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

Regenerable oxygen-deficient Ni/γ-Al2O3 catalyst for efficient glycerol aqueous phase reforming

Jung Hyun Park a, Hong Lu a, Brajendra K. Sharma b, David Johnston b, Nandakishore Rajagopalan a and Jaemin Kim *a
aIllinois Sustainable Technology Center, University of Illinois at Urbana-Champaign, 1Hazelwood Drive, Champaign, IL 61820, USA. E-mail: jaemin@illinois.edu
bU.S. Department of Agriculture, Agricultural Research Service, Eastern Regional Research Center, Sustainable Biofuels and Co-Products Research Unit, 600 E. Mermaid Lane, Wyndmoor, PA 19038, USA

Received 13th October 2025 , Accepted 2nd December 2025

First published on 5th December 2025


Abstract

Aqueous phase reforming (APR) of glycerol represents a promising pathway for sustainable fuel gas generation. Nickel-immobilized gamma alumina (Ni/γ-Al2O3) has been recognized as an effective alternative to noble metal catalysts, but the phase transformation from γ-Al2O3 to AlOOH under hydrothermal conditions negatively affects its long-term catalytic performance. To address this challenge, we synthesized a Ni/γ-Al2O3 catalyst via a Ni-exsolution technique from NiAl2O4 spinel oxide. The catalyst achieved a gasification yield of 49.2% with a fuel gas energy of 9.2 MJ kg−1 of glycerol in 45 min at 250 °C, producing hydrogen, carbon monoxide, and methane, which is comparable to that of the Ru-catalyst. The spent catalyst was regenerated, resulting in an increased gasification yield of 52.6% and fuel gas energy of 10.3 MJ kg−1 of glycerol, with enhanced H2 (106.7%) and CH4 (123.0%) production compared to the fresh catalyst. This remarkable performance is primarily attributed to improved crystallinity of γ-Al2O3 and strengthened Ni and γ-Al2O3 interactions induced by increased oxygen vacancies and electron density. This study highlights the significance of the metal exsolution approach in catalyst preparation, demonstrating that chemical structure modulation through regeneration is crucial for enhancing both the catalytic activity and durability of γ-Al2O3 supported catalysts in glycerol APR.


Introduction

Producing fuel gases such as hydrogen (H2) and methane (CH4) from sustainable feedstock is key to developing cleaner energy routes. Glycerol, a renewable byproduct from biodiesel manufacturing1–3 and bioethanol production,4–7 is an attractive feedstock for the aqueous phase reforming (APR) process due to its oversupply relative to demand. The APR of glycerol produces carbon monoxide (CO) and H2 (eqn (1)), while the produced CO is most likely converted to carbon dioxide (CO2) and H2via the water–gas shift (WGS) reaction (eqn (2)). Overall, APR of glycerol is described by using eqn (3):8
 
C3H8O3 (l) → 3CO + 4H2 ΔH° = 245 kJ mol−1(1)
 
CO + H2O ↔ CO2 + H2 ΔH° = −41 kJ mol−1(2)
 
C3H8O3 (l) + 3H2O ↔ 7H2 + 3CO2 ΔH° = 123 kJ mol−1(3)

The produced CO, CO2, and H2 can be further converted to methane through reactions in eqn (4) and (5), respectively:

 
CO + 3H2 ↔ CH4 + H2O ΔH° = −206 kJ mol−1(4)
 
CO2 + 4H2 ↔ CH4 + 2H2O ΔH° = −165 kJ mol−1(5)

Fuel gas yields are strongly influenced by the catalyst used under typical APR reaction conditions (∼250 °C; ∼580 psig).9–12 For practical applications, catalysts must exhibit excellent hydrothermal stability to sustain fuel gas production over time. The desired catalyst possesses uniform dispersion of active metals, small particle size, strong metal–support interactions, and highly exposed triple-phase boundaries among the catalyst, support, and reactants.13 In this context, careful selection of catalyst materials with tailored synthesis methods is crucial for developing durable, practical catalysts for sustainable fuel gas production.

Nickel-based catalysts are attractive alternatives to noble metal-based catalysts such as Ru, Pd, and Pt in biomass treatment due to their earth abundance and excellent catalytic performance. Nickel as a catalytic center is active for C–C bond cleavage, water–gas shift, and methanation reactions.14 In addition to Ni, metal oxide supports such as Al2O3, ZrO2, and CeO2 can enhance the catalytic performance of Ni by modulating the electronic structure and dispersing Ni nanoparticles to optimize active site exposure.15,16 γ-Al2O3 features a porous structure with abundant acidic and basic sites that facilitate catalytic reactions through strong interaction with reactants.17,18 Therefore, Ni supported on γ-Al2O3 (Ni/γ-Al2O3) has been extensively studied for hydrothermal biomass treatment16,19–21 and methanation processes.22–25 γ-Al2O3, however, suffers from low hydrothermal stability, which undergoes phase transformation to AlOOH, leading to structural collapse and activity loss.26–28 Nonetheless, effective strategies to improve the hydrothermal stability of γ-Al2O3 without compromising catalytic performance remain scarce, making this an ongoing research challenge.29

Metal exsolution techniques, unlike conventional catalyst reduction, enable the formation of homogeneously dispersed, finely anchored metal nanoparticles on an oxide support,30–33 leading to exceptional catalytic activity.34–36 Additionally, oxygen vacancies within the oxide lattice serve as crucial active sites that promote water activation and enhance CO2 chemisorption.37–39 Here, we present a Ni/γ-Al2O3 catalyst, prepared by exsolving Ni from NiAl2O4 spinel oxide, for producing fuel gases from glycerol APR (Scheme 1). The Ni/γ-Al2O3 catalyst, characterized by uniformly distributed Ni nanoparticles on a γ-Al2O3 support, demonstrated high yields of H2, CO, and CH4 in glycerol APR. We carefully investigated the catalytic pathways and active sites using X-ray photoelectron spectroscopy (XPS) and proton nuclear magnetic resonance (1H-NMR), and examined the phase segregation process during APR. The spent, phase-segregated catalysts were regenerated through consecutive calcination and exsolution processes, which presented enhanced catalytic activity with improved hydrothermal stability. Comprehensive XPS characterization confirmed that increased oxygen vacancies and enhanced γ-Al2O3 crystallinity significantly contributed to the improved catalytic activity and stability in glycerol APR.


image file: d5ta08347h-s1.tif
Scheme 1 Illustration depicting the exsolution of Ni nanoparticles from NiAl2O4 spinel oxide to generate the Ni/γ-Al2O3 catalyst for enhanced aqueous phase reforming (APR) of glycerol.

Results and discussion

Synthesis of Ni/γ-Al2O3 through Ni-exsolution

Ni/γ-Al2O3 was prepared via Ni-exsolution from NiAl2O4 spinel oxide. The initial Ni–Al complex was prepared by a sol–gel method using Ni(NO3)2·6H2O and Al(NO3)3·9H2O as metal precursors and citric acid as a chelating agent. NiAl2O4 spinel was obtained through calcination of the Ni–Al complex at 800 °C for 6 h in an air environment. The scanning electron microscope (SEM) image in Fig. S1a supports that as-prepared NiAl2O4 has irregular morphology with relatively large, undefined particles. A higher-magnification SEM image in Fig. 1a further confirmed the smooth surface feature of NiAl2O4. After the exsolution process, Ni/γ-Al2O3 exhibited well-dispersed Ni nanoparticles (Ni NPs, 20–30 nm) anchored on the support surface, while its overall morphology remained largely unchanged (Fig. 1b and S1b).
image file: d5ta08347h-f1.tif
Fig. 1 Characterization of the prepared compounds. SEM images of (a) NiAl2O4 spinel oxide and (b) the Ni/γ-Al2O3 catalyst. (c) X-ray diffraction patterns before (NiAl2O4) and after reduction (Ni/γ-Al2O3).

The structure of NiAl2O4 and Ni/γ-Al2O3 was further investigated using X-ray diffraction (XRD) spectroscopy, as shown in Fig. 1c. The diffraction patterns of the prepared NiAl2O4 matched well with the reference peaks of NiAl2O4 (PDF# 01-078-6956), with the Fd-3ms space group and cubic crystal structure. Exsolved Ni NPs from the parent NiAl2O4 spinel were clearly observed from the XRD patterns after the reduction process, while characteristic peaks for NiAl2O4 were not detected. Well-developed diffraction peaks at 2θ degrees of 44.4, 51.7, and 76.3 correspond to the (111), (200), and (220) facets, respectively, of cubic structured Fm-3m Ni (PDF# 01-076-4179). The Ni NP size was calculated to be 25.7 nm using the Scherrer equation, which is in good agreement with the particle size observed in the SEM images (Fig. 1b). The rest of the diffraction peaks are matched with γ-phase Al2O3 (PDF# 01-076-4179), whose space group is Fd-3mZ with a cubic crystal structure. These suggest the successful exsolution of Ni NPs from NiAl2O4 with complete phase transformation. The structure of the Ni-exsolved catalyst was examined at various temperatures, indicating that 800 °C is the minimum temperature required to fully convert NiAl2O4 into Ni/γ-Al2O3 (Fig. S2). This finding is consistent with the hydrogen temperature-programmed reduction (H2-TPR) results reported in previous literature,40 which indicated complete reduction of Ni species near this temperature.

For controls, Ni NPs and Ni/γ-Al2O3 were additionally prepared by the hydrothermal process and conventional wet impregnation method, respectively. Ni nanoparticles prepared via the hydrothermal process had a spherical shape with wide particle size distribution in the 100–300 nm range (Fig. S3 and S4). Ni/γ-Al2O3, prepared by the conventional wet impregnation (Ni/γ-Al2O3_w) method, presented Ni nanoparticle sizes from 40 to 110 nm on γ-Al2O3, resulting from the uncontrollable Ni agglomeration at the reduction step (Fig. S5 and S6). It is noteworthy that, unlike Ni/γ-Al2O3_w, Ni NPs in Ni/γ-Al2O3 are strongly anchored on the surface of γ-Al2O3,41 resulting in a homogeneous particle size distribution (20–30 nm). The uniform dispersion of Ni NPs provides a high density of accessible Ni active sites and triple-phase boundaries, which are essential for efficient glycerol APR and methanation reactions.

APR of glycerol

The glycerol APR was conducted in a Parr high-pressure reactor using 1.0 g of the Ni/γ-Al2O3 catalyst and 30 mL of a 0.2 M glycerol aqueous solution at 250 °C. The evolution of gaseous products (H2, CO, CH4, and CO2) was monitored every 15 min for 45 min (Fig. 2). In Ni/γ-Al2O3-catalyzed APR, 12 mmol H2 and 1.9 mmol CO were produced within the first 15 min (Fig. 2a and b). Then H2 production yield remained largely unchanged for 45 min whereas CO decreased to <0.5 mmol. In the meantime, CH4 production gradually increased and reached a maximum yield of 1.3 mmol at 45 min (Fig. 2c). The observed CH4 formation is likely attributed to the methanation reactions of CO and CO2 with H2, where the γ-Al2O3 support facilitates the chemisorption of CO and CO2 (Fig. 2d).42–45
image file: d5ta08347h-f2.tif
Fig. 2 The catalytic performance of the Ni/γ-Al2O3 catalyst for APR of glycerol at 250 °C. The production of (a) hydrogen, (b) carbon monoxide, (c) methane, and (d) carbon dioxide. (e) Gasification yield of glycerol and (f) energy of the gas products as a function of time.

The gasification yield of glycerol is presented in Fig. 2e. The yield was determined by a carbon balance calculation, comparing the total carbon in the glycerol feed with the carbon detected in the gaseous products, CO, CH4, and CO2. The gasification yield reached 49.2% at 45 min. The observed yield is comparable to that of Ru catalysts,46,47 which is attributed to the increased triple phase boundaries surrounding Ni NPs achieved through the metal exsolution approach. To further evaluate the fuel quality, the energy of the product gas was determined (Fig. 2f). The gas product energy was stabilized at 9.2 MJ kg−1 of glycerol after 30 min and remained unchanged for an additional 15 min. The higher heating value (HHV) of the gas product obtained under the Ni/γ-Al2O3 system was calculated to be 13.7 MJ kg−1 (Table S1).

The catalytic performance of Ni/γ-Al2O3 was evaluated with control catalysts, Ni NPs, γ-Al2O3, and Ni/γ-Al2O3_w, in at least three independent experiments, as shown in Fig. 3. In the tests, Ni NPs produced on average 8.7 mmol H2, 0.5 mmol CO, 0.8 mmol CH4, and 6.9 mmol CO2. The gasification yield of Ni NPs was 45.8%, which was similar to that of Ni/γ-Al2O3 (49.2%). However, Ni/γ-Al2O3 produced 52.1% more H2 and 62.2% more CH4, despite containing only 35.5% Ni content by chemical composition. These results indicate that the enhanced H2 and CH4 production of Ni/γ-Al2O3 is mainly attributed to the increased chemisorption of CO and CO2 on the γ-Al2O3 support.42,45 γ-Al2O3 alone, in contrast, showed negligible catalytic activity for glycerol APR, producing minimal amounts of H2 (0.08 mmol), CO (0.02 mmol), CH4 (0.04 mmol), and CO2 (1.2 mmol), comparable to the non-catalyzed system (H2: 0 mmol, CO: 0 mmol, CH4: 0 mmol, and CO2: 0.8 mmol). The gasification yield of γ-Al2O3 alone was only 6.8%, confirming that active metallic Ni NPs are the primary catalytic sites and γ-Al2O3 mainly serves for chemisorption of CO and CO2 rather than direct fuel gas production. For the Ni/γ-Al2O3_w control catalyst, the average gasification yield was 31.1%, most likely due to reduced active sites and triple phase boundaries resulting from the larger Ni particle size (40–110 nm) compared to that of Ni/γ-Al2O3 (20–30 nm). The HHV for the gas product, which was influenced by the relative proportion of the combustible gases, was evaluated to be 9.7, 1.2 and 13 MJ kg−1 in the presence of Ni NPs, γ-Al2O3, and Ni/γ-Al2O3_w, respectively (Table S1). Overall, the performed control experiments highlight the significance of the Ni exsolution strategy for APR catalysts, which generates smaller, uniformly dispersed Ni nanoparticles on the support.


image file: d5ta08347h-f3.tif
Fig. 3 Control experiments for APR of glycerol at 250 °C under an N2 atmosphere. Comparison of gas productions: (a) hydrogen, (b) carbon monoxide, (c) methane, and (d) carbon dioxide. (e) Gasification yields on a carbon basis and (f) energy of the gas products. All experiments were conducted at least three times to ensure reproducibility.

Ni/γ-Al2O3 was further tested in a pure water system, in the absence of glycerol, to elucidate the origin of hydrogen and carbon in fuel gas production. The measured gas products were 1.5 mmol H2, 1.1 mmol CO2, and a negligible amount of CH4 (0.03 mmol). This result indicates that Ni/γ-Al2O3 can activate water to produce H2. Taken together with the negligible H2 production observed from the γ-Al2O3 system, it is evident that Ni NPs play a crucial role in H2 evolution from water. The low CH4 yield is likely attributed to dissolved CO2 in water, as reflected in the CO2 production from the γ-Al2O3 system. NiAl2O4 was also tested for the glycerol APR, where 10.4 mmol of H2 and 2.2 mmol of CO2 were produced with negligible CO and CH4 generation and a low gasification yield of 13.7% (Fig. S7).

Glycerol APR reaction pathways

To investigate the glycerol APR reaction pathway in the Ni/γ-Al2O3 system, individual reaction tests were conducted using γ-Al2O3, Ni NPs, and Ni/γ-Al2O3 catalysts, respectively. The intermediate species generated from each reaction were characterized using nuclear magnetic resonance (NMR) spectroscopy (Fig. S8–S11 and 4). The APR of glycerol has been understood to proceed primarily through two pathways:48 (1) a dehydration pathway yielding hydroxyacetone as an intermediate and (2) a dehydrogenation pathway yielding glyceraldehyde as an intermediate.
image file: d5ta08347h-f4.tif
Fig. 4 Suggested reaction mechanisms for the APR of glycerol over the Ni/γ-Al2O3 catalyst at 250 °C under a N2 atmosphere.

In the glycerol APR test using γ-Al2O3 alone, the 1H-NMR spectrum exhibited a multiplet at 3.78 ppm along with two doublets of doublet at 3.65 ppm and 3.55 ppm, corresponding to the characteristic proton peaks for glycerol (Fig. S8). This observation aligns with the negligible gasification yield shown in Fig. 3e. Hydroxyacetone (singlets at 4.37 ppm and 2.14 ppm) and propyleneglycol (doublet of doublet at 3.43 ppm and doublet at 1.13 ppm) were detected, indicating that the dehydration pathway is dominant in the presence of γ-Al2O3.36,48 Ethanol was identified by a triplet at 1.18 ppm, whereas the expected quartet at 3.65 ppm was not resolved due to overlap with the glycerol peak. The trace ethanol likely results in minor CO production (Fig. 3b), which is produced alongside hydrogen.40,48 This hydrogen is possibly utilized for the hydrogenation of hydroxyacetone to propyleneglycol, with the remaining H2 detected in Fig. 3a.

In the Ni NP-catalyzed glycerol APR system, glycerol was rarely detected, while weak peaks attributed to acetaldehyde appeared at 9.67 ppm (quartet) and 2.24 ppm (doublet) (Fig. S9). Relatively intense peaks corresponding to ethanol (triplet at 1.18 ppm; quartet at 3.65 ppm) and acetone (singlet at 2.23 ppm) further indicate that both dehydration and dehydrogenation pathways are active. However, considering the high H2 and CO production yields in Fig. 3a and b, it can be inferred that the dehydrogenation pathway is predominant, as glyceraldehyde formation produces more H2 and CO than hydroxyacetone.48,49

In the presence of Ni/γ-Al2O3, glycerol APR reaction solutions were analyzed at 30 min and 45 min. The 1H-NMR spectra of the 30 min sample (Fig. S10) presented a distinct quartet peak at 9.67 ppm along with an intense doublet at 2.24 ppm for acetaldehyde. In addition, a quartet at 5.25 ppm and a doublet at 1.32 ppm correspond to ethane-1,1′-diol, likely formed via hydration of acetaldehyde.50,51 The simultaneous presence of Ni NPs and γ-Al2O3 facilitates glycerol conversion to both hydroxyacetone and glyceraldehyde.40 Since Ni/γ-Al2O3 actively produces H2 from glycerol and water (Fig. 3a), sufficient hydrogen is available to hydrogenate hydroxyacetone to propyleneglycol. This intermediate subsequently dehydrates to acetone (2.23 ppm, singlet), which can further generate CO and CH4. In 45 min (Fig. S11), most of the hydroxyacetone, propyleneglycol, and acetone were consumed. The observed decrease in ethanol and methanol (singlet at 3.36 ppm) further supports the predominance of the dehydrogenation pathway, consistent with increased CO, H2, and CO2 production.

Overall, Ni/γ-Al2O3 converts glycerol into both hydroxyacetone and glyceraldehyde via dehydration and dehydrogenation pathways, respectively, leading to the formation of H2, CO, CH4, and CO2. These gaseous products undergo water–gas shift and methanation reactions on the catalyst surface, resulting in fuel gas production. The strong chemisorption of CO2 and CO on γ-Al2O3, combined with the increased triple phase boundaries in Ni/γ-Al2O3, underpins its excellent catalytic activity for fuel gas generation. Although acetic acid and 2-propanol are potential intermediates in glycerol APR, they were not detected in the reaction mixture of this study.

Catalyst stability and regeneration

To evaluate the durability of the Ni/γ-Al2O3 catalyst, the spent catalyst was recovered after glycerol APR and subsequently characterized. XRD analysis in Fig. 5a showed that the majority of γ-Al2O3 was converted into AlOOH (PDF# 01-073-9093) with an orthorhombic crystal structure (Cmcm space group) during the reaction. Notably, the characteristic reflections of γ-Al2O3 were completely absent after the reaction, indicating a thorough phase conversion. Only a weak diffraction peak for γ-Al2O3 was observed at a 2θ degree of 45.8, in line with previous reports that γ-Al2O3 hydrates to form AlOOH under aqueous phase conditions.26–28 XRD peaks for metallic Ni became more prominent after glycerol APR without further oxidation into NiOx. The SEM image of the used catalyst showed a reduction in particle size as a result of the phase transition from γ-Al2O3 to AlOOH (Fig. S12). While agglomeration of Ni NPs was observed due to the instability of the support, some Ni NPs remained anchored, retaining their original particle size, as shown in Fig. 5b.
image file: d5ta08347h-f5.tif
Fig. 5 Catalytic performance of the used Ni/γ-Al2O3 catalyst. (a) XRD patterns and (b) SEM image of the used Ni/γ-Al2O3 catalyst. Comparison of fresh Ni/γ-Al2O3 and used Ni/γ-Al2O3 catalysts: (c) hydrogen, (d) carbon monoxide, (e) methane, and (f) carbon dioxide production. (g) Gasification yields of glycerol on a carbon basis and (h) HHV of the gas products. At least three independent experiments were conducted to ensure reliability.

The spent Ni/γ-Al2O3 catalyst was reused for glycerol APR to clarify the catalytically active species (Fig. 5c–h). Upon reuse, the catalyst produced 8.2 mmol H2, 0.52 mmol CO, 0.79 mmol CH4, and 5.7 mmol CO2, resulting in a decreased gasification yield of 39.3% and product gas energy of 5.8 MJ kg−1 of glycerol. This result closely resembled that of Ni NPs or Ni/γ-Al2O3_w (Fig. 3). Given the smaller Ni particle size observed on the used catalyst compared to the fresh Ni NPs and Ni/γ-Al2O3_w, it is likely that the catalytic activity originated primarily from the Ni NPs present in the used catalyst. These results indicate that the catalyst exhibits excellent activity when Ni is strongly bound to γ-Al2O3 rather than AlOOH, thereby emphasizing the importance of restoring the Ni/γ-Al2O3 structure to maintain optimal catalytic performance.

To extend the lifetime of the spent catalyst by regeneration, the used Ni/γ-Al2O3 was collected by centrifugation and subjected to sequential treatments as shown in Fig. 6a. The used catalyst was first calcined at 800 °C for 6 h in air to form NiAl2O4 spinel oxide (Regen. NiAl2O4). It is worth noting that Ni/γ-Al2O3 prepared by the conventional wet impregnation method can partially be converted to NiAl2O4 under similar conditions, but the entire phase transformation has not been reported,26,52,53 and the conversion of Ni/AlOOH to NiAl2O4 remains unexplored. We have previously demonstrated a similar regeneration strategy for catalysts prepared via the metal exsolution technique.34 The converted phase after calcination was characterized by XRD (Fig. 6b). Notably, the peak at a 2θ degree of 37.3 became more intense, which can be attributed to the overlap between the NiAl2O4 spinel phase and the (111) facet of newly formed NiO. Additionally, XRD peaks at 2θ degrees of 43.3, 62.9, 75.44, and 79.4 correspond to the (111), (200), (220), (311), and (222) facets of NiO, respectively, confirming the presence of residual NiO in the regenerated NiAl2O4. The presence of NiO peaks suggests incomplete incorporation of Ni into the NiAl2O4 lattice, which may result from agglomeration of Ni NPs during the conversion of Al2O3 to AlOOH in the APR process.


image file: d5ta08347h-f6.tif
Fig. 6 (a) Sequential regeneration route of the used Ni/γ-Al2O3 catalyst and XRD spectral comparison of (b) fresh and regenerated NiAl2O4 spinel oxides and (c) fresh and regenerated Ni/γ-Al2O3 catalysts. SEM images of (d) and (e) regenerated NiAl2O4 spinel oxide and (f) and (g) regenerated Ni/γ-Al2O3.

The Regen. NiAl2O4 was further reduced at 800 °C for 6 h under a 10% H2 atmosphere to re-form Ni/γ-Al2O3, and its structure was characterized in Fig. 6c. The resulting diffraction patterns closely matched those of the fresh Ni/γ-Al2O3, while sharper Ni peaks were observed after the regeneration process. Notably, the peaks corresponding to γ-Al2O3 at 2θ degrees of 37.3, 39.3 and 45.6 became more pronounced, indicating increased crystallinity of the γ-Al2O3 support after regeneration. This improved crystallinity is likely to contribute to enhanced hydrothermal stability of γ-Al2O3 under APR conditions, which will be discussed further in a later section.

The SEM image in Fig. 6d shows a well-defined structure of the Regen. NiAl2O4, with particle sizes comparable to those of the used catalyst, ranging from hundreds of nanometers to micrometers (Fig. S12). Smaller particles observed in Regen. NiAl2O4 in Fig. 6e are NiO particles, as confirmed by XRD data. The overall morphology of the Regen. NiAl2O4 was retained during the reduction process to produce Regen. Ni/γ-Al2O3, resulting in numerous exsolved Ni NPs on the γ-Al2O3 surface (Fig. 6f and g). Energy dispersive X-ray spectroscopy (EDS) mapping presented that the larger, well-defined particles predominantly consist of Al and O, while the smaller particles are Ni NPs (Fig. S13), indicating agglomeration and a less uniform Ni distribution compared to the fresh Ni/γ-Al2O3 catalyst. This agglomeration likely arises from the presence of large NiO particles formed on the spinel oxide during regeneration.

To evaluate the influence of AlOOH and NiO on particle size and morphology during regeneration, fresh Ni/γ-Al2O3 was further regenerated without conducting the APR process (Regen. NiAl2O4 w/o APR) for comparison. The XRD patterns of Regen. NiAl2O4 w/o APR displayed less pronounced NiO peaks (2θ degrees of 43.3, 62.9, 75.44, and 79.4) compared to Regen. NiAl2O4 that was regenerated after the glycerol APR process (Fig. S14). This difference implies that the absence of phase transformation from γ-Al2O3 to AlOOH provides better anchoring, and less agglomeration and formation of NiO. Consequently, the regenerated Ni/γ-Al2O3 without the glycerol APR process (Regen. Ni/γ-Al2O3 w/o APR) exhibited a more intense Ni diffraction peak (2θ degrees of 44.4, 51.7, and 76.3), which correlates with the increased Ni NP size observed in SEM images (Fig. S15). Detailed glycerol APR performance using the Regen. Ni/γ-Al2O3 w/o APR catalyst will be shown and discussed in a later section.

The catalytic performance of Regen. Ni/γ-Al2O3 is presented in Fig. 7. Regen. Ni/γ-Al2O3 produced 14.2 mmol H2, 0.8 mmol CO, 1.6 mmol CH4, and 7.0 mmol CO2, which are comparable to or higher than those of the fresh Ni/γ-Al2O3 catalyst. It showed a high gasification yield of 52.6% with a product gas energy of 10.3 MJ kg−1 of glycerol and HHV of 14.8 MJ kg−1. This higher fuel gas energy, compared to the fresh catalyst (9.2 MJ kg−1 of glycerol), is mainly attributed to the increased fuel gas production of H2 (106.7%) and CH4 (123%), despite the Regen. Ni/γ-Al2O3 catalyst containing larger Ni NPs with fewer exposed triple-phase boundaries.


image file: d5ta08347h-f7.tif
Fig. 7 Glycerol APR performance of the regenerated Ni/γ-Al2O3 catalyst. The amounts of gas products: (a) hydrogen, (b) carbon monoxide, (c) methane, and (d) carbon dioxide. (e) Gasification yield of glycerol on a carbon basis, and (f) HHV of the gas products. An arrow (red) indicates the regeneration process.

The improved hydrothermal stability of the Regen. Ni/γ-Al2O3 catalyst was observed through consecutive cyclic tests (cycles 3–5 in Fig. 7) conducted without further regeneration: H2 production gradually decreased to 9.2 mmol by the 5th cycle, whereas CO and CH4 yields remained stable at 0.8 mmol and 1.1 mmol, respectively. XRD analysis for the cyclic tests (Fig. S16) further confirmed the enhanced stability of the Regen. Ni/γ-Al2O3 catalyst. The catalyst exhibited only minor structural changes over the cycles, which is in great contrast to the fresh Ni/γ-Al2O3 that underwent complete transformation into Ni/AlOOH (Fig. 5a). The diffraction peaks of Regen. γ-Al2O3 remained largely unchanged with increased γ-Al2O3 crystallinity, although peaks corresponding to AlOOH gradually appeared during repeated glycerol APR cycles. Overall, the regenerated catalyst, Regen. Ni/γ-Al2O3, demonstrated improved hydrothermal stability and sustained catalytic performance compared to the fresh catalyst.

The catalytic activity of Regen. Ni/γ-Al2O3 w/o APR was further evaluated as shown in Fig. S17. Despite possessing smaller Ni NPs and a greater exposure of triple phase boundaries than Regen. Ni/γ-Al2O3, the catalyst exhibited lower activity in glycerol APR by producing 9.6 mmol H2, 0.5 mmol CO, 1.0 mmol CH4, and 5.1 mmol CO2 with a HHV of 6.8 MJ kg−1 of glycerol. These results importantly indicate that the phase transformation from γ-Al2O3 to AlOOH not only promotes Ni NP agglomeration but also alters the catalyst's chemical structure in ways that critically impact its catalytic activity for glycerol APR.

To examine the correlation between glycerol APR performance (activity and durability) and the chemical structure of the catalysts, X-ray photoelectron spectroscopy (XPS) was performed on fresh Ni/γ-Al2O3, Regen. Ni/γ-Al2O3, and Regen. Ni/γ-Al2O3 w/o APR (Fig. 8 and S19). All spectra were deconvoluted using the C–C bond from the C 1s peak at 284.8 eV as a reference (Fig. S18) as the initial step for analysis. In the Ni 2p XPS spectrum of the fresh Ni/γ-Al2O3 catalyst (Fig. 8a), metallic Ni (Ni0) was identified at 852.4 eV, while peaks for Ni2+ and Ni3+ appeared at 853.8 and 855.5 eV, respectively. The Al 2p XPS spectrum (Fig. 8b) shows a peak at 74.5 eV corresponding to Al3+, and a characteristic peak of γ-Al2O3, with an additional peak at 70–65 eV attributed to the Ni 3p orbital. The O 1s XPS spectrum in Fig. 8c shows lattice oxygen (Al–O) at 530.6 eV, an oxygen vacancy at 531.3 eV, and adsorbed oxygen at 532.0 eV.54–57


image file: d5ta08347h-f8.tif
Fig. 8 XPS spectra of fresh and regenerated Ni/γ-Al2O3. (a) Ni 2p, (b) Al 2p, and (c) O 1s spectra of fresh Ni/γ-Al2O3, and (d) Ni 2p, (e) Al 2p, and (f) O 1s spectra of regenerated Ni/γ-Al2O3.

For Regen. Ni/γ-Al2O3, notable shifts and changes were observed in all three spectra. The Ni0 peak in the Ni 2p spectrum shifted by 0.1 eV to lower binding energy (852.3 eV), indicating enhanced electron density, and the proportion of Ni0 increased from 28.9 to 31.6% (Fig. 8d). As metallic Ni serves as the active site, its increased content strongly correlates with improved catalytic performance.58–60 In contrast, Ni2+ and Ni3+ peaks shifted by 0.1 eV to higher binding energy, which is attributed to the strengthened interactions between the support and exsolved Ni NPs. Similarly, the Al 2p peak in the regenerated catalyst shifted by 0.1 eV to higher binding energy (Fig. 8e). Considering the increased crystallinity observed in XRD data (Fig. 6c), this shift suggests stronger interactions between Ni and γ-Al2O3, while improved support crystallinity contributes to enhanced hydrothermal stability.

Oxygen vacancies, recognized as active sites for water activation and CO/CO2 chemisorption,61 showed a significant increase for Regen. Ni/γ-Al2O3. In the O 1s XPS spectrum (Fig. 8f), the lattice oxygen peak remained unchanged, while the oxygen vacancy peak shifted from 531.3 to 531.2 eV, and the adsorbed oxygen peak shifted from 532.0 to 532.1 eV, indicating the increased relative concentration of oxygen vacancies. The calculated oxygen vacancy content increased from 42.5% to 51.9%. XPS data suggest that the increased amounts of metallic Ni and oxygen vacancies in Regen. Ni/γ-Al2O3 likely contributed to its sustained and robust catalytic performance, despite the presence of larger Ni NPs and fewer exposed triple phase boundaries.35,62

The origin of the oxygen vacancy was examined by comparing XPS data of Regen. NiAl2O4 and Regen. Ni/γ-Al2O3 w/o APR (Fig. S19–S21 and Table S2). The regenerated NiAl2O4 spinel oxide exhibited increased oxygen vacancies, estimated at 51.6% based on the relative area of the vacancy-associated peak in the O 1s spectrum, compared to 39.0% in the fresh NiAl2O4. Consequently, Regen. NiAl2O4 showed a lower Ni2+ binding energy of 853.3 eV, compared to 853.9 eV in fresh NiAl2O4, consistent with the lower binding energy of metals in oxygen-deficient metal oxides.38,62–64 This confirms the presence of increased oxygen vacant sites in Regen. Ni/γ-Al2O3. In contrast, Regen. Ni/γ-Al2O3 w/o APR exhibited a similar level of oxygen vacancy content (42.4%) to that of fresh Ni/γ-Al2O3. In summary, the phase transformation (γ-Al2O3 →AlOOH → γ-Al2O3) during regeneration is crucial to generate abundant oxygen vacancies in the γ-Al2O3 support while enhancing its crystallinity as evidenced by a 0.3 eV shift in the Al3+ peak (from 74.2 eV to 74.5 eV). Together, these changes contributed to the improved catalytic activity and durability of the Regen Ni/γ-Al2O3 catalyst. The detailed XPS peak deconvolution and quantitative results are summarized in Table 1, providing a comparative overview of the electronic states and elemental compositions for fresh and regenerated catalysts.

Table 1 XPS analysis results of fresh and regenerated Ni/γ-Al2O3
  Nickel Aluminum Oxygen
Ni0 Ni2+ Ni3+ Al–O Ov Oad
Fresh Ni/γ-Al2O3 Binding energy [eV] 852.4 853.8 855.5 74.5 530.6 531.3 532.0
Amount [%] 28.9 8.3 62.8   21.9 42.5 35.7
Regen. Ni/γ-Al2O3 Binding energy [eV] 852.3 853.9 855.6 74.6 530.6 531.2 532.1
Amount [%] 31.6 3.8 64.6   18.4 51.9 29.7


Conclusions

In this study, we demonstrated the enhanced catalytic performance of the Ni/γ-Al2O3 catalyst for fuel gas production via glycerol APR. The Ni/γ-Al2O3 catalyst was prepared using a Ni-exsolution technique starting from NiAl2O4 spinel oxide as the parent material. With homogeneously distributed Ni nanoparticles, the catalyst exhibited excellent activity towards H2, CO, and CH4 production under mild reaction conditions. Mechanistic analysis revealed that the Ni/γ-Al2O3 catalyst promotes both dehydration and dehydrogenation pathways during glycerol conversion to gaseous products. To assess durability, the spent catalyst was regenerated through sequential heat treatments. The regenerated catalyst (Regen. Ni/γ-Al2O3) showed improved catalytic activity and durability in consecutive cyclic tests. These enhancements are mainly attributed to the increased crystallinity of the oxide support and strengthened interactions between Ni and γ-Al2O3, facilitated by increased oxygen vacancies and electron density, as confirmed by XPS analysis. Overall, these findings not only advance the understanding of glycerol APR, but also create new opportunities to enhance the catalytic activity and stability of γ-Al2O3-supported catalysts for diverse aqueous phase reactions and sustainable fuel gas production from various biomass-derived waste streams, thereby promoting integrated waste-to-fuel strategies within circular bioeconomy frameworks.

Experimental

Materials

All chemicals were used without further purification. Nickel nitrate hexahydrate (98%), aluminum nitrate nonahydrate (98%), hydrazine hydrate solution (>98.0%), γ-Al2O3 (99.95%), and sodium hydroxide (98%) were purchased from Sigma-Aldrich (USA). Citric acid (99%), nickel chloride hexahydrate (99.3%), and glycerol (99%) were purchased from Fisher Scientific (USA).

Synthesis of the Ni/γ-Al2O3 catalyst

NiAl2O4 spinel oxide was prepared via a sol–gel process using citric acid as the chelating agent. In a typical preparation, 2.9 g (10 mmol) of Ni(NO3)2·6H2O, 7.5 g (20 mmol) of Al(NO3)3·9H2O, and 12.6 g (60 mmol) of citric acid were dissolved in 50 mL of deionized water in a 200 mL beaker under continuous stirring. The solution was heated at 150 °C overnight, yielding a dried solid complex. The resulting product was ground into a fine powder using a mortar and pestle and subsequently calcined in air at 800 °C for 6 h with a ramping rate of 5 °C min−1 to produce NiAl2O4. Finally, the NiAl2O4 powder was reduced under 10% H2 (90% N2) at 800 °C for 6 h leading to exsolution of Ni nanoparticles on γ-Al2O3 (denoted as Ni/γ-Al2O3).

As a control, Ni/γ-Al2O3 was also prepared via a conventional wet impregnation method. 8.43 g (20 mmol) of Ni(NO3)2·6H2O was dissolved in 100 mL of methanol and 2 g (20 mmol) of γ-Al2O3 powder was added. The suspension was stirred at 80 °C overnight to ensure uniform adsorption of the nickel precursor onto the support. The resulting dry powder was collected and reduced under 10% H2 (90% N2) at 800 °C for 6 h to yield Ni nanoparticle immobilized γ-Al2O3 (denoted as Ni/γ-Al2O3_w).

Synthesis of Ni nanoparticles

For Ni nanoparticle preparation, 4 g of NaOH and 2 g of NiCl2·6H2O were dissolved in 32 mL and 20 mL of an ethanol–water mixture (2[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v), respectively. These solutions were then mixed, and 10 mL of hydrazine monohydrate was added dropwise. The mixture was transferred to a Teflon-lined autoclave (200 mL) and heated at 115 °C for 2 h. The product was washed three times using water and ethanol, respectively, followed by centrifugation at 8000 rpm.

APR of glycerol

The as-prepared Ni/γ-Al2O3 catalyst (1 g) and 30 mL of 0.2 M glycerol solution were added to a 250 mL Parr reactor (Series 4570 HP/HT reactor, Parr Instrument Company, USA) equipped with a reactor controller (Model: 4848, Parr Instrument Company, USA). The reactor was sealed, evaluated for leakage using 400 psig of nitrogen gas, and purged with 400 psig of nitrogen six times. The reactor was heated to 250 °C for about 50 min, and the glycerol hydrothermal treatment was conducted. The measured pressure at 250 °C was 600 psig. Gas samples from the headspace of the reactor were collected by water substitution and analyzed by using a thermal conductivity detector (TCD)-equipped gas chromatograph (GC, 5890 Series II, Hewlett Packard, USA) using helium as a carrier gas. Calibration of the instrument was achieved using standard reference gases of ultra-high purity grade, and the volume change during the hydrothermal gasification was compensated by measuring the N2 concentration in the headspace of the reactor. The gasification yield was calculated using eqn (6):
 
image file: d5ta08347h-t1.tif(6)
where G.Y. is the gasification yield of glycerol in aqueous phase reforming under subcritical water conditions, MCO, MCH4, and MCO2 are molar amounts of CO, CH4, and CO2 obtained after the reaction, and MC,glycerol is the molar amount of carbon in the glycerol solution. The quantification of H2 includes generation from both glycerol and water.

The higher heating value (HHV) of the fuel gases produced from glycerol was calculated using the following equation:

 
image file: d5ta08347h-t2.tif(7)
where HHV is the higher heating value of the fuel gas produced from glycerol (MJ kg−1), i is a compound in the product fuel gas, m is the mass of the compound in the fuel gas (kg kg−1), and ΔcH° is the standard heat of combustion (MJ kg−1).

Regeneration of the spent catalyst

The spent catalyst was collected by using a centrifuge at 8000 rpm for 10 min and was rinsed using water and ethanol three times. Then the used catalyst was heated at 800 °C for 6 h with a ramping rate of 5 °C min−1 in an air atmosphere. The resulting NiAl2O4 spinel (denoted as Regen. NiAl2O4) was further reduced by a thermal treatment at 800 °C for 6 h with a ramping rate of 5 °C min−1 under 10% H2 (90% N2) to obtain regenerated Ni/γ-Al2O3 (denoted as Regen. Ni/γ-Al2O3).

Author contributions

Jung Hyun Park: conceptualization, methodology, data curation, writing – original draft, visualization, investigation. Hong Lu: methodology, writing – review & editing. Brajendra Kumar Sharma: writing – review & editing. David Johnston: writing – review & editing. Nandakishore Rajagopalan: writing – review & editing, supervision. Jaemin Kim: funding acquisition, conceptualization, writing – review & editing, supervision.

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/d5ta08347h.

Acknowledgements

This work was supported by the USDA Agriculture and Food Research Initiative (AFRI), project award no. 2024-67021-42492, from the U.S. Department of Agriculture's National Institute of Food and Agriculture. Powder XRD was performed in George L. Clark X-Ray Facility and 3M Materials Laboratory in School of Chemical Sciences (SCS) at the University of Illinois, Urbana-Champaign (UIUC). The NMR study was carried out in the NMR laboratory in SCS at UIUC. SEM and XPS were carried out in Materials Research Laboratory (MRL) Central Facilities at UIUC.

References

  1. C. Venu, D. Palanisamy, S. Jaganathan and S. Rajendran, Renew. Energy, 2023, 219, 119594 CrossRef CAS.
  2. F. Yang, M. A. Hanna and R. Sun, Biotechnol. Biofuels, 2012, 5, 13 CrossRef CAS PubMed.
  3. S. Haosagul, N. Vikromvarasiri, V. Sawasdee and N. Pisutpaisal, Int. J. Hydrogen Energy, 2019, 44, 29568–29574 CrossRef CAS.
  4. Y. Kim, N. S. Mosier, R. Hendrickson, T. Ezeji, H. Blaschek, B. Dien, M. Cotta, B. Dale and M. R. Ladisch, Bioresour. Technol., 2008, 99, 5165–5176 CrossRef CAS PubMed.
  5. K. Liu, J. Agr. Food Chem., 2011, 59, 1508–1526 CrossRef CAS PubMed.
  6. F. Adom, J. Fan, J. Davis, P. Dunn and D. Shonnard, ACS Sustain. Chem. Eng., 2014, 2, 1139–1146 CrossRef CAS.
  7. K. Ratanapariyanuch, Y. Y. Shim, S. Emami and M. J. T. Reaney, J. Agr. Food Chem., 2016, 64, 9488–9496 CrossRef CAS PubMed.
  8. N. A. Roslan, S. Z. Abidin, A. Ideris and D.-V. N. Vo, Int. J. Hydrogen Energy, 2020, 45, 18466–18489 CrossRef CAS.
  9. S. Leng, S. Barghi and C. Xu, npj Mater. Sustain., 2024, 2, 19 CrossRef.
  10. L. Santiago-Martínez, M. Li, P. Munoz-Briones, J. Vergara-Zambrano, S. Avraamidou, J. A. Dumesic and G. W. Huber, Green Chem., 2024, 26, 7212–7230 RSC.
  11. F. Bastan, M. Kazemeini and A. S. Larimi, Renew. Energy, 2017, 108, 417–424 CrossRef CAS.
  12. A. Iriondo, V. L. Barrio, J. F. Cambra, P. L. Arias, M. B. Güemez, R. M. Navarro, M. C. Sánchez-Sánchez and J. L. G. Fierro, Top. Catal., 2008, 49, 46–58 CrossRef CAS.
  13. J. Oh, S. Joo, C. Lim, H. J. Kim, F. Ciucci, J.-Q. Wang, J. W. Han and G. Kim, Angew. Chem., Int. Ed., 2022, 61, e202204990 CrossRef CAS PubMed.
  14. R. R. Davda, J. W. Shabaker, G. W. Huber, R. D. Cortright and J. A. Dumesic, Appl. Catal., B, 2003, 43, 13–26 CrossRef CAS.
  15. J. Li, Y. Tu, K. He, C. Chen, L. Liang, C. Ruan and Q. Zhang, Molecules, 2025, 30, 1310 CrossRef CAS PubMed.
  16. P. Azadi, E. Afif, F. Azadi and R. Farnood, Green Chem., 2012, 14, 1766–1777 RSC.
  17. S. Carré, N. S. Gnep, R. Revel and P. Magnoux, in Studies in Surface Science and Catalysis, ed. A. Gédéon, P. Massiani and F. Babonneau, Elsevier, 2008, vol. 174, pp. 973–976 Search PubMed.
  18. H. Knozinger, in Studies in Surface Science and Catalysis, ed. B. Imelik, C. Naccache, G. Coudurier, Y. B. Taarit and J. C. Vedrine, Elsevier, 1985, vol. 20, pp. 111–125 Search PubMed.
  19. F. Bastan and M. Kazemeini, Biomass Convers. Bior., 2023, 13, 237–246 CrossRef.
  20. M. M. Rahman, T. L. Church, A. I. Minett and A. T. Harris, ChemSusChem, 2013, 6, 1006–1013 CrossRef CAS PubMed.
  21. M. Z. Hossain, M. R. Karim, S. Sutradhar, M. B. I. Chowdhury and P. A. Charpentier, Int. J. Hydrogen Energy, 2023, 48, 39791–39804 CrossRef CAS.
  22. S. Abelló, C. Berrueco and D. Montané, Fuel, 2013, 113, 598–609 CrossRef.
  23. J. Kumar Prabhakar, P. A. Apte and G. Deo, Chem.–Eng. J., 2023, 471, 144252 CrossRef CAS.
  24. G. Garbarino, C. Wang, T. Cavattoni, E. Finocchio, P. Riani, M. Flytzani-Stephanopoulos and G. Busca, Appl. Catal., B, 2019, 248, 286–297 CrossRef CAS.
  25. G. Garbarino, P. Kowalik, P. Riani, K. Antoniak-Jurak, P. Pieta, A. Lewalska-Graczyk, W. Lisowski, R. Nowakowski, G. Busca and I. S. Pieta, Ind. Eng. Chem. Res., 2021, 60, 6554–6564 CrossRef CAS.
  26. R. S. Zhou and R. L. Snyder, Acta Cryst., 1991, 47, 617–630 CrossRef.
  27. X. Liang, J. Tang, L. Li and Y. Wu, JOM, 2023, 75, 4689–4700 CrossRef CAS.
  28. A. P. Amrute, K. Jeske, Z. Łodziana, G. Prieto and F. Schüth, Chem. Mater., 2020, 32, 4369–4374 CrossRef CAS.
  29. Y. Zhang, B. Huang, M. K. Mardkhe and B. F. Woodfield, Microporous Mesoporous Mater., 2019, 284, 60–68 CrossRef CAS.
  30. J. Mei, T. Liao and Z. Sun, Mater. Today Energy, 2023, 31, 101216 CrossRef CAS.
  31. M. L. Weber, B. Šmíd, U. Breuer, M.-A. Rose, N. H. Menzler, R. Dittmann, R. Waser, O. Guillon, F. Gunkel and C. Lenser, Nat. Mater., 2024, 23, 406–413 CrossRef CAS PubMed.
  32. A. Bonkowski, M. J. Wolf, J. Wu, S. C. Parker, A. Klein and R. A. De Souza, J. Am. Chem. Soc., 2024, 146, 23012–23021 CrossRef CAS PubMed.
  33. D. Neagu, T.-S. Oh, D. N. Miller, H. Ménard, S. M. Bukhari, S. R. Gamble, R. J. Gorte, J. M. Vohs and J. T. S. Irvine, Nat. Commun., 2015, 6, 8120 CrossRef PubMed.
  34. J. H. Park, J. W. Scott, H. Lu, N. Rajagopalan, B. K. Sharma, S. D. Cosper, K. J. Hay and J. Kim, Chem.–Eng. J., 2024, 486, 150330 CrossRef CAS.
  35. A. Kumar, D. R. Kanchan, A. Banerjee, B. P. Singh and R. Srivastava, ACS Catal., 2025, 15, 8239–8258 CrossRef.
  36. M. Shokrollahi Yancheshmeh, O. Alizadeh Sahraei, M. Aissaoui and M. C. Iliuta, Appl. Catal., B, 2020, 265, 118535 CrossRef CAS.
  37. S. Zhao, Y. Wen, C. Du, T. Tang and D. Kang, Chem.–Eng. J., 2020, 402, 126180 CrossRef CAS.
  38. T. Zhang, W. Wang, F. Gu, W. Xu, J. Zhang, Z. Li, T. Zhu, G. Xu, Z. Zhong and F. Su, Appl. Catal., B, 2022, 312, 121385 CrossRef CAS.
  39. H. Fu and H. Lian, Chem.–Eng. J., 2024, 489, 151021 CrossRef CAS.
  40. A. Morales-Marín, J. L. Ayastuy, U. Iriarte-Velasco and M. A. Gutiérrez-Ortiz, Appl. Catal., B, 2019, 244, 931–945 CrossRef.
  41. O. Kwon, S. Joo, S. Choi, S. Sengodan and G. Kim, J. Phys. Energy, 2020, 2, 032001 CrossRef CAS.
  42. K. Föttinger, W. Emhofer, D. Lennon and G. Rupprechter, Top. Catal., 2017, 60, 1722–1734 CrossRef PubMed.
  43. S.-n. Yin, J. Zhao, S. Wu, X. Han and J. Ren, Chem.–Eng. J., 2025, 504, 158723 CrossRef CAS.
  44. Y. Li, C. Shen, C. Jiang, C. Liang, B. Chen, W. Ding and X. Guo, ACS Catal., 2025, 15, 9728–9737 CrossRef CAS.
  45. S. Liu, Z. Zhou, J. Chen, Y. Fu and C. Cai, Appl. Surf. Sci., 2023, 611, 155645 CrossRef CAS.
  46. S. Jeon, Y. M. Park, J. Park, K. Saravanan, H.-K. Jeong and J. W. Bae, Appl. Catal., A, 2018, 551, 49–62 CrossRef CAS.
  47. P. Gogoi, A. S. Nagpure, P. Kandasamy, C. V. V. Satyanarayana and T. Raja, Sustain. Energy Fuels, 2020, 4, 678–690 RSC.
  48. A. Fasolini, D. Cespi, T. Tabanelli, R. Cucciniello and F. Cavani, Catalysis, 2019, 9, 722 CAS.
  49. B. C. M. Morales and B. A. O. Quesada, Catal. Today, 2021, 372, 115–125 CrossRef CAS.
  50. G. Socrates, J. Org. Chem., 1969, 34, 2958–2961 CrossRef CAS.
  51. P. Diederich, T. Geisberger, Y. Yan, C. Seitz, A. Ruf, C. Huber, N. Hertkorn and P. Schmitt-Kopplin, Commun. Chem., 2023, 6, 38 CrossRef CAS PubMed.
  52. M. A. Goula, N. D. Charisiou, K. N. Papageridis, A. Delimitis, E. Pachatouridou and E. F. Iliopoulou, Int. J. Hydrogen Energy, 2015, 40, 9183–9200 CrossRef CAS.
  53. G. Garbarino, C. Wang, I. Valsamakis, S. Chitsazan, P. Riani, E. Finocchio, M. Flytzani-Stephanopoulos and G. Busca, Appl. Catal., B, 2015, 174–175, 21–34 CrossRef CAS.
  54. D. Li, Y. Li, X. Liu, Y. Guo, C.-W. Pao, J.-L. Chen, Y. Hu and Y. Wang, ACS Catal., 2019, 9, 9671–9682 CrossRef CAS.
  55. G. Caravaggio, L. Nossova and M. J. Turnbull, Chem.–Eng. J., 2021, 405, 126862 CrossRef CAS.
  56. M. Han, Z. Wang, Y. Xu, R. Wu, S. Jiao, Y. Chen and S. Feng, Mater. Chem. Phys., 2018, 215, 251–258 CrossRef CAS.
  57. J. Wang, D. N. Mueller and E. J. Crumlin, J. Eur. Ceram. Soc., 2024, 44, 116709 CrossRef CAS.
  58. Z. Refaat, M. E. Saied, A. O. A. E. Naga, S. A. Shaban, H. B. Hassan, M. R. Shehata and F. Y. E. Kady, Sci. Rep., 2023, 13, 4855 CrossRef CAS PubMed.
  59. Y. Kong, X. Jia, X. Chai, Z. Chen, C. Shang, X. Jiang, H. Cai, L. Jing, Q. Hu, H. Yang, X. Zhang and C. He, Natl. Sci. Rev., 2025, 12, nwaf173 CrossRef CAS PubMed.
  60. L. Shen, J. Xu, M. Zhu and Y.-F. Han, ACS Catal., 2020, 10, 14581–14591 CrossRef CAS.
  61. Z. Wei, M. Zhao, Z. Yang, X. Duan, G. Jiang, G. Li, F. Zhang and Z. Hao, Proc. Natl. Acad. Sci. U. S. A., 2023, 120, e2217148120 CrossRef CAS PubMed.
  62. G. S. More, B. P. Singh, R. Bal and R. Srivastava, Inorg. Chem., 2023, 62, 13069–13080 CrossRef CAS.
  63. B. Bharti, S. Kumar, H.-N. Lee and R. Kumar, Sci. Rep., 2016, 6, 32355 CrossRef CAS PubMed.
  64. T. F. Qahtan, T. O. Owolabi and T. A. Saleh, J. Mol. Liq., 2024, 393, 123556 CrossRef CAS.

Footnote

Mention of trade names or commercial products in this article is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the U.S. Department of Agriculture. USDA is an equal opportunity provider and employer.

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.