Direct dehydrogenation of ethylbenzene over C60–Ni/SiO2 catalysts: mechanistic insight into C60 as a molecular promoter
Received
4th September 2025
, Accepted 21st October 2025
First published on 21st October 2025
Abstract
Extensive research in heterogeneous catalysis has highlighted the potential of non-noble metal catalysts for efficient alkane dehydrogenation. Although Ni species are widely employed for activating C–H bonds, their utilization as active catalysts for alkane dehydrogenation is limited by thermal sintering, coke deposition, and undesired side reactions that compromise the stability and selectivity of the catalyst. This work reports the synthesis of a C60-modified nickel-based catalyst (C60–Ni/SiO2), which was employed for the direct dehydrogenation (DDH) of ethylbenzene (EB) to styrene (ST). Owing to its tailored electronic structure, the C60–Ni/SiO2 catalyst reached an ST formation rate of 2.7 mmol g−1 h−1 while maintaining a selectivity exceeding 99.0%. XRD, Raman, TEM, and XPS characterization revealed that the C60 served as an electronic promoter, which decreased the electron density of Ni species without disturbing its crystalline structure. Such modulation of the electronic structure of Ni centers effectively suppresses the cracking side reactions and coke formation, thereby improving both selectivity and stability of the catalyst during EB DDH. The present work introduces a promising Ni-based catalyst for EB dehydrogenation, potentially offering prospects for developing advanced non-noble metal catalysts for alkene production via the DDH process.
1. Introduction
Styrene (ST) is considered a key raw material in the modern petrochemical industry and is widely utilized for producing polymers, such as polystyrene and butadiene rubber. It is estimated that global ST production exceeds 30 million tons annually, leading to more than 27 million tons of CO2 emissions each year.1 Presently, industrial ST production predominantly relies on the direct dehydrogenation of ethylbenzene (EB DDH) over Fe–K–Cr-based catalysts with co-feeding overheated steam. However, Fe–K–Cr catalysts exhibit several limitations, including the instability of active Fe3+ sites, loss and redistribution of potassium promoters, and rapid deactivation caused by coke deposition.2–4 Moreover, a significant excess of steam used in DDH normally results in substantial energy consumption.5 Therefore, the development of novel, active, and coke-resistant catalysts that can operate with reduced steam consumption or even under steam-free conditions is crucial for improving the energy efficiency of ST production.
Ni-based catalysts have demonstrated remarkable performance in diverse catalytic reactions, including steam reforming,6,7 the water–gas shift reaction,8,9 and biomass gasification.10 As a non-noble metal, Ni has great advantages in terms of low cost and ease of preparation. Recently, Ni-based catalysts have been extensively applied in methane reforming, demonstrating excellent C–H bond activation ability.11–14 However, Ni-based materials have been rarely applied as catalysts for EB DDH, since Ni species tend to promote both C–H and C–C bond activation due to their strong adsorption capacity of both EB and ST, which accelerates coke formation.15–17 As a result, the rapid catalyst deactivation accompanied by diminished ST selectivity is commonly observed. A major challenge in solving this problem is that Ni species exhibit high electron density, which facilitates the strong metal–π interactions with aromatic compounds such as benzene. These interactions significantly alter the electron density distribution within the aromatic ring, thereby reducing the activation barriers for undesirable side reactions, including ring opening, cracking, and polymerization.18,19 These parasitic reactions are acknowledged as primary pathways for the coke generation, which in turn deactivates the catalyst and diminishes both the efficiency and stability of the reaction system. Recent studies have shown increasing efforts in exploring the incorporation of metallic promoters or elemental doping as effective strategies to engineer the electronic structure of Ni catalysts, thereby enhancing their activity and stability. For instance, several successful examples using modified Ni-based catalysts, such as NiP,20 NiSn,21 and NiAu,17 have been reported to improve alkane dehydrogenation performance. However, these modified catalysts still face serious scientific challenges, including decreased catalytic performance caused by the loss of accessible Ni active sites and sustained coke accumulation after long-term operation.
C60 (fullerene), which was discovered in 1985 and produced on a large scale by the 1990s,22 has shown intriguing properties in diverse fields, including chemistry, physics, materials science, electronics, and medicine.23 C60 continues to attract considerable scientific interest, particularly due to its emerging roles in energy and catalytic applications.24–26 Owing to its highly symmetric, π-conjugated structure, C60 exhibits excellent electron mobility and low reorganization energy, facilitating the efficient electron transportation.27,28 Moreover, C60 serves as an exceptional electron acceptor, capable of accepting electrons from metal species due to its unique degenerate lowest unoccupied molecular orbital (LUMO) at low energy.29,30 These characteristics make it a promising candidate for modulating the electronic structure and catalytic performance of Ni centers.31–33 In contrast, other carbon-based materials such as graphene, carbon nanotubes (CNTs) and fullerene derivatives primarily act through physical dispersion or π–π interactions rather than direct electronic modulation. Graphene and CNTs possess extended π-electron systems that interact weakly with metal orbitals, thus exhibiting limited ability to alter the electron density of Ni atoms.34–36 C60 with its discrete electron-accepting capability can efficiently withdraw electrons from Ni and regulate the d-orbital occupancy, thereby optimizing the adsorption and dehydrogenation behavior of ethylbenzene on Ni sites. This unique electronic interaction distinguishes C60 from other carbonaceous modifiers and underpins its effectiveness as an electronic promoter in Ni-based dehydrogenation catalysts.
This work presents a novel and promising strategy to tailor the electronic properties of Ni-based catalysts through introducing C60 as the promoter. The hybrid catalyst (C60–Ni/SiO2), composed of C60 promoters and Ni active sites, not only promotes the C–H bond activation during EB DDH but also significantly suppresses carbon deposition. Moreover, comprehensive characterization and mechanistic studies reveal that the electron transfer from Ni active sites to C60 modulates the electronic environment of Ni active sites and effectively mitigates sintering and coke formation in EB DDH. To our knowledge, this work introduces C60 as the promoter to modify Ni-based catalysts for EB DDH for the first time, offering new avenues for designing high-performance Ni-based catalytic systems.
2. Experimental
2.1. Preparation of materials
The C60–Ni/SiO2 catalyst was prepared by dissolving 0.10 g of Ni(NO3)2·6H2O in deionized water, followed by impregnation onto 1.00 g of nano-sized SiO2. The resulting mixture was dried at 80 °C for 12 h to obtain the Ni/SiO2 precursor. Separately, 0.15 g of C60 was dispersed in 60 mL of toluene and sonicated for 30 min until the C60 was fully dispersed. Subsequently, the C60–toluene solution was added dropwise to the Ni/SiO2 precursor under continuous mechanical stirring. Then, the resulting mixture was stirred at 60 °C for 12 h. The obtained precipitate was separated by filtration, followed by sequential washing with deionized water and ethanol to remove residual inorganic salts, nitrate ions, and organic species such as toluene and unbound C60. The washed solid was dried under vacuum overnight at 80 °C, and subsequently reduced at 600 °C under a 10% H2/Ar gas flow for 3 h. The final sample was denoted as C60–Ni/SiO2. For comparison, C60/SiO2 and Ni/SiO2 were prepared following the same procedure while omitting either Ni or C60, respectively.
2.2. Characterization of materials
The morphology and elemental distribution were examined via high-resolution transmission electron microscopy (HR-TEM), and energy-dispersive X-ray spectroscopy (EDX) mapping was carried out on an FEI Talos F200S G2 instrument (FEI, USA) operating at 200 kV. XRD patterns were collected on a Bruker D8 ADVANCE diffractometer employing Cu Kα radiation to probe crystal structures. Nitrogen adsorption–desorption isotherms were carried out at 77 K using a Micromeritics ASAP 2020 analyzer, and the specific surface areas (SBET, m2 g−1) were calculated according to the Brunauer–Emmett–Teller (BET). Thermogravimetric analysis (TG) was conducted on a NETZSCH STA 449 C at 40–900 °C under air or argon atmosphere. A HORIBA LabRam HR800 Raman spectrometer, operated with a 532 nm laser, was employed to collect the Raman spectra. The chemical states of the samples were analyzed via X-ray photoelectron spectroscopy (XPS) on an ESCALAB 250 system under ultra-high vacuum (UHV), with a monochromatic aluminum K-alpha (Al Kα) X-ray source with a photon energy of 1486.6 electron volts, a power of 150 watts, and a pass energy of 50.0 electron volts to investigate the chemical states of the samples. Fourier transform infrared (FT-IR) spectra were recorded with a Bruker-V70 spectrometer. For temperature programmed surface reaction (TPSR) experiments, methanol and ethylbenzene (EB) were tested on a Pfeiffer Omnistar mass spectrometer, with a helium flow rate of 15 milliliters per minute, a heating rate of 5 degrees Celsius per minute, and 50 mg of catalyst; the evolutions of EB (mass-to-charge ratio 106), styrene (mass-to-charge ratio 104), and formaldehyde (mass-to-charge ratio 30) were monitored online.
2.3. Measurement of EB DDH activity
The catalytic performance for ethylbenzene direct dehydrogenation (EB DDH) was evaluated in a tubular fixed-bed reactor operating under standard atmospheric conditions, using helium as the carrier gas under plug-flow conditions. Ethylbenzene was metered into a preheated evaporator using a Cole Parmer (78-9100C) micro-syringe pump and carried by the helium stream into the reactor while maintaining the desired partial pressure. Effluent gases were monitored in real time with an Agilent 7890A gas chromatograph fitted with flame ionization (FID) and thermal conductivity (TCD) detectors. The FID equipped with a methyl silicone HP-5 capillary column was used to achieve separation and measurement of EB and ST, while He, CO, and CO2 were determined using two Porapak Q packed columns in combination with a molecular sieve 5A column interfaced with the TCD. The overall carbon balance was maintained within 100 ± 5%.
The EB conversion (XEB), the ST selectivity (SST), and carbon balance (Cbalance) were determined using:
3. Results and discussion
3.1. Chemical and structural features of C60–Ni/SiO2 catalysts
The C60–Ni/SiO2 hybrid catalysts were prepared through a simple sequential process involving drying and thermal reduction procedures. Briefly, the SiO2 support was sequentially impregnated with an aqueous nickel nitrate solution and a C60–toluene dispersed solution, followed by washing with ethanol to remove the residual toluene. The resulting mixture was then dried and thermally reduced at 600 °C in a mixed H2/Ar gas flow for 2 h to obtain the C60–Ni/SiO2 catalyst.
Fig. 1a displays the spherical morphology of the SiO2 support, with particle sizes ranging from 20 to 30 nm. A similar spherical shape morphology was retained in the C60–Ni/SiO2 catalyst, consistent with that of the Ni/SiO2 catalyst (Fig. S1). The well-dispersed Ni nanoparticles (NPs) exhibit sizes ranging from 5 to 10 nm, and C60 could be observed on the surface of Ni NPs. The selected lattice spacing of 0.204 nm (Fig. 1b) corresponds to the (111) plane of Ni, and an enlarged image (Fig. 1c) shows the diameter of 0.70 nm features corresponding to C60 molecules (theoretical diameter: 0.71 nm),37 confirming that C60 is successfully anchored on the surface of the Ni NPs. Furthermore, HAADF-STEM and EDX elemental mapping (Fig. 1d–i) images show uniform distributions of C and Ni elements, further confirming the presence of well-dispersed C60 molecules and Ni species on the surface of C60–Ni/SiO2.
 |
| | Fig. 1 (a–c) TEM and high-magnification TEM images of C60–Ni/SiO2. (d–i) HAADF-STEM images and the corresponding EDX mappings of C60–Ni/SiO2. | |
The Ni/SiO2 and C60–Ni/SiO2 catalysts show similar diffraction peaks corresponding to Ni at approximately 2θ = 44.5°, 51.8° and 76.3°, and no additional specific diffraction peaks could be detected in their XRD patterns (Fig. 2a). The incorporation of C60 causes no shift in the diffraction peaks, indicating that the addition of C60 does not affect the crystalline structure of Ni.38 These XRD results are consistent with the observations from TEM images, which reveal that the Ni nanoparticles maintain a similar size distribution and dispersion on the surface of SiO2 support regardless of C60 modification, confirming that C60 incorporation does not induce significant morphological changes to the Ni phase. The Raman spectra of C60 show two intense signals at approximately 495 cm−1 and 1469 cm−1, corresponding to the Ag(1) mode and the Ag(2) mode, respectively, which are recognized as the characteristic vibration signal of C60 molecules (Fig. 2b).39–41 Similarly, the C60–Ni/SiO2 catalysts also showed typical Ag(1) and Ag(2) Raman bands, confirming the successful incorporation of C60 into the C60–Ni/SiO2 catalysts. The textural properties, including the specific surface area, were measured using N2 adsorption–desorption isotherms (Fig. 2c). All the samples exhibited adsorption–desorption behaviors similar to typical type IV isotherms according to the IUPAC classifications,42,43 which are characteristics of mesoporous materials. Capillary condensation occurred between the relative pressures of 0.8 and 1.0, and therefore exhibited the type H3 hysteresis loop. Pristine SiO2 exhibited the highest BET surface area (330.5 m2 g−1), indicative of well-developed mesoporosity. In contrast, the incorporation of C60 or Ni reduces the surface area primarily due to partial pore blockage. Interestingly, the C60–Ni/SiO2 composite maintained a relatively high surface area (121.9 m2 g−1), suggesting the high dispersion of Ni and C60 components. Furthermore, the C60 content in the C60–Ni/SiO2 catalyst was estimated to be approximately 1.5 wt% based on thermogravimetric analysis (TGA) as shown in Fig. 2d. The TGA curve of C60–Ni/SiO2 under an atmosphere of Ar (Fig. S2) shows that no significant weight loss could be observed below 900 °C, indicating the relatively high thermal stability of the C60–Ni/SiO2 catalyst under inert conditions.
 |
| | Fig. 2 (a) XRD diffraction patterns of C60–Ni/SiO2, Ni/SiO2, C60/SiO2, and SiO2. (b) Raman spectra of C60–Ni/SiO2, Ni/SiO2, C60/SiO2 and C60. (c) N2 adsorption–desorption isotherms of C60–Ni/SiO2, Ni/SiO2, C60/SiO2 and SiO2. (d) TG profiles under air atmosphere of C60–Ni/SiO2, Ni/SiO2, C60/SiO2 and C60 (with the inset showing the magnified view of the dashed-line region). | |
3.2. Catalytic activity of C60–Ni/SiO2 catalysts in the EB DDH reaction
After successful construction of the C60–Ni/SiO2 catalysts, their catalytic performance in the EB DDH reaction, an industrially relevant reaction for ST synthesis, was systematically evaluated.44,45 The reaction was performed at 550 °C using a feed gas composed of 1% EB balanced in He under ambient pressure. Fig. 3a shows the conversion of EB over time on stream for the C60–Ni/SiO2 catalyst with Ni/SiO2 and C60/SiO2 catalysts for comparison. The C60/SiO2 catalyst displayed minimal catalytic performance, with a steady state EB conversion of 9.5% (Fig. S3). In contrast, the Ni/SiO2 catalyst exhibited a carbon balance close to zero, owing to significant carbon accumulation on the catalyst. For comparison, the C60–Ni/SiO2 catalysts achieved a significantly higher steady EB conversion (30.8%) than both C60/SiO2 and Ni/SiO2, highlighting the promotional role of C60 and the synergistic effect between C60 and Ni in optimizing the catalytic performance of the catalyst in the EB DDH process. TGA and Raman spectra of the spent catalysts (Fig. S9 and S10) demonstrate that carbon deposition on C60–Ni/SiO2 is significantly lower than that on Ni/SiO2, confirming the coke-resistant effect of C60 modification. The EB conversion over C60–Ni/SiO2 remained at 25.7% after 18 hours on stream at 550 °C (Fig. 3b), and the selectivity to ST on C60–Ni/SiO2 kept over 95% even without steam protection, indicating the relatively high stability of the C60–Ni/SiO2 catalyst. This result indicates that the incorporation of C60 successfully inhibits the undesired side reactions. A 60 h long-term stability test was conducted under identical conditions (Fig. S11). The C60–Ni/SiO2 catalyst maintained an EB conversion of 20–25% and a consistent ST selectivity above 95%, demonstrating its excellent structural robustness and strong resistance to coke formation. Moreover, the catalytic performance of C60–Ni/SiO2 surpasses the combined activities of Ni/SiO2 and C60/SiO2, implying a synergistic interaction between the two components. Because the direct dehydrogenation (DDH) of ethylbenzene is endothermic, elevating the reaction temperature can promote the activation and breaking of the C–H bond. Consequently, EB conversion rises from 18.4% to 49.5% with the reaction temperature rising from 530 to 600 °C (Fig. 3c). However, it should be noted that elevated temperatures also promote the C–C bond cleavage as a secondary reaction, which causes a decrease in ST selectivity from 97.1% to 91.5%. The C60–Ni/SiO2 catalyst achieved an ST formation rate of 2.7 mmol g−1 h−1 at 550 °C (Fig. 3d and Table S1), exceeding that of commercial K–Fe (∼1.0 mmol g−1 h−1) and other reported typical catalysts under the same reaction conditions, demonstrating excellent potential for practical styrene production. For commercial EB dehydrogenation processes, conventional Fe–K catalysts are typically operated under high space velocities and elevated temperatures, delivering long lifetimes but with moderate selectivity.46 Pt–Sn catalysts exhibit superior selectivity and durability, but they rely on costly and scarce noble metals.47 In contrast, the C60–Ni/SiO2 catalyst, though evaluated at the laboratory scale, bridges the gap between non-noble and noble-metal catalysts, approaching the performance of Pt–Sn systems while maintaining cost efficiency. These results underscore the critical role of C60 as an electronic promoter and demonstrate the potential of this non-noble system for sustainable styrene production.
 |
| | Fig. 3 (a) Conversion of EB over time on C60–Ni/SiO2, C60/SiO2, and Ni/SiO2; (b) EB conversion and ST selectivity of C60–Ni/SiO2 versus reaction time; (c) EB conversion and corresponding ST selectivity over C60–Ni/SiO2 versus reaction temperature. Conditions: 100 mg catalysts, 550 °C, 1.0 kPa EB, total flow 6.12 mL min−1 with He as balance gas, carbon balance: 100 ± 5%. (d) Comparisons of the ST formation rates of C60–Ni/SiO2 with those of other typical reported catalysts at different temperatures. | |
Fig. 4a shows a positive correlation between the EB partial pressure and ST formation rate over the C60–Ni/SiO2 catalyst, suggesting that EB activation kinetically controls the reaction under these conditions. The reaction order for EB activation was 0.79 (Fig. 4b), implying that EB activation likely serves as the rate-determining step (RDS) over C60–Ni/SiO2, consistent with other metal catalysts.44,48–51 The kinetic isotope effect (KIE), quantifying the relative rates of isotopically labeled and unlabeled reactants, is commonly employed to elucidate the kinetic relevance of specific elementary steps. A KIE experiment was conducted on the C60–Ni/SiO2 catalyst using deuterated EB (D-EB, C8D10) to further validate the proposed mechanism. As shown in Fig. 4c, the EB conversion over the C60–Ni/SiO2 catalyst decreases sharply from 24.5% to 13.4%. A reaction rate ratio of 1.82 between EB and deuterated EB (REB/RD-EB) was observed under steady-state conditions when EB was substituted with D-EB. A similar KIE is also observed for the C60/SiO2 catalyst (Fig. S4). This result confirms that C–H bond activation and dissociation in EB serve as the RDS under the tested conditions. Furthermore, the KIE results indicate that the activation energy is closely correlated with the capability of the catalysts in activating the C–H bond.51,52 As shown in Fig. 4d, under conditions where C–H bond activation was the RDS, the apparent activation energies of EB DDH on C60–Ni/SiO2 and C60/SiO2 were determined. The activation energy for C60–Ni/SiO2 was measured at 80 kJ mol−1, significantly lower than that for C60/SiO2 at 152 kJ mol−1, suggesting that the incorporation of C60 substantially lowers the activation energy. This effect highlights the promotional role of C60 in C–H bond activation. C60–Ni/SiO2 demonstrates a lower activation energy than the 105.0 kJ mol−1 typically reported for EB DDH over commercial Fe–K catalysts.53 These results highlight the unique kinetic advantage of C60–Ni/SiO2 in facilitating EB dehydrogenation.
 |
| | Fig. 4 (a) ST formation rates versus partial pressure of EB and (b) corresponding reaction order for the C60–Ni/SiO2 catalyst. Conditions: 100 mg catalyst, 550 °C, EB partial pressure 0.5–7.0 kPa, total flow 6.12 mL min−1 with He as balance gas, carbon balance: 100 ± 5%. (c) Time-on-stream EB conversion over the C60–Ni/SiO2 catalyst in EB (C8H10)/He and D-EB (C8D10)/He. Conditions: 100 mg catalyst, 550 °C, EB or D-EB partial pressure 1.0 kPa, total flow 6.12 mL min−1 with He as balance gas, carbon balance: 100 ± 5%. (d) Arrhenius plots for EB DDH over the C60–Ni/SiO2 and C60/SiO2 catalysts. Conditions: 100 mg catalyst, 530–570 °C, EB partial pressure 1.0 kPa, total flow 6.12 mL min−1 with He as balance gas, carbon balance: 100 ± 5%. | |
3.3. Promotional effect of C60 on Ni species in C60–Ni/SiO2 catalysts
In the Raman spectra of C60, an intense signal is observed at 1469 cm−1, which is the characteristic vibration signal of the C60 molecule due to the pentagonal pinch mode Ag(2).39,40 The Raman spectra of C60/SiO2 exhibit a superposition of signals from C60 moieties, with no discernible shift relative to that of pristine C60, indicating no strong intermolecular interactions between C60 and SiO2. In contrast, a noticeable downshift of the Ag(2) vibration mode is observed in the Raman spectra of C60–Ni/SiO2, with an approximately 5 cm−1 peak shifting from 1469 to 1464 cm−1. As reported, the position of the Ag(2) mode Raman signal reflects changes in the charge state of C60;40,54 therefore, the observed downshift is indicative that there is electron migration from Ni to C60, attributable to the excellent electron-acceptor capability of C60.54,55 Elemental analysis and surface chemical state identification were carried out via X-ray photoelectron spectroscopy. As displayed in Fig. S5, four main peaks corresponding to C 1s, O 1s, Si 1s, and Si 2p XPS signals were identified, consistent with TEM and elemental mapping images. XPS characterization was used to reveal the promotional role of C60 by comparing the C60–Ni/SiO2 and unmodified Ni/SiO2 catalysts. The binding energy of Ni0 2p3/2 in the XPS spectra of C60–Ni/SiO2 appears at 856.9 eV, which is +1.0 eV higher than that of Ni/SiO2 (855.9 eV). A similar positive shift is observed for the Ni0 2p1/2 signal (873.93 → 874.68 eV). These increases in binding energy indicate a decrease of electron density at Ni sites upon C60 modification, consistent with the result that partial charge transfers from Ni to the electron-accepting C60. XPS atomic composition analysis reveals nearly identical surface Ni contents (0.49 at%) for C60–Ni/SiO2 and Ni/SiO2, confirming that the observed electronic changes arise from interfacial charge redistribution rather than variations of the loading content of Ni. The CO-FTIR results further corroborate this interpretation (Fig. S7). The IR band corresponding to linearly adsorbed CO on Ni/SiO2 appears at 2026 cm−1, whereas it is centered at 2054 cm−1 for C60–Ni/SiO2, representing a 28 cm−1 blueshift. Such a blueshift of the C–O stretching vibration is typically attributed to the weakened π back-donation from d orbitals of the metal to the 2π* antibonding orbital of CO, directly reflecting a reduction in the d-electron density at metal sites.56–58 The concurrent observation of a 1.0 eV XPS shift and a 28 cm−1 CO frequency blueshift is consistent with Raman analysis, thus providing mutually consistent and semi-quantitative evidence that C60 modification results in a decreased electron density of Ni species.
Temperature programmed surface reaction (TPSR) experiments with methanol and EB were performed over the C60–Ni/SiO2 catalyst to elucidate the promotional effect of C60 in the EB DDH process.59 In these TPSR experiments, the selected reactant was first adsorbed onto the catalyst surface, and then purged with helium overnight ensured the removal of physisorbed species. The catalyst was subsequently heated to 500 °C at a rate of 5 °C min−1, while product evolution was continuously monitored by in situ mass spectrometry. The analysis focused on formaldehyde (FA) formation from methanol dehydrogenation and styrene (ST) production from EB dehydrogenation as a function of reaction temperature. Methanol was employed as a probe molecule to assess the redox properties of the catalytically active sites on the catalysts.60 The temperature at which FA desorbed reflected redox ability, with lower desorption temperatures corresponding to higher redox capability.60,61 The redox capacity is crucial for C–H bond activation during EB DDH. As shown in Fig. 5c, the desorption temperature of formaldehyde on both Ni/SiO2 and C60–Ni/SiO2 catalysts was observed at 160 °C. Notably, the C60–Ni/SiO2 catalyst exhibited a larger peak area than Ni/SiO2, suggesting that it possesses a higher density of accessible active sites and is less susceptible to carbon deposition. As shown in Fig. 5d and S8, no detectable signals corresponding to EB and ST were observed for Ni/SiO2 catalysts, possibly due to the strong adsorption of EB on Ni sites. This strong interaction may facilitate deep dehydrogenation and C–C bond dissociation, leading to substantial carbon deposition and a resulting decrease in catalytic performance. In contrast, the signal intensity of EB and ST on C60–Ni/SiO2 and C60/SiO2 catalysts is much higher than that on Ni/SiO2 at temperatures over 200 °C, suggesting more efficient desorption of chemisorbed EB and enhanced ST formation on these C60-based catalysts.
 |
| | Fig. 5 (a) Raman spectra of C60–Ni/SiO2, Ni/SiO2, C60/SiO2, and C60. (b) Deconvolution of the Ni 2p XPS spectra of C60–Ni/SiO2 and Ni/SiO2. (c) MS signals under temperature-programmed reaction conditions of methanol on C60–Ni/SiO2, Ni/SiO2, and C60/SiO2. A heating rate of 5 °C per minute. (d) MS signals of ST under temperature-programmed reaction conditions of EB on the C60–Ni/SiO2, Ni/SiO2, and C60/SiO2 surfaces. A heating rate of 5 °C per minute. | |
4. Conclusions
In summary, the C60–Ni/SiO2 catalyst was successfully synthesized through a simple impregnation method, yielding a SiO2 microsphere structure with Ni and C60 uniformly distributed on its surface. Unlike conventional nickel-based catalysts, which normally suffer from carbon deposition, the proposed C60–Ni/SiO2 catalyst not only suppresses coking, but also exhibits enhanced catalytic activity for EB DDH reactions. The C60–Ni/SiO2 catalyst exhibited superior catalytic performance in the EB DDH reactions, achieving a remarkable ST selectivity of 99.2% and an ST formation rate of 2.7 mmol g−1 h−1. This performance surpassed that of most reported Ni-based catalysts and commercial catalysts under similar reaction conditions. KIE results confirmed that the rate-determining step in EB DDH reactions was the activation of the C–H bond under the catalysis of C60–Ni/SiO2. The electron-withdrawing properties of C60 facilitate the electron transfer from Ni NPs to C60 molecular promoters and this transfer is driven by the higher Fermi level of Ni relative to the LUMO of C60, resulting in favorable adsorption and desorption behavior of different intermediates on C60–Ni/SiO2 compared to traditional Ni-based catalysts. The high catalytic performance of C60–Ni/SiO2 composites offers a highly effective and promising strategy in the informed design of efficient catalysts in alkane dehydrogenation, offering valuable insights for future informed design of Ni catalysts targeting alkane dehydrogenation.
Conflicts of interest
The authors declare no competing interests.
Data availability
Data are available from the corresponding author upon reasonable request. Supplementary materials include extended characterization of the SiO2 support and SiO2-based composites, providing additional structural, textural, and thermal analysis relevant to catalyst performance.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5cy01075f.
Acknowledgements
The authors acknowledge the financial support from the National Natural Science Foundation of China (U23A20545 and 22072163), the Natural Science Foundation of Liaoning Province of China (2024JH3/50100016), the China Baowu Low Carbon Metallurgy Innovation Foundation (BWLCF202113), and the Shccig-Qinling Program.
References
- Z. Qin, J. Liu, A. Sun and J. Wang, Ind. Eng. Chem. Res., 2003, 42, 1329–1333 CrossRef CAS.
- K. Zha, T. Zeng, M. Zhu, C. Wei, L. Song and C. Miao, Appl. Catal., A, 2023, 666, 119372 CrossRef CAS.
- Z. Zhang, T. Zeng, C. Wei, L. Song and C. Miao, Mol. Catal., 2023, 540, 113058 CAS.
- Y. Sekine, R. Watanabe, M. Matsukata and E. Kikuchi, Catal. Lett., 2008, 125, 215–219 CrossRef CAS.
- N. Mimura, I. Takahara, M. Saito, T. Hattori, K. Ohkuma and M. Ando, Catal. Today, 1998, 45, 61–64 CrossRef CAS.
- R.-J. Zhang, G.-F. Xia, M.-F. Li, W. Yu, N. Hong and D.-D. Li, J. Fuel Chem. Technol., 2015, 43, 1359–1365 CrossRef CAS.
- K. Hou and R. Hughes, Chem. Eng. J., 2001, 82, 311–328 CrossRef CAS.
- A. Mohsenzadeh, A. Borjesson, J.-H. Wang, T. Richards and K. Bolton, Int. J. Mol. Sci., 2013, 14, 23301–23314 CrossRef PubMed.
- J. Carrasco, L. Barrio, P. Liu, J. A. Rodriguez and M. V. Ganduglia-Pirovano, J. Phys. Chem. C, 2013, 117, 8241–8250 CrossRef CAS.
- F. L. Chan and A. Tanksale, Renewable Sustainable Energy Rev., 2014, 38, 428–438 CrossRef CAS.
- B. Abdullah, N. A. Abd Ghani and D.-V. N. Vo, J. Cleaner Prod., 2017, 162, 170–185 CrossRef CAS.
- W. Shen, Y. Wang, X. Shi, N. Shah, F. Huggins, S. Bollineni, M. Seehra and G. Huffman, Energy Fuels, 2007, 21, 3520–3529 CrossRef CAS.
- Z. J. Wang, H. Song, H. Liu and J. Ye, Angew. Chem., Int. Ed., 2020, 59, 8016–8035 CrossRef CAS PubMed.
- Y. Guo, Y. Li, Y. Ning, Q. Liu, L. Tian, R. Zhang, Q. Fu and Z.-J. Wang, Ind. Eng. Chem. Res., 2020, 59, 15506–15514 CrossRef CAS.
- J. J. Sattler, J. Ruiz-Martinez, E. Santillan-Jimenez and B. M. Weckhuysen, Chem. Rev., 2014, 114, 10613–10653 CrossRef CAS.
- E. Gianotti, M. Taillades-Jacquin, J. Rozière and D. J. Jones, ACS Catal., 2018, 8, 4660–4680 CrossRef CAS.
- Z. Yan and D. W. Goodman, Catal. Lett., 2012, 142, 517–520 CrossRef CAS.
- S. Strömberg, M. Svensson and K. Zetterberg, Organometallics, 1997, 16, 3165–3168 CrossRef.
- C. Massera and G. Frenking, Organometallics, 2003, 22, 2758–2765 CrossRef CAS.
- G. Zhang, C. Yang and J. Miller, ChemCatChem, 2018, 10, 961–964 CrossRef CAS.
- H. Wang, H. Wang, X. Li and C. Li, Appl. Surf. Sci., 2017, 407, 456–462 CrossRef CAS.
- H. W. Kroto, J. R. Heath, S. C. O'Brien, R. F. Curl and R. E. Smalley, Nature, 1985, 318, 162–163 CrossRef CAS.
- M. A. Lebedeva, T. W. Chamberlain and A. N. Khlobystov, Chem. Rev., 2015, 115, 11301–11351 CrossRef CAS PubMed.
- R. Wakimoto, T. Kitamura, F. Ito, H. Usami and H. Moriwaki, Appl. Catal., B, 2015, 166, 544–550 CrossRef.
- L. Jia, M. Chen and S. Yang, Mater. Chem. Front., 2020, 4, 2256–2282 RSC.
- J. Chen, J. Luo, E. Hou, P. Song, Y. Li, C. Sun, W. Feng, S. Cheng, H. Zhang and L. Xie, Nat. Photonics, 2024, 18, 464–470 CrossRef CAS.
- Q. Xie, E. Perez-Cordero and L. Echegoyen, J. Am. Chem. Soc., 1992, 114, 3978–3980 CrossRef CAS.
- L. Echegoyen and L. E. Echegoyen, Acc. Chem. Res., 1998, 31, 593–601 CrossRef CAS.
- C. A. Reed and R. D. Bolskar, Chem. Rev., 2000, 100, 1075–1120 CrossRef CAS PubMed.
- L. M. Andersson, Org. Electron., 2011, 12, 300–305 CrossRef CAS.
- J. Zheng, L. Huang, C.-H. Cui, Z.-C. Chen, X.-F. Liu, X. Duan, X.-Y. Cao, T.-Z. Yang, H. Zhu and K. Shi, Science, 2022, 376, 288–292 CrossRef CAS PubMed.
- Y. Zhang, X. Peng, H.-R. Tian, B. Yang, Z.-C. Chen, J. Li, T. Zhang, M. Zhang, X. Liang and Z. Yu, Nat. Chem., 2024, 16, 1781–1787 CrossRef CAS PubMed.
- R. Zhang, Y. Li, X. Zhou, A. Yu, Q. Huang, T. Xu, L. Zhu, P. Peng, S. Song and L. Echegoyen, Nat. Commun., 2023, 14, 2460 CrossRef CAS.
- A. Dahal and M. Batzill, Nanoscale, 2014, 6, 2548–2562 RSC.
- J. Li, E. Croiset and L. Ricardez-Sandoval, Appl. Surf. Sci., 2014, 317, 923–928 CrossRef CAS.
- K. Miyaura, Y. Miyata, B. Thendie, K. Yanagi, R. Kitaura, Y. Yamamoto, S. Arai, H. Kataura and H. Shinohara, Sci. Rep., 2018, 8, 8098 CrossRef.
- J. Guan, X. Zhong, X. Chen, X. Zhu, P. Li, J. Wu, Y. Lu, Y. Yu and S. Yang, Nanoscale, 2018, 10, 2473–2480 RSC.
- J. Wu, S. Wang, Z. Lei, R. Guan, M. Chen, P. Du, Y. Lu, R. Cao and S. Yang, Nano Res., 2021, 1–10 Search PubMed.
- D. Saha and S. Deng, Langmuir, 2011, 27, 6780–6786 CrossRef CAS PubMed.
- F. Leng, I. C. Gerber, P. Lecante, S. Moldovan, M. Girleanu, M. R. Axet and P. Serp, ACS Catal., 2016, 6, 6018–6024 CrossRef CAS.
- M. Matus, H. Kuzmany and W. Krätschmer, Solid State Commun., 1991, 80, 839–842 CrossRef CAS.
- F. Jing, B. Katryniok, S. Paul, L. Fang, A. Liebens, M. Shen, B. Hu, F. Dumeignil and M. Pera-Titus, ChemCatChem, 2017, 9, 258–262 CrossRef CAS.
- S. Brunauer, L. S. Deming, W. E. Deming and E. Teller, J. Am. Chem. Soc., 1940, 62, 1723–1732 CrossRef CAS.
- W. Liu, B. Chen, X. Duan, K.-H. Wu, W. Qi, X. Guo, B. Zhang and D. Su, ACS Catal., 2017, 7, 5820–5827 CrossRef CAS.
- J. Zhang, X. Wang, Q. Su, L. Zhi, A. Thomas, X. Feng, D. S. Su, R. Schlögl and K. Müllen, J. Am. Chem. Soc., 2009, 131, 11296–11297 CrossRef CAS PubMed.
- E. H. Lee, Catal. Rev.: Sci. Eng., 1974, 8, 285–305 CrossRef.
- L. Deng, T. Arakawa, T. Ohkubo, H. Miura, T. Shishido, S. Hosokawa, K. Teramura and T. Tanaka, Ind. Eng. Chem. Res., 2017, 56, 7160–7172 CrossRef CAS.
- J. Zhang, D. S. Su, A. Zhang, D. Wang, R. Schlögl and C. Hébert, Angew. Chem., Int. Ed., 2007, 46, 7319–7323 CrossRef CAS PubMed.
- S. Mao, B. Li and D. Su, J. Mater. Chem. A, 2014, 2, 5287–5294 RSC.
- X. Zhang, X. Dai, Z. Xie and W. Qi, Small, 2024, 20, 2401532 CrossRef CAS.
- X. Dai, T. Cao, X. Lu, Y. Bai and W. Qi, Appl. Catal., B, 2023, 324, 122205 CrossRef CAS.
- W. Qi, P. Yan and D. S. Su, Acc. Chem. Res., 2018, 51, 640–648 CrossRef CAS PubMed.
- J. Sheng, B. Yan, W.-D. Lu, B. Qiu, X.-Q. Gao, D. Wang and A.-H. Lu, Chem. Soc. Rev., 2021, 50, 1438–1468 RSC.
- A. Talyzin and U. Jansson, Thin Solid Films, 2003, 429, 96–101 CrossRef CAS.
- X. Zhu, T. Zhang, D. Jiang, H. Duan, Z. Sun, M. Zhang, H. Jin, R. Guan, Y. Liu and M. Chen, Nat. Commun., 2018, 9, 4177 CrossRef PubMed.
- J. I. Cohen and R. Tobin, J. Chem. Phys., 2018, 148, 22 CrossRef PubMed.
- J. Lauterbach, M. Wittmann and J. Küppers, Surf. Sci., 1992, 279, 287–296 CrossRef CAS.
- Y. Peng, A. Melillo, R. Shi, A. Forneli, A. Franconetti, J. Albero and H. García, Appl. Catal., B, 2023, 339, 123143 CrossRef CAS.
- H. Huang, F. Tang, J. Sheng, Z. Liu, J. Fan, R.-P. Zhang, D. Wang and A.-H. Lu, J. Catal., 2024, 436, 115626 CrossRef CAS.
- M. Badlani and I. E. Wachs, Catal. Lett., 2001, 75, 137–149 CrossRef CAS.
- J. Sheng, W.-C. Li, W.-D. Lu, B. Yan, B. Qiu, X.-Q. Gao, R.-P. Zhang, S.-Z. Zhou and A.-H. Lu, Appl. Catal., B, 2022, 305, 121070 CrossRef CAS.
Footnote |
| † These authors contributed equally to this work. |
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.