, Nat. Commun., 2024, 15, 4678 CrossRef CAS PubMed.
  24. K. Tang, Y. Ren, W. Liu, J. Wei, J. Guo, S. Wang and Y. Yang, Insight investigation of active palladium surface sites in palladium-ceria catalysts for NO + CO reaction, ACS Appl. Mater. Interfaces, 2018, 10, 13614–13624 CrossRef CAS.
  25. G. Wang, Y. Yan, X. Zhang, X. Gao and Z. Xie, Three-dimensional porous hexagonal boron nitride fibers as metal-free catalysts with enhanced catalytic activity for oxidative dehydrogenation of propane, Ind. Eng. Chem. Res., 2021, 60, 17949–17958 CrossRef CAS.
  26. Z.-L. Du, Q.-Q. Dang and X.-M. Zhang, Heptazine-based porous framework supported palladium nanoparticles for green suzuki–miyaura reaction, Ind. Eng. Chem. Res., 2017, 56, 4275–4280 CrossRef CAS.
  27. J. T. Grant, W. P. McDermott, J. M. Venegas, S. P. Burt, J. Micka, S. P. Phivilay, C. A. Carrero and I. Hermans, Boron and boron-containing catalysts for the oxidative dehydrogenation of propane, ChemCatChem, 2017, 9, 3623–3626 CrossRef CAS.
  28. T. Fovanna, M. Nachtegaal, A. H. Clark, O. Kröcher and D. Ferri, Preparation, quantification, and reaction of Pd hydrides on Pd/Al2O3 in liquid environment, ACS Catal., 2023, 13, 3323–3332 CrossRef CAS.
  29. M. Yarar, A. Bouziani and D. Uner, Pd as a reduction promoter for TiO2: Oxygen and hydrogen transport at 2D and 3D Pd interfaces with TiO2 monitored by TPR, operando 1H NMR and CO oxidation studies, Catal. Commun., 2023, 174, 106580 CrossRef CAS.
  30. R. W. Dorn, L. O. Mark, I. Hung, M. C. Cendejas, Y. Xu, P. L. Gor'kov, W. Mao, F. Ibrahim, Z. Gan, I. Hermans and A. J. Rossini, An atomistic picture of boron oxide catalysts for oxidative dehydrogenation revealed by ultrahigh field 11B-17O solid-state NMR spectroscopy, J. Am. Chem. Soc., 2022, 144, 18766–18771 CrossRef CAS PubMed.
  31. R. W. Dorn, M. C. Cendejas, K. Chen, I. Hung, N. R. Altvater, W. P. McDermott, Z. Gan, I. Hermans and A. J. Rossini, Structure determination of boron-based oxidative dehydrogenation heterogeneous catalysts with ultra-high field 35.2 T 11B solid-state NMR spectroscopy, ACS Catal., 2020, 10, 13852–13866 CrossRef CAS.
  32. G. Bilgiç, N. Korkmaz, M. Şahin and A. Karadağ, Synthesis, structural, and electrochemical properties of boron-based ionic liquid, Ionics, 2022, 28, 3289–3300 CrossRef.
  33. S. Liu, Y. Deng, L. Fu, L. Huang, L. Ouyang and S. Yuan, Understanding the role of boron on the interface modulation of the Pd/TiO2 catalyst for direct synthesis of H2O2, ACS Sustainable Chem. Eng., 2022, 10, 3264–3275 CrossRef CAS.
  34. R. Liu, C. Wang, Y. Yan, R. Wang and G. Chen, Reversed charge transfer to modulate the d-band center of Pd for efficient direct H2O2 synthesis, ACS Catal., 2024, 14, 3955–3965 CrossRef CAS.
  35. R. Rajesh, S. Lingalwar, P. Challa, P. Balla, B. Shanmugavelu, V. P. Sruthi, S. Kim, S. Senthilkumar and V. Perupogu, Low-temperature total oxidation of propane using silver-decorated MnO2 nanorods, ACS Appl. Nano Mater., 2023, 6, 12258–12270 CrossRef CAS.
  36. M. Shamzhy, B. Gil, M. Opanasenko, W. J. Roth and J. Čejka, MWW and MFI frameworks as model layered zeolites: structures, transformations, properties, and activity, ACS Catal., 2021, 11, 2366–2396 CrossRef CAS.
  37. M. G. Bridelli and P. R. Crippa, Infrared and water sorption studies of the hydration structure and mechanism in natural and synthetic melanin, J. Phys. Chem. B, 2010, 114, 9381–9390 CrossRef CAS PubMed.
  38. J. Wu, M. Lin, X. Cong, H. Liu and P. Tan, Raman spectroscopy of graphene-based materials and its applications in related devices, Chem. Soc. Rev., 2018, 47, 1822 RSC.
  39. W. Cen, Y. Liu, Z. Wu, J. Liu, H. Wang and X. Weng, Cl Species Cl Species transformation on CeO2(111) surface and its effects on CVOCs catalytic abatement: A first-principles investigation, J. Phys. Chem. C, 2014, 118, 6758–6766 CrossRef CAS.
  40. L. M. Petrick, S. Sabach and Y. Dubowski, Degradation of VX surrogate profenofos on surfaces via in situ photo-oxidation, Environ. Sci. Technol., 2013, 47, 8751–8758 CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.