Next Article in Journal
Carbonation Deactivation of Limestone in a Micro-Fluidized Bed Reactor
Previous Article in Journal
Highly Efficient and Stable Ni-Cs/TS-1 Catalyst for Gas-Phase Propylene Epoxidation with H2 and O2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Alkaline Earth Carbonate Engineered Pt Electronic States for High-Efficiency Propylene Oxidation at Low Temperatures

1
State Key Laboratory of Green Chemical Engineering and Industrial Catalysis, Research Institute of Industrial Catalysis, School of Chemistry and Molecular Engineering, East China University of Science and Technology, Shanghai 200237, China
2
School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
3
State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering, College of Chemistry and Chemical Engineering, Ningxia University, Yinchuan 750021, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2025, 15(8), 696; https://doi.org/10.3390/catal15080696
Submission received: 25 June 2025 / Revised: 13 July 2025 / Accepted: 14 July 2025 / Published: 22 July 2025
(This article belongs to the Section Catalytic Materials)

Abstract

Alkaline earth elements have emerged as crucial electronic modifiers for regulating active sites in catalytic systems, yet the influence of metal–support interactions (MSIs) between alkaline earth compounds and active metals remains insufficiently understood. This study systematically investigated Pt nanoparticles supported on alkaline earth carbonates (Pt/MCO3, M = Mg, Ca, Ba) for low-temperature propylene combustion. The Pt/BaCO3 catalyst exhibited outstanding performance, achieving complete propylene conversion at 192 °C, significantly lower than Pt/MgCO3 (247 °C) and Pt/CaCO3 (282 °C). The enhanced activity stemmed from distinct MSI effects among the supports, with Pt/BaCO3 showing the poorest electron enrichment and lowest propylene adsorption energy. Through kinetic analyses, 18O2 isotope labeling, and comprehensive characterization, the reaction was confirmed to follow the Mars–van Krevelen (MvK) mechanism. Pt/BaCO3 achieves an optimal balance between propylene and oxygen adsorption, a critical factor underlying its superior activity.

Graphical Abstract

1. Introduction

Volatile organic compounds (VOCs), as a class of carbon-based compounds readily volatilized at ambient temperature, not only serve as primary precursors for photochemical smog and ozone layer depletion but also exhibit significant correlations with the incidence of respiratory diseases, posing dual threats to both ecological systems and public health [1,2]. Among these, propylene, a representative unsaturated hydrocarbon VOC emitted by the petrochemical industry, presents a critical challenge for atmospheric pollution control in chemical industrial parks [3,4,5]. Among various VOC combustion catalysts, Pt-based catalysts have demonstrated prominent application potential in catalytic combustion due to their exceptional C-H bond activation and oxygen dissociation capabilities [6,7]. Notably, the catalytic combustion mechanism of propylene fundamentally differs from that of propane; its carbon–carbon double bond structure induces strong chemisorption on noble metal surfaces, while competitive adsorption with oxygen at active sites leads to complex catalytic kinetic behaviors [8,9]. Consequently, regulating the electronic states of active sites through metal–support interactions to balance the adsorption energies of propylene and oxygen has emerged as a pivotal scientific challenge for achieving efficient olefin catalytic combustion [10,11].
Alkali and alkaline earth metal elements (e.g., Na+, K+, Ca2+, Mg2+, Ba2+), with their distinctive electron transfer properties, offer an effective strategy for precisely modulating the electronic structure of metal active centers [12]. Among these, alkaline earth metal carbonates (e.g., CaCO3) have demonstrated remarkable performance in catalytic systems such as CO2 hydrogenation and transesterification reactions, owing to their moderate Lewis basicity, superior CO2 adsorption capacity, and high-temperature stability [13,14,15]. Recent studies have further unveiled novel mechanisms underlying metal–carbonate interfacial effects; Zhu et al. reported H2 pretreatment-induced dynamic reconstruction in Ni/BaCO3 systems, where the surface migration of BaCO3 formed a mesoporous overlayer encapsulating Ni nanoparticles, successfully constructing electron-enriched Ni-BaCO3 interface. This unique strong metal–support interaction (SMSI) not only suppressed high-temperature sintering of metal particles (stabilizing particle size at ~5 nm) but also significantly enhanced CO2 activation efficiency [16]. Nevertheless, critical knowledge gaps remain regarding structure-activity relationships in alkaline earth metal carbonate-supported Pt catalysts for low-temperature propylene catalytic combustion, particularly the regulatory mechanisms of carbonate carrier types on Pt electronic structures and the quantitative correlation between metal–support interaction strength and propylene–oxygen adsorption equilibrium.
This study systematically investigates the influence of carrier properties on low-temperature propylene catalytic combustion performance by constructing a series of alkaline earth metal carbonate-supported Pt catalysts (Pt-MCO3, M = Mg/Ca/Ba). Activity evaluations reveal that Pt-BaCO3 exhibits optimal performance, with its T90 (complete propylene conversion temperature) reduced by 55 °C and 90 °C compared to Pt/MgCO3 and Pt/CaCO3, respectively. Combining kinetic analysis and 18O isotopic labeling experiments, we first confirm that the Pt-MCO3 system follows a modified Mars–van Krevelen (MvK) reaction mechanism, where adsorbed propylene directly reacts with lattice oxygen at Pt-MCO3 interfaces. Through in situ DRIFTS characterization and density functional theory (DFT) calculations, we elucidate that the weaker metal–support interaction in the Pt-BaCO3 system results in a higher proportion of Pt4+ species. This electronic configuration effectively mitigates excessive site occupation induced by propylene adsorption, while simultaneously accelerating the regeneration kinetics of interfacial lattice oxygen. This research not only provides new insights for designing high-efficiency VOC purification catalysts but also deepens the theoretical understanding of electronic synergy mechanisms at metal–carbonate interfaces.

2. Result and Discussion

2.1. Synthesis and Characterization of Pt-MCO3

The Pt-MgCO3, CaCO3, and BaCO3 samples were synthesized through a wet-impregnation method, with Pt loading controlled at approximately 1 wt.% (ICP analysis confirmed actual loadings of 0.92 wt.%, 0.94 wt.%, and 0.98 wt.% for MgCO3, CaCO3, and BaCO3 substrates, respectively). The XRD patterns of Pt/MCO3 (M = Mg, Ca, Ba) catalysts and their corresponding supports reveal that both Pt-CaCO3 and Pt-MgCO3 catalysts exhibit characteristic diffraction patterns corresponding to the calcite phase with rhombohedral symmetry, while Pt-BaCO3 displays distinct witherite phase reflections with orthorhombic symmetry (Figure 1a). This adoption of the BaCO3 witherite phase (instead of the calcite phase) is necessitated by the limited synthetic accessibility of calcite-phase BaCO3 [17,18]. Notably, no diffraction peaks associated with metallic Pt or PtOX species are detected in any of the Pt-loaded samples, suggesting that Pt is highly dispersed [19]. The Raman spectra of Pt-BaCO3, Pt-CaCO3, and Pt-MgCO3 exhibit characteristic carbonate vibrational modes, with all catalysts showing the ν3(CO32−) asymmetric stretching vibration at 1088 cm−1 (Figure 1b). Significantly, Pt-BaCO3 displays an additional symmetric stretching band at 1062 cm−1 [20]. In Pt-CaCO3 and Pt-MgCO3, the high symmetry of the trigonal crystal system preserves the degeneracy of CO32− vibrational modes (yielding a single peak), whereas the diminished symmetry in orthorhombic Pt-BaCO3 lifts this degeneracy, inducing ν3 vibrational splitting. These spectroscopic observations correlate precisely with XRD structural determinations [21,22].
To elucidate the morphological and textural characteristics of Pt-MCO3 (M = Mg, Ca, Ba) catalysts, a combination of N2 physisorption analysis and scanning electron microscopy (SEM) is systematically employed. As shown in Figure 1c, all three materials demonstrated limited N2 adsorption capacities across the relative pressure range (P/P0 = 0.05–0.95). BET surface area calculations established the following specific surface area sequence: Pt-MgCO3 (27.6 m2·g−1) > Pt-BaCO3 (3.5 m2·g−1) > Pt-CaCO3 (1.4 m2·g−1). SEM characterization reveals distinct morphological features that correlate precisely with the crystallographic phases identified by XRD and Raman spectroscopy. Pt-BaCO3 exhibits characteristic columnar structures (Figure S1), consistent with the orthorhombic symmetry of witherite, while both Pt-CaCO3 and Pt-MgCO3 display well-defined cubic particles (Figures S2 and S3), matching the expected morphology for trigonal calcite systems.
Thermal gravimetric (TG) tests were conducted on Pt-MCO3 (M = Mg, Ca, Ba) catalysts to study their thermal stability under reaction temperature conditions (150–300 °C). As Figure S4 shows, all three catalysts exhibit excellent thermal stability within the reaction temperature range (150–300 °C), with no significant mass loss observed. Pt-BaCO3 and Pt-CaCO3 in particular demonstrate superior high-temperature structural stability, with almost no weight loss within this range. In contrast, Pt-MgCO3 exhibits relatively lower thermal stability, beginning to decompose at temperatures between 300 and 500 °C. However, it remains structurally stable within the reaction temperature range (below 300 °C), with a weight loss rate of less than 5.6%.
The dispersion characteristics of Pt nanoparticles supported on various carbonate substrates (MCO3, M = Mg, Ca, Ba) were quantitatively evaluated by CO pulse chemisorption measurements (Figure 1d). CO pulse chemisorption analysis indicates that the Pt-BaCO3 catalyst exhibits the highest CO uptake, with a quantitatively derived average Pt particle size of 1.6 nm, confirming the predominance of highly exposed sub-nanoclusters or small nanoparticles. This phenomenon is presumably attributed to the enhanced surface alkalinity of BaCO3 during wet impregnation, which facilitates superior Pt dispersion [23]. In contrast, Pt-CaCO3 shows the lowest CO adsorption capacity, corresponding to a larger Pt grain size of 3.5 nm. Pt-MgCO3 presents an intermediate adsorption behavior, with an estimated particle size of 2.6 nm. The slightly distinct Pt particle sizes observed across the three catalysts, likely originating from variations in morphology and specific surface area, remain confined within a narrow nanoscale range. More investigation of Pt-MCO3 (M = Mg, Ca, Ba) samples is performed by HRTEM, HAADF-STEM, and EDS mapping. For Pt-BaCO3 (Figure 2a–e), lattice-resolved TEM reveals highly dispersed Pt species, as evidenced by well-defined 0.21 nm fringes corresponding to the Pt (111) plane. In contrast, Pt-CaCO3 (Figure 2f–j) exhibits distorted lattice fringes with a spacing of 0.24 nm [6,24]. The observed bending and local distortion indicate a significantly larger Pt grain size with pronounced lattice strain [25]. The observed bending and local distortion suggest the presence of larger Pt domains and indicate a higher degree of lattice strain at the Pt-CaCO3 interface. All catalysts exhibit highly dispersed Pt species, as evidenced by the homogeneous Pt distribution in EDS mapping, with minimal variation in dispersion across different samples. The Pt remains uniformly anchored on the support surfaces, primarily existing as small nanoparticles or sub-nanometric clusters without significant differences among the catalysts.

2.2. The Catalytic Performance Tests

The catalytic performance of Pt catalysts supported on alkaline earth carbonates (MgCO3, CaCO3, BaCO3) for low-temperature propylene combustion was systematically evaluated through temperature programmed oxidation experiments. As depicted in Figure 3a, distinct support effects manifested in the light-off profiles, as follows: Pt-BaCO3 demonstrated exceptional low-temperature activity with catalytic ignition initiating at 140 °C, achieving complete conversion by 192 °C. Comparatively, Pt-MgCO3 exhibited a higher activation threshold, requiring gradual temperature elevation to 247 °C for full conversion. The Pt-CaCO3 system showed the most sluggish kinetics, initiating conversion at 180 °C but necessitating heating to 282 °C for complete oxidation. These results conclusively establish the following support-dependent activity hierarchy: Pt-BaCO3 > Pt-MgCO3 > Pt-CaCO3, highlighting the critical role of support composition in modulating the metal–support interaction and redox properties.
To systematically compare the propylene catalytic combustion performance of three catalysts, catalytic reaction rates under different temperature conditions were also systematically tested. As shown in Figure 3b, after eliminating internal/external diffusion effects, the Pt-BaCO3 catalyst demonstrated a propylene combustion rate of 370 mmol·g−1·s−1 at 200 °C, representing 10-fold enhancements over Pt-CaCO3 and Pt-MgCO3, respectively, further confirming its catalytic superiority. While all three catalysts showed significant activity improvement with increasing temperature, Pt-BaCO3 maintained a 10 times rate advantage. Activation energy measurements revealed that the Pt-BaCO3 system possessed the lowest energy barrier (41.7 kJ·mol−1), substantially lower than those of Pt-CaCO3 (65.1 kJ·mol−1) and Pt-MgCO3 (53.2 kJ·mol−1). (Figure 3c) These experimental findings conclusively demonstrate that Pt-BaCO3 exhibits optimal propylene catalytic combustion performance among carbonate-supported catalysts.
Complementary GHSV experiments (20,000–160,000 mL·g−1·h−1) on the optimal Pt-BaCO3 catalyst (Figure 3d) revealed robust operational stability. While incremental flow rate increases caused moderate activity attenuation, the system exhibited remarkable process tolerance above 80,000 mL·g−1·h−1, where conversion rates stabilized within <5% variation despite the quadrupled space velocity. This threshold behavior suggests practical applicability in high-throughput industrial settings. Meanwhile, catalyst durability was also rigorously assessed through extended time-on-stream (50 h) and thermal cycling (3 cycles) experiments under harsh reaction conditions. Continuous operation at 190 °C (Figure 3e) revealed remarkable stability with propylene conversion maintaining stable conversion rates between 85–95% throughout the 50 h duration, without detectable deactivation. Thermal stress testing through repeated heating–cooling cycles (150–200 °C) demonstrated exceptional structural robustness, as evidenced by the superimposable conversion profiles in Figure 3f. Each cycle consistently achieved >90% conversion at 190 °C and complete oxidation at 200 °C, with less than 2% variation in T50 values between cycles. This excellent stability indicates that under operating conditions, common deactivation mechanisms, such as active phase sintering or carbonate carrier decomposition, are effectively alleviated in Pt-BaCO3.

2.3. The Reaction Path in Pt/MCO3

To elucidate the divergent catalytic activities, we performed systematic kinetic analyses of the above catalysts. As shown in Figure 4a under fixed propylene partial pressure (1 kPa), the oxygen reaction orders for propane combustion rates were determined as 0.98 (Pt-BaCO3), 1.08 (Pt-MgCO3), and 1.61 (Pt-CaCO3), respectively. The lower oxygen reaction order of Pt-BaCO3 and Pt-MgCO3 compared to Pt-CaCO3 suggested that the influence of O2 partial pressure to the combustion rate in the former two catalysts was weaker than Pt-CaCO3 (Figure 4a). Conversely, the propylene reaction orders exhibited negative values (−0.36 for Pt-BaCO3, −0.42 for Pt-MgCO3, −0.56 for Pt-CaCO3), indicating stronger propylene adsorption on Pt-CaCO3 relative to the other catalysts (Figure 4b) [26,27]. The observed positive oxygen reaction orders coupled with negative substrate dependencies collectively support a Mars–van Krevelen (M-K) redox mechanism, where adsorbed hydrocarbons react with lattice oxygen species during the catalytic combustion process.
18O2 isotopic labeling experiments unambiguously demonstrate the prevalence of the Mars–van Krevelen (M-K) mechanism in propylene catalytic combustion. Time-resolved isotopic tracking (Figure 4c) reveals distinct oxygen participation kinetics on Pt-BaCO3; pre-adsorbed surface O16 species are completely consumed within 30 min under He flow, while subsequent C3H6/O182/He exposure initially generates CO442 (from lattice O16) with progressively increasing intensity. The absence of CO462 and CO482 signals provides definitive evidence for the M-K pathway [4]. Isotopic kinetic analysis (Figure 4d) quantifies the rate-determining role of lattice oxygen regeneration, showing a twofold enhancement in reaction rates with O162 versus O182 at elevated temperatures, confirming oxygen diffusion as the rate-limiting step. This mechanistic understanding rationalizes the observed activity trend; strong propylene adsorption on Pt-CaCO3 induces competitive site blocking, simultaneously suppressing oxygen adsorption and compromising lattice oxygen replenishment, ultimately diminishing catalytic efficiency [28].
To validate the competitive adsorption mechanism, systematic O2-TPD and C3H6-TPD analyses were performed. As illustrated in Figure 5a, the O2-TPD profile of Pt-BaCO3 exhibited a broad multi-shouldered desorption peak spanning 427–640 °C with a dominant feature at ~427 °C, indicating abundant surface oxygen species. In contrast, Pt-MgCO3 demonstrated a single sharp desorption peak shifted to higher temperatures (~440 °C), suggesting reduced reactivity of its surface lattice oxygen. Notably, Pt-CaCO3 showed negligible oxygen desorption, reflecting its poor oxygen activation capability [29,30]. Complementary C3H6-TPD results (Figure 5b) revealed minimal propylene desorption from Pt-BaCO3 and Pt-MgCO3, implying weak propylene adsorption. However, Pt-CaCO3 displayed a prominent C3H6 desorption peak centered at ~411 °C. These TPD findings align with kinetic analyses, confirming that the propylene combustion pathway involves adsorbed C3H6 reacting with surface lattice oxygen. The superior performance of Pt-BaCO3 correlates with its optimal combination of low C3H6 surface coverage and enhanced oxygen activation capacity [31].

2.4. Understanding the Structure–Performance Relationship in Pt/MCO3

Given the poor lattice oxygen reactivity of carbonate-supported catalysts, the active lattice oxygen in Pt-BaCO3 and Pt-CaCO3 likely originates from PtO2 or PtO species. To verify this hypothesis, a comprehensive XPS analysis was conducted to probe the electronic states of Pt. As shown in Figure 6a, all three catalysts exhibited characteristic Pt4+ and Pt2+ signatures, with binding energies at 72.5 eV and 74.8 eV corresponding to the d7/2 orbitals of Pt2+ and Pt4+, respectively, confirming the predominantly oxidized state of Pt on carbonate supports. Notably, distinct Pt2+/Pt4+ ratios were observed across different carriers, following the sequence Pt-CaCO3 (0.60) > Pt-BaCO3 (0.39) > Pt-MgCO3 (0.31), revealing the highest proportion of Pt2+ species in Pt-CaCO3. These findings suggest that Pt2+ species may facilitate propylene adsorption, while Pt4+ ions likely correspond to reactive surface lattice oxygen species critical for catalytic oxidation.
While the Pt2+/Pt4+ ratio correlates with variations in propylene adsorption energy and lattice oxygen reactivity within Pt-MCO3 catalytic systems, the subtle differences in these ratios fail to account for the distinct properties between Pt-BaCO3 and Pt-MgCO3. Subsequently, H2-TPR analyses were performed to investigate the reducibility of PtO2 and PtO species. As shown in Figure 6b, only Pt-BaCO3 displays a distinct hydrogen consumption peak around 100 °C, whereas Pt-CaCO3 and Pt-MgCO3 show negligible reduction signals in this range, indicating divergent metal–support interactions that render PtO2/PtO species on CaCO3 or MgCO3 more reducible compared with those on BaCO3. This electronic modulation likely originates from interfacial charge transfer mechanisms.
To gain atomic-level insights into the electronic interactions between Pt and alkaline earth carbonate supports, density functional theory (DFT) calculations were performed for PtO2 clusters anchored on the (104) facets of CaCO3, MgCO3, and BaCO3. As illustrated by the charge density difference plots and Mulliken population analysis (Figures S5–S7), significant electron donation from the carbonate supports to Pt, with the magnitude of electron accumulation on Pt following the order MgCO3 (+0.52 e) > CaCO3 (+0.42 e) > BaCO3 (+0.39 e), demonstrating the support-dependent electronic modulation of Pt nanoparticles. The diminished electron donation capacity of Ba2+, arising from its low electronegativity and highly ionic nature, accounts for its inefficient charge transfer to Pt in comparison to Ca2+ and Mg2+. This electronically inert environment at the Pt-BaCO3 interface gives rise to weak electrostatic interactions, thereby favoring the stabilization of Pt species in a highly dispersed and oxidized state [32,33]. Consequently, BaCO3 demonstrates a superior ability to maintain ultrasmall Pt structures, which can be rationalized by its intrinsically weaker metal–support interaction with Pt [34,35].
Although the Pt valence states are initially determined under static conditions via XPS analysis, the inherently reductive nature of Pt species suggests that the oxidative propylene reaction environment may lead to dynamic valence changes (Figure 5b). To probe this possibility, in situ CO desorption infrared spectroscopy is employed, where characteristic CO binding configurations are used to distinguish Pt oxidation states based on their distinct chemisorption patterns (Figure 7). Typically, Ptδ+ species possess fewer d-electrons than Pt0, which limits effective d–π* back-donation, resulting in weaker CO adsorption and activation. Accordingly, high-frequency bands above 2090 cm−1 are assigned to CO adsorbed on Pt4+ sites, while the lower-frequency region (2088–2068 cm−1) corresponds to CO bound to Pt0. Notably, Pt–BaCO3 shows a much stronger signal at 2108 cm−1, corresponding to CO adsorbed on Pt4+ sites, indicating a higher proportion of Pt4+, consistent with the XPS results. In addition, Pt-CaCO3 displays a distinct band at 2016 cm−1, attributable to bridged CO, suggesting Pt aggregation and the formation of larger particles, in contrast to the more uniform dispersion of Pt species observed in Pt-BaCO3 and Pt-MgCO3 [36]. This observation aligns with the particle size trend obtained from CO pulse chemisorption. Moreover, compared to the other two catalysts, the CO adsorption band on Pt-BaCO3 is shifted to a higher wavenumber, reflecting weaker Pt-CO bonding and lower electron density around Pt, further supporting the presence of weaker metal–support interactions (MSIs) in the Pt-BaCO3 system.
To probe the evolution of surface intermediates under reaction conditions, Pt-BaCO3 is exposed to a C3H6/O2 mixture in a DRIFTS cell (Figure 8). In situ IR spectroscopy reveals the formation of carbonyl-containing intermediates on the catalyst surface. As temperature increases, a distinct band at 1745 cm−1 emerges, assigned to the C=O stretch of oxidation products (e.g., acrolein or ketones). Bands at 1619 and 1591 cm−1 are attributed to C2 enolates, while signals at 1577, 1560, 1545, 1408, and 1200–1400 cm−1 correspond to surface-stabilized C1 (carbonates/formates) and C2 (carboxylates) species, indicating progressive oxygenate formation. A minor increase in the 2800–3000 cm−1 region suggests σ-bonded alcohols. These results demonstrate that Pt-BaCO3 facilitates selective propylene oxidation to unsaturated aldehydes/ketones (e.g., acrolein, acetone), which further convert to C2/C1 species prior to final oxidation to CO2 and H2O [37,38].
Combining the results of kinetic studies, isotopic labeling, XPS characterization, and in situ FT-IR spectroscopy reveals that the catalytic combustion of propylene over carbonate-supported catalysts follows a classical M-K mechanism. Among these systems, the BaCO3 support exhibits the weakest interaction with PtO2, as evidenced by its limited electron transfer capacity and the higher oxidation state of Pt species (as illustrated in Figure 6a). These characteristics synergistically lead to (i) the weakest propylene adsorption on active sites and (ii) highly reactive lattice oxygen species. The concerted effects of these two factors collectively contribute to the superior catalytic performance observed on Pt-BaCO3.

3. Experimental

3.1. Preparation of Catalysts

The Pt-MCO3 catalysts (M = Mg, Ca, Ba) with 1 wt.% Pt loading were synthesized via an impregnation method [39,40]. Specifically, 1.0 g of MCO3 support was dispersed in 40 mL deionized water to form a slurry. Subsequently, 0.067 g of Pt(NO3)2 solution (containing 14.94% Pt) was added dropwise to the suspension under continuous stirring to achieve the designated platinum loading. The mixture was continuously stirred at room temperature for 6 h to ensure homogeneous dispersion of the active component, followed by 12 h of drying under ambient conditions. The dried precursors were then calcined in static air at 400 °C for 3 h with a heating rate of 2 °C/min [41].

3.2. Catalyst Performance Evaluation

The catalytic activity of the catalysts was evaluated in a fixed-bed quartz reactor (8 mm in diameter and 40 cm in length). To mitigate thermal effects and improve heat dissipation during the reaction, 30 mg of catalyst powder (40–60 mesh) was mixed with 100 mg of quartz sand (40–60 mesh) [28]. The reactant gas flow rate was controlled using a mass flow controller at 10 mL/min. The reactant mixture consisted of 1 vol% C3H6, 20 vol% O2, 79 vol% N2, with a gas hourly space velocity (GHSV) maintained at 20,000 mL·g−1·h−1. Product stream composition was continuously monitored by online gas chromatography (GC-9750, FULI Analytical Instruments, Wenling, China) equipped with a TDX-01 packed column and dual thermal conductivity detectors (TCD).
The C3H6 conversion was calculated using the following relationship:
X C 3 H 6 = C 3 H 6 inlet - C 3 H 6 outlet C 3 H 6 inlet   ×   100 %
where [C3H6]inlet and [C3H6]outlet represent the inlet and outlet concentrations of propylene, respectively [37].

3.3. Characterization of Catalysts

X-ray diffraction (XRD) patterns were acquired using a Bruker AXS D8 Focus diffractometer equipped (Karlsruhe, Germany) with Cu Kα radiation (λ = 0.15406 nm) operated at 40 kV and 40 mA. The morphologies of the samples were characterized using scanning electron microscopy (SEM, Nova Nano SEM 450, Waltham, MA, USA) and transmission electron microscopy (TEM, Nion UltraSTEM 100, Karlsruhe, Germany). For TEM analyses, the dry carbon samples were first dispersed in ethanol through ultrasonication to form homogeneous suspensions. Approximately 5 μL aliquots of these suspensions were then drop-cast onto TEM grids and allowed to dry under ambient conditions prior to imaging. X-ray photoelectron spectroscopy (XPS) measurements were conducted on a Thermo ESCALAB 250 spectrometer (Waltham, MA, USA) with a monochromatized Al Kα X-ray source (1486.6 eV). The analyzer pass energy was set to 20 eV for high-resolution scans. All binding energies were calibrated against the C1s peak of adventitious carbon (284.6 eV). Oxygen temperature programmed desorption (O2-TPD) was performed on a Micromeritics AutoChem 2920 II analyzer (Shanghai, China) coupled with a Hiden HPR 20 mass spectrometer (Warrington, UK). Approximately 50 mg of sample (40–60 mesh) was loaded into a quartz reactor and pretreated under He flow (30 mL/min) at 250 °C for 30 min. After cooling to room temperature, the sample was exposed to a 3 vol% O2/He mixture (30 mL/min) at 200 °C for 30 min (heating rate: 10 °C/min), followed by cooling to room temperature under He purge (20 mL/min). The desorption profile was recorded by heating the reactor to 600 °C at 10 °C min−1 under He flow (20 mL/min), with m/z = 32 (O2) monitored by mass spectrometry. H2 temperature programmed reduction (H2-TPR) experiments were carried out using conventional flow system equipment with a thermal conductivity detector (TCD). A 30 mg sample was loaded into the quartz tube reactor. The reduction gas consisted of 5% H2/N2 (40 mL/min). TPR was run from 30 °C to 450 °C at the rate of 5 °C/min.

3.4. Isotope-Labeling Experiments

The isotopic labeling experiments for the C3H6 + 18O2 continuous reactions were monitored by real-time mass spectrometry. Prior to testing, all samples underwent pretreatment at 300 °C for 60 min under 5 vol% 16O2/He flows, followed by He purging to eliminate residual 16O2. Upon cooling the reaction beds to the target temperatures, the carrier gases were immediately switched to the reaction mixtures (20 vol% C3H6 + 5048 ppm 18O2 balanced with He) while heating to 280 °C at 2 °C/min for 1 h, with the effluents directly analyzed by mass spectrometry. Reaction products were continuously quantified using online mass spectrometers (Thermo Fisher Delta V, Waltham, MA, USA) by tracking the characteristic signals of C16O2 (m/z = 44), C16O18O (m/z = 46), and C18O2 (m/z = 48).

3.5. Computational Method

All density functional theory (DFT) calculations were performed using CASTEP with the projector augmented wave method. The exchange–correlation effects were treated with the Perdew–Burke–Ernzerhof functional. After rigorous convergence tests, we employed plane-wave cutoff energies of 450 eV and 1 × 1 × 1 k-point meshes for Brillouin zone sampling. Vacuum layers of 15 Å were applied along the z-axes of the slab models to eliminate periodic interactions, with convergence criteria set at 10−5 eV for energies and 0.02 eV Å−1 for maximum stresses [42,43].

4. Conclusions

In summary, this study reveals a pronounced structure–composition correlation in alkaline earth metal carbonate-supported Pt catalysts for propylene combustion, with BaCO3 identified as the optimal support. Notably, the propylene combustion rate over Pt/BaCO3 surpasses those of Pt/CaCO3 and Pt/MgCO3 by an order of magnitude. Through kinetic studies and 18O2 isotope labeling experiments, we demonstrate that while all Pt-MCO3 systems follow the Mars–van Krevelen (M-K) mechanism, there are distinct differences emerge in propylene adsorption and activation pathways. The weakened metal–support interaction in Pt/BaCO3 results in minimal electron transfer from the BaCO3 support to PtO2, rendering the Pt4+ centers in Pt/BaCO3 the most electrophilic yet least capable of propylene chemisorption. Intriguingly, this electronic configuration modification effectively prevents poisoning of Pt active sites, while stabilizing Pt species in higher oxidation states, thereby dramatically enhancing the catalytic combustion performance. These findings provide fundamental insights into the structure–performance relationships of Pt/MCO3 catalysts, establishing a framework for the rational design of advanced metal carbonate-supported catalytic systems.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal15080696/s1, Figure S1: SEM images of Pt-BaCO3.; Figure S2: SEM images of Pt-CaCO3; Figure S3: SEM images of Pt-MgCO3; Figure S4: TG curves of Pt-MCO3 (M=Mg, Ca, Ba) catalysts; Figure S5: Mulliken population analysis of Pt-CaCO3; Figure S6: Mulliken population analysis of Pt-MgCO3; Figure S7: Mulliken population analysis of Pt-BaCO3.

Author Contributions

X.S.: writing—review and editing, methodology, data curation, conceptualization. Y.S.: writing—review and editing, methodology, data curation, conceptualization. Y.L.: writing—review and editing, writing—original draft, visualization, formal analysis, data curation, conceptualization. Y.G.: writing—review and editing, supervision, funding acquisition, conceptualization. P.Z.: writing—review and editing, supervision, methodology, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article/supplementary material. Further inquiries can be directed to the corresponding author(s).

Acknowledgments

This work was supported by the National Natural Science Foundation of China (grant no. U24A20529, 22308172, 22178219), Natural Science Foundation of Ningxia Province [grant number 2023BSB03068], National Key R & D Program Plan (2022YFA1504803), Inner Mongolia R &D Program Plan (2021ZD0042, 2021EEDSCXSFQZD006, 2021GG0350), Ordos R&D Program (2121HZ231-8), The Central Guidance for Local Scientific and Technological Development Funds of China (No. 2022FRD05017).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, H.; Wang, Z.; Wei, L.; Liu, Y.; Dai, H.; Deng, J. Recent progress on VOC pollution control via the catalytic method. Chin. J. Catal. 2024, 61, 71–96. [Google Scholar] [CrossRef]
  2. Yang, C.; Miao, G.; Pi, Y.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Abatement of various types of VOCs by adsorption/catalytic oxidation: A review. Chem. Eng. J. 2019, 370, 1128–1153. [Google Scholar] [CrossRef]
  3. Zhang, B.; Yang, J.; Mu, Y.; Ji, X.; Cai, Y.; Jiang, N.; Xie, S.; Qian, Q.; Liu, F.; Tan, W.; et al. Fabrication of Highly Dispersed Ru Catalysts on CeO2 for Efficient C3H6 Oxidation. Environ. Sci. Technol. 2024, 58, 19533–19544. [Google Scholar] [CrossRef] [PubMed]
  4. Hosokawa, S.; Shibano, T.; Koga, H.; Matsui, M.; Asakura, H.; Teramura, K.; Okumura, M.; Tanaka, T. Excellent Catalytic Activity of a Pd-Promoted MnOx Catalyst for Purifying Automotive Exhaust Gases. ChemCatChem 2020, 12, 4276–4280. [Google Scholar] [CrossRef]
  5. Chenglin, S.; Yonghui, B.; Guangsuo, Y.; Xiaqing, L.; Li, M.; Pengcheng, F.; Fang, L. Co-combustion Characteristic and Gas Emission Characteristic of Entrained Flow Gasification Fine Slag Residual Carbon and Coal. J. Ningxia Univ. (Nat. Sci. Ed.) 2024, 45, 282–289. [Google Scholar] [CrossRef]
  6. Lee, S.; Lin, C.; Kim, S.; Mao, X.; Kim, T.; Kim, S.-J.; Gorte, R.J.; Jung, W. Manganese Oxide Overlayers Promote CO Oxidation on Pt. ACS Catal. 2021, 11, 13935–13946. [Google Scholar] [CrossRef]
  7. Park, J.; Kim, D.; Byun, S.W.; Shin, H.; Ju, Y.; Min, H.; Kim, Y.J.; Heo, I.; Hazlett, M.J.; Kim, M.; et al. Impact of Pd:Pt ratio of Pd/Pt bimetallic catalyst on CH4 oxidation. Appl. Catal. B Environ. Energy 2022, 316, 121623. [Google Scholar] [CrossRef]
  8. Liccardo, G.; Cendejas, M.C.; Mandal, S.C.; Stone, M.L.; Porter, S.; Nhan, B.T.; Kumar, A.; Smith, J.; Plessow, P.N.; Cegelski, L.; et al. Unveiling the Stability of Encapsulated Pt Catalysts Using Nanocrystals and Atomic Layer Deposition. J. Am. Chem. Soc. 2024, 146, 23909–23922. [Google Scholar] [CrossRef]
  9. Zhang, B.; Yang, Y.; Zheng, J.; Zhang, D.; Chen, W.; Yuan, W.; Chen, X.; Liu, R.; Chen, B.; Li, L.; et al. Diverse Effects of SO2-Induced Pt–O–SO3 on the Catalytic Oxidation of C3H6 and C3H8. Environ. Sci. Technol. 2024, 58, 18020–18032. [Google Scholar] [CrossRef]
  10. You, Y.; Xu, A.; Tang, X.; Guo, Y.; Zhan, W.; Wang, L.; Dai, S.; Guo, Y. Refining Metal–Support Interactions via Surface Modification of Irreducible Oxide Support for Enhanced Complete Propane Oxidation. ACS Catal. 2024, 14, 11457–11467. [Google Scholar] [CrossRef]
  11. An, Y.; Chen, S.-Y.; Wang, B.; Zhou, L.; Hao, G.; Wang, Y.; Chen, J.; Tsung, C.-K.; Liu, Z.; Chou, L.-Y. Controlling the metal–support interaction with steam-modified ceria to boost Pd activity towards low-temperature CO oxidation. J. Mater. Chem. A 2023, 11, 16838–16845. [Google Scholar] [CrossRef]
  12. Fang, Q.; Du, H.; Zhang, X.; Ding, Y.; Zhang, Z.C. Alteration of Hydrodeoxygenation Pathways of Ni/TiO2-A Catalyst through Controlled Regulation of Strong Metal–Support Interactions and Surface Acidity. ACS Catal. 2024, 14, 5047–5063. [Google Scholar] [CrossRef]
  13. Song, Y.; Beyazay, T.; Tüysüz, H. Effect of Alkali- and Alkaline-Earth-Metal Promoters on Silica-Supported Co−Fe Alloy for Autocatalytic CO2 Fixation. Angew. Chem. Int. Ed. 2023, 63, e202316110. [Google Scholar] [CrossRef] [PubMed]
  14. Xie, Z.; Sun, Z.; Shao, B.; Zhu, Y.; Ma, R.; Li, S.; Li, J.; Chen, Y.; Liu, H.; Hu, J. Highly efficient hydrogenation of carbonate to methanol for boosting CO2 mitigation. Chem. Eng. J. 2024, 495, 153465. [Google Scholar] [CrossRef]
  15. Naicheng, L.; Mingxiao, M.; Mingiie, M. Research on lnfluencing Factors af Early Strength of CaO Fly Ash Geopolymer. J. Ningxia Univ. (Nat. Sci. Ed.) 2022, 43, 378–382. [Google Scholar]
  16. Zhang, W.; Li, D.; Liu, C.; Jiang, Z.; Gao, C.; Shen, L.; Liu, Q.; Qiu, R.; Gu, H.; Lian, C.; et al. Facilitated CO2 hydrogenation by strong metal-support interaction between Ni and BaCO3. Chem Catal. 2024, 4, 101113. [Google Scholar] [CrossRef]
  17. Xiong, P.; Xu, Z.; Wu, T.-S.; Yang, T.; Lei, Q.; Li, J.; Li, G.; Yang, M.; Soo, Y.-L.; Bennett, R.D.; et al. Synthesis of core@shell catalysts guided by Tammann temperature. Nat. Commun. 2024, 15, 420. [Google Scholar] [CrossRef]
  18. Wang, P.; Shi, M.; Shu, Y.; Niu, Q.; Zhang, P. Porous CaCO3 fabricated by Ca-containing solid wastes for effective CO2 adsorption. J. Environ. Chem. Eng. 2025, 13, 115815. [Google Scholar] [CrossRef]
  19. Ren, H.; Yu, W.; Lv, M.; Gao, J.; Hu, H.; Wang, M.; Cui, X.; Liu, J.; Jiang, L. Carbon-Encapsulated Single-Atom Platinum Nickel Alloy for Efficient and Durable Alkaline Hydrogen Oxidation Through Enhanced Charge Polarization. Adv. Funct. Mater. 2024, 35, 2413754. [Google Scholar] [CrossRef]
  20. Hou, Z.; Sun, Z.; Cui, C.; Zhu, D.; Yang, Y.; Zhang, T. Ru Coordinated ZnIn2S4 Triggers Local Lattice-Strain Engineering to Endow High-Efficiency Electrocatalyst for Advanced Zn-Air Batteries. Adv. Funct. Mater. 2022, 32, 2110572. [Google Scholar] [CrossRef]
  21. Ma, X.; Yang, T.; He, D.; Gao, X.; Jiang, W.; Li, D.; Sun, Y.; Lin, X.; Xu, J.; Wang, H.; et al. Carbonate shell regulates CuO surface reconstruction for enhanced CO2 electroreduction. Nat. Synth. 2024, 4, 53–66. [Google Scholar] [CrossRef]
  22. Zhu, H.; Lv, X.; Wu, Y.; Wang, W.; Wu, Y.; Yan, S.; Chen, Y. Carbonate-carbonate coupling on platinum surface promotes electrochemical water oxidation to hydrogen peroxide. Nat. Commun. 2024, 15, 8846. [Google Scholar] [CrossRef]
  23. Cui, H.; Cheng, Z.; Zhou, Z. Unravelling the role of alkaline earth metal carbonates in intermediate temperature CO2 capture using alkali metal salt-promoted MgO-based sorbents. J. Mater. Chem. A 2020, 8, 18280–18291. [Google Scholar] [CrossRef]
  24. Khan, H.A.; Natarajan, P.; Jung, K.-D. Stabilization of Pt at the inner wall of hollow spherical SiO2 generated from Pt/hollow spherical SiC for sulfuric acid decomposition. Appl. Catal. B Environ. Energy 2018, 231, 151–160. [Google Scholar] [CrossRef]
  25. Xinxin, H.; Lijun, H.; Yanming, D.; Jiankang, Z.; Ting, L.; Yaojun, G.; Mingze, S. Numerical Analysis of the Influence of Residual Stress and Grain Size on the Indentation Behavior of Beryllium. J. Ningxia Univ. (Nat. Sci. Ed.) 2024, 45, 31–35. [Google Scholar]
  26. Lachkov, P.T.; Chin, Y.-H. Catalytic consequences of reactive oxygen species during C3H6 oxidation on Ag clusters. J. Catal. 2018, 366, 127–138. [Google Scholar] [CrossRef]
  27. Hazlett, M.J.; Moses-Debusk, M.; Parks, J.E.; Allard, L.F.; Epling, W.S. Kinetic and mechanistic study of bimetallic Pt-Pd/Al2O3 catalysts for CO and C3H6 oxidation. Appl. Catal. B Environ. Energy 2016, 202, 404–417. [Google Scholar] [CrossRef]
  28. Shu, Y.; Wang, M.; Duan, X.; Liu, D.; Yang, S.; Zhang, P. Low-temperature total oxidation of methane by pore- and vacancy-engineered NiO catalysts. AIChE J. 2022, 68, e17664. [Google Scholar] [CrossRef]
  29. Bae, W.B.; Kim, D.Y.; Byun, S.W.; Lee, S.J.; Kuk, S.K.; Kwon, H.J.; Lee, H.C.; Hazlett, M.J.; Liu, C.; Kim, Y.J.; et al. Direct NO decomposition over Rh-supported catalysts for exhaust emission control. Chem. Eng. J. 2023, 475, 146005. [Google Scholar] [CrossRef]
  30. Xia, H.; Dang, C.; Zhou, D.; Cai, W. Lamellar cross-linking Ni/CeO2 as an efficient and durable catalyst for dry reforming of methane. Chem. Eng. J. 2024, 489, 151365. [Google Scholar] [CrossRef]
  31. Chai, Y.; Chen, S.; Chen, Y.; Wei, F.; Cao, L.; Lin, J.; Li, L.; Liu, X.; Lin, S.; Wang, X.; et al. Dual-Atom Catalyst with N-Colligated Zn1Co1 Species as Dominant Active Sites for Propane Dehydrogenation. J. Am. Chem. Soc. 2023, 146, 263–273. [Google Scholar] [CrossRef]
  32. Hu, X.; Xu, D.; Jiang, J. Strong Metal-Support Interaction between Pt and TiO2 over High-Temperature CO2 Hydrogenation. Angew. Chem. Int. Ed. 2024, 64, e202419103. [Google Scholar] [CrossRef]
  33. Shaodi, S.; Xiaomin, W.; Zhiwei, H.; Huazheng, S.; Huawang, Z.; Guohua, J. Engineering stable Pt nanoclusters on defective two-dimensional TiO2 nanosheets by introducing SMSI for efficient ambient formaldehyde oxidation. Chem. Eng. J. 2022, 435, 135035. [Google Scholar] [CrossRef]
  34. Haus, M.O.; Meledin, A.; Leiting, S.; Louven, Y.; Roubicek, N.C.; Moos, S.; Weidenthaler, C.; Weirich, T.E.; Palkovits, R. Correlating the Synthesis, Structure, and Catalytic Performance of Pt–Re/TiO2 for the Aqueous-Phase Hydrogenation of Carboxylic Acid Derivatives. ACS Catal. 2021, 11, 5119–5134. [Google Scholar] [CrossRef]
  35. Guo, M.; Meng, Q.; Gao, M.-L.; Zheng, L.; Li, Q.; Jiao, L.; Jiang, H.-L. Single-Atom Pt Loaded on MOF-Derived Porous TiO2 with Maxim-Ized Pt Atom Utilization for Selective Hydrogenation of Halonitro-benzene. Angew. Chem. Int. Ed. 2024, 64, e202418964. [Google Scholar] [CrossRef] [PubMed]
  36. Fang, Y.; Li, H.; Zhang, Q.; Wang, C.; Xu, J.; Shen, H.; Yang, J.; Pan, C.; Zhu, Y.; Luo, Z.; et al. Oxygen Vacancy-Governed Opposite Catalytic Performance for C3H6 and C3H8 Combustion: The Effect of the Pt Electronic Structure and Chemisorbed Oxygen Species. Environ. Sci. Technol. 2022, 56, 3245–3257. [Google Scholar] [CrossRef] [PubMed]
  37. Bi, S.; Tian, X.; Shu, Y.; Zhang, Z.; Liu, Q.; Zhang, P.; Luo, Z. Ideal Co-NiO solid solution with excellent low-temperature propene catalytic combustion performance. Chem. Eng. Sci. 2024, 302, 120923. [Google Scholar] [CrossRef]
  38. Liu, P.; Zheng, C.; Liu, W.; Wu, X.; Liu, S. Oxidative Redispersion-Derived Single-Site Ru/CeO2 Catalysts with Mobile Ru Complexes Trapped by Surface Hydroxyls Instead of Oxygen Vacancies. ACS Catal. 2024, 14, 6028–6044. [Google Scholar] [CrossRef]
  39. Yang, H.; Wang, X.; Xie, Y.-D.; Lu, A.-H. Evolution of the Active Phase of Pt/Sn–Al2O3 Catalysts During Acidic Impregnation and Their Use in Propane Dehydrogenation. ACS Appl. Mater. Interfaces 2024, 16, 47773–47783. [Google Scholar] [CrossRef]
  40. Wang, X.; Wang, X.; Yu, H.; Wang, X. The functions of Pt located at different positions of HZSM-5 in H2-SCR. Chem. Eng. J. 2018, 355, 470–477. [Google Scholar] [CrossRef]
  41. Xiao, Z.; Lin, X.; Feng, W.; Chen, B.; Meng, Q.; Wang, T. Efficient Hydrogen Production from the Aqueous-Phase Reforming of Biomass-Derived Oxygenated Hydrocarbons over an Ultrafine Pt Nanocatalyst. Catalysts 2023, 13, 1428. [Google Scholar] [CrossRef]
  42. Singh Pal, R.; Rana, S.; Kumar Sharma, S.; Khatun, R.; Khurana, D.; Suvra Khan, T.; Kumar Poddar, M.; Sharma, R.; Bal, R. Enhancement of oxygen vacancy sites of La2-xMxCe2O7-δ (M = Ca, Ba, Sr) catalyst for the low temperature oxidative coupling of Methane: A combined DFT and experimental study. Chem. Eng. J. 2023, 458, 141379. [Google Scholar] [CrossRef]
  43. Xiaohui, R.; Fei, L.; Qi, W.; Hui, S.; Shunqin, L.; Sijie, L.; Gaoliang, Y.; Bowen, D.; Zongyu, H.; Xu-Sheng, W.; et al. Engineering interfacial charge transfer channel for efficient photocatalytic H2 evolution: The interplay of CoPx and Ca2+ dopant. Appl. Catal. B Environ. Energy 2021, 303, 120887. [Google Scholar] [CrossRef]
Figure 1. Structural characterization of Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (a) XRD patterns of Pt-MCO3 (M = Ba, Mg, Ca) catalysts. (b) Raman spectra of Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (c) N2 adsorption and desorption curves and the pore size distribution of the Pt-MCO3 (M = Ba, Mg, Ca) catalysts. (d) CO pulse chemisorption profiles at room temperature.
Figure 1. Structural characterization of Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (a) XRD patterns of Pt-MCO3 (M = Ba, Mg, Ca) catalysts. (b) Raman spectra of Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (c) N2 adsorption and desorption curves and the pore size distribution of the Pt-MCO3 (M = Ba, Mg, Ca) catalysts. (d) CO pulse chemisorption profiles at room temperature.
Catalysts 15 00696 g001
Figure 2. (ac) TEM images of Pt-BaCO3. (d) HAADF-STEM image of Pt-BaCO3. (e) The corresponding EDS mapping of Pt-BaCO3. (fh) TEM images of Pt-CaCO3. (i) HAADF-STEM image of Pt-CaCO3. (j) The corresponding EDS mapping of Pt-CaCO3.
Figure 2. (ac) TEM images of Pt-BaCO3. (d) HAADF-STEM image of Pt-BaCO3. (e) The corresponding EDS mapping of Pt-BaCO3. (fh) TEM images of Pt-CaCO3. (i) HAADF-STEM image of Pt-CaCO3. (j) The corresponding EDS mapping of Pt-CaCO3.
Catalysts 15 00696 g002
Figure 3. The performance of Pt-MCO3 (M = Mg, Ca, Ba) catalysts in C3H6 combustion. (a) Light-off curves of C3H6 combustion over the Pt-MCO3 samples. (b) C3H6 combustion rates in the Pt-MCO3 catalysts. (c) The C3H6 catalytic combustion activation energies of Pt-MCO3 catalysts. (d) Light-off curves of C3H6 combustion over the Pt-BaCO3 at WHSV = 20,000 mL·g−1·h−1, 40,000 mL·g−1·h−1, 80,000 mL·g−1·h−1, and 160,000 mL·g−1·h−1. (e) Stability test on Pt-BaCO3. (f) Light-off curves of C3H6 combustion over Pt-BaCO3 in three consecutive runs. All the reactions were kept at WHSV = 20,000 mL·g−1·h−1.
Figure 3. The performance of Pt-MCO3 (M = Mg, Ca, Ba) catalysts in C3H6 combustion. (a) Light-off curves of C3H6 combustion over the Pt-MCO3 samples. (b) C3H6 combustion rates in the Pt-MCO3 catalysts. (c) The C3H6 catalytic combustion activation energies of Pt-MCO3 catalysts. (d) Light-off curves of C3H6 combustion over the Pt-BaCO3 at WHSV = 20,000 mL·g−1·h−1, 40,000 mL·g−1·h−1, 80,000 mL·g−1·h−1, and 160,000 mL·g−1·h−1. (e) Stability test on Pt-BaCO3. (f) Light-off curves of C3H6 combustion over Pt-BaCO3 in three consecutive runs. All the reactions were kept at WHSV = 20,000 mL·g−1·h−1.
Catalysts 15 00696 g003
Figure 4. The kinetic analysis of Pt-MCO3 (M = Mg, Ca, Ba) catalysts in C3H6 combustion. (a) Dependence of reaction rate on partial pressure of C3H6. (b) Dependence of reaction rate on partial pressure of O2 over Pt-MCO3. (c) Time-dependent mass spectra of CO162, CO16O18, and CO18O18 species during the C3H6 combustion reaction over Pt-BaCO3. (d) The C3H6 combustion reaction rates for Pt-BaCO3 were measured in the presence of O16 or O18.
Figure 4. The kinetic analysis of Pt-MCO3 (M = Mg, Ca, Ba) catalysts in C3H6 combustion. (a) Dependence of reaction rate on partial pressure of C3H6. (b) Dependence of reaction rate on partial pressure of O2 over Pt-MCO3. (c) Time-dependent mass spectra of CO162, CO16O18, and CO18O18 species during the C3H6 combustion reaction over Pt-BaCO3. (d) The C3H6 combustion reaction rates for Pt-BaCO3 were measured in the presence of O16 or O18.
Catalysts 15 00696 g004
Figure 5. The characterization of TPD in Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (a) O2-TPD profiles of Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (b) C3H6-TPD profiles of Pt-MCO3 (M = Mg, Ca, Ba) catalysts.
Figure 5. The characterization of TPD in Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (a) O2-TPD profiles of Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (b) C3H6-TPD profiles of Pt-MCO3 (M = Mg, Ca, Ba) catalysts.
Catalysts 15 00696 g005
Figure 6. The characterization of Pt species in Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (a) Pt 4f XPS profiles of Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (b) H2-TPR profiles of Pt-MCO3 (M = Mg, Ca, Ba) catalysts.
Figure 6. The characterization of Pt species in Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (a) Pt 4f XPS profiles of Pt-MCO3 (M = Mg, Ca, Ba) catalysts. (b) H2-TPR profiles of Pt-MCO3 (M = Mg, Ca, Ba) catalysts.
Catalysts 15 00696 g006
Figure 7. The characterization of in situ FTIR spectra of CO desorption in (a) Pt-CaCO3 (b) Pt-MaCO3 (c) Pt-BaCO3.
Figure 7. The characterization of in situ FTIR spectra of CO desorption in (a) Pt-CaCO3 (b) Pt-MaCO3 (c) Pt-BaCO3.
Catalysts 15 00696 g007
Figure 8. The characterization of in situ FTIR spectra under reaction conditions in Pt-BaCO3 ((a): 3000 cm−1 to 2800 cm−1, (b): 1800 cm−1 to 1400 cm−1, (c): 1400 cm−1 to 1200 cm−1).
Figure 8. The characterization of in situ FTIR spectra under reaction conditions in Pt-BaCO3 ((a): 3000 cm−1 to 2800 cm−1, (b): 1800 cm−1 to 1400 cm−1, (c): 1400 cm−1 to 1200 cm−1).
Catalysts 15 00696 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sun, X.; Lv, Y.; Shu, Y.; Guo, Y.; Zhang, P. Alkaline Earth Carbonate Engineered Pt Electronic States for High-Efficiency Propylene Oxidation at Low Temperatures. Catalysts 2025, 15, 696. https://doi.org/10.3390/catal15080696

AMA Style

Sun X, Lv Y, Shu Y, Guo Y, Zhang P. Alkaline Earth Carbonate Engineered Pt Electronic States for High-Efficiency Propylene Oxidation at Low Temperatures. Catalysts. 2025; 15(8):696. https://doi.org/10.3390/catal15080696

Chicago/Turabian Style

Sun, Xuequan, Yishu Lv, Yuan Shu, Yanglong Guo, and Pengfei Zhang. 2025. "Alkaline Earth Carbonate Engineered Pt Electronic States for High-Efficiency Propylene Oxidation at Low Temperatures" Catalysts 15, no. 8: 696. https://doi.org/10.3390/catal15080696

APA Style

Sun, X., Lv, Y., Shu, Y., Guo, Y., & Zhang, P. (2025). Alkaline Earth Carbonate Engineered Pt Electronic States for High-Efficiency Propylene Oxidation at Low Temperatures. Catalysts, 15(8), 696. https://doi.org/10.3390/catal15080696

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop