Next Article in Journal
Simultaneous Adsorption and Purification of Low-Concentration SO2 and H2S
Previous Article in Journal
Flavonoids Identified in Terminalia spp. Inhibit Gastrointestinal Pathogens and Potentiate Conventional Antibiotics via Efflux Pump Inhibition
Previous Article in Special Issue
The Application of Multifunctional Metal–Organic Frameworks for the Detection, Adsorption, and Degradation of Contaminants in an Aquatic Environment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Hydrogen Production from Ammonia Using Ru Nanoparticles on Ce-Based Metal–Organic Framework (MOF)-Derived CeO2 with Oxygen Vacancies

Key Laboratory of Cluster Science Ministry of Education, Beijing Key Laboratory of Photoelectronic/Electrophotonic Conversion Materials, Advanced Research Institute of Multidisciplinary Science, School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing 100081, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(11), 2301; https://doi.org/10.3390/molecules30112301
Submission received: 20 March 2025 / Revised: 20 May 2025 / Accepted: 21 May 2025 / Published: 23 May 2025

Abstract

:
Ammonia is a promising hydrogen storage material because it is easy to store and decompose into COX-free hydrogen. A Ru-based catalyst exhibits good catalytic performance in ammonia decomposition, and enhancing the interaction between the Ru atoms and the support is an important way to further improve its catalytic activity. In this study, CeO2 was prepared by calcination using a cerium-based metal–organic framework (MOF) as the precursor, and the number of oxygen vacancies on the surface of CeO2 was regulated by hydrogen reduction. The XPS and Raman results showed that abundant oxygen vacancies were formed on the surface of these CeO2, and their number increased with an increase in the reduction time. The Ru/CeO2-4 h catalyst, using CeO2 reduced for 4 h as the support, exhibited good catalytic activity in ammonia decomposition, reaching 98.9% ammonia conversion and 39.74 mmol gcat−1 min−1 hydrogen yield under the condition of GHSV = 36,000 mL gcat−1 h−1 at 500 °C. The XAFS results demonstrated that Ru was stably anchored with oxygen vacancies on the surface of CeO2 via Ru-O-Ce bonds. Density functional theory calculations further showed that these bondings lower the reaction energy barrier for N-H bond cleavage, thereby significantly enhancing the catalytic activity.

1. Introduction

Hydrogen energy storage and transportation is an important link restricting the development of hydrogen energy. The flammable and explosive characteristics of hydrogen make its transportation a difficult problem [1,2]. Ammonia is considered to be a promising hydrogen storage medium, which has the advantages of easy liquefaction (25 °C, 8.6 bar), high-volume hydrogen density (120 kg cm−3), low manufacturing cost, and itself being a carbon free fuel [1,3]. The properties of ammonia are conducive to the overall improvement of production, storage, and transportation infrastructure [4,5,6].
The decomposition reaction of ammonia is an endothermic process, manifested as follows:
2NH3 (g) → N2 (g) + 3H2 (g)  ΔH = 91.2 kJ mol−1
To achieve the complete conversion of ammonia, the decomposition temperature usually needs to be maintained over 800 °C, which would consume a large amount of energy. Thus, developing novelly efficient catalysts to reduce the energy consumption is very important for the industrial application of ammonia decomposition to hydrogen [7]. In recent years, the development of ammonia decomposition catalysts has made great progress [8,9,10,11]. Fe, Co, Ni, Ru, and other metal catalysts have been widely researched, providing a variety of options for the development of ammonia decomposition technology. Among these metals, Ru is the most effective active center due to its excellent catalytic activity at low and medium temperatures [12,13,14,15]. Meanwhile, the selection of the carrier is also very important. As reported in previous work, the strong metal–support interaction (SMSI) between Ru and the carrier could further enhance the catalytic activity of the Ru catalyst for NH3 decomposition [16,17,18,19,20,21,22].
Metal oxides, such as CeO2 [23,24,25], MgO [26,27], La2O3 [28], Al2O3 [29,30,31], ZrO2 [32,33], are widely used as carriers for Ru catalysts in the ammonia decomposition process due to their surface acidity/basicity, surface oxygen vacancy, reoxidation properties, and the interaction between metal and support. Among these carriers, CeO2 showed higher NH3 decomposition catalytic activity than Ru/Al2O3 due to the strong SMSI and electronic modification of Ru active sites by CeO2 [34]. Hu et al. [35] loaded Ru single atoms onto cerium oxide nanospheres (CeO2-Nss) prepared by an improved colloidal deposition method and cerium oxide nanorods (CeO2-NRs) prepared by the hydrothermal method. N2 and H2 on CeO2-Ns and CeO2-NR catalysts are more easily desorbed than MgO-supported catalysts prepared by other methods. According to the research, Ru/CeO2 has great potential as the ammonia decomposition catalyst.
For the last few years, the application of the metal–organic framework (MOF) in catalyst preparation has been widely studied [36]. The MOF structure combines inorganic metals and organic linkers with a high surface area, diversity of assemblies, and uniform porosity. These advantages give the MOF the potential to be the catalyst, catalyst support, or precursor in many chemical reactions [37]. There have been numerous reports on the preparation of CeO2 using MOF structures. Sivan et al. [38] reported a Ce-BTC-derived Ru/CeO2 catalyst with highly dispersed Ru, and it shows excellent and more stable performance in NH3 synthesis. He et al. [39] reported that the high-porosity Ru/CeO2 catalyst prepared with Ce-UiO-66 improved the catalytic performance of CO2 methanation compared with the CeO2-supported Ru catalyst prepared by the traditional method. Chen et al. [40] used CeO2 obtained by pyrolysis of Ce-MOF for toluene combustion; when compared with CeO2 prepared by the precipitation method, MOF-CeO2 exhibits a better catalytic activity due to its structure and abundant oxygen vacancy. According to the research, CeO2 derived from Ce-MOF could exhibit superior surface properties to support the Ru metal, and this strategy is a potential path to prepare high-performance ammonia decomposition catalysts.
In this study, the Ce-BPDC was synthesized as the precursor, followed by calcination in air and reduction in hydrogen atmosphere to obtain CeO2 with abundant oxygen vacancies. The number of oxygen vacancies on the surface of CeO2 were adjusted by controlling the reduction time. Then, the Ru catalysts supported on these CeO2 were prepared by impregnation, and their catalytic performances on ammonia decomposition were evaluated. Furthermore, these catalysts were comprehensively characterized to explore the relationship between the structure and catalytic properties, as well as the reaction mechanism.

2. Results and Discussion

2.1. Structural Characterization of Carrier and Catalyst

The characterization results of Ce-BPDC are consistent with the results in the literature (see Supporting Information for detailed description). The PXRD results of CeO2-t obtained by calcining Ce-BPDC at 500 °C and then reducing in the H2/Ar atmosphere for different times (t = 0, 0.5, 1, 1.5, 2, 3, 4 h) are given in Figure S4. The pattern of the CeO2 standard sample showed diffraction peaks at 28.5, 33.1, 47.5, 56.3, 59.1, 69.4, 76.7, 79.1, and 87.4, corresponding with a face-centered cubic phase of the CeO2 fluorite structure. Compared with that of CeO2, the diffraction peaks of CeO2-t shifted toward a higher angle, which could be attributed to the transformation of Ce4+ to Ce3+ during the calcination process of Ce-BPDC [41]. The N2 adsorption/desorption isotherms of CeO2-t are given in Figure S5. The specific surface area of cerium oxide obtained by the calcination of Ce-BPDC decreased significantly from 1502 m2 g−1 to about 45 m2 g−1. After loading Ru, the specific surface area of the catalyst did not change significantly. The morphology of CeO2-t was characterized by SEM and TEM (Figures S6 and S7). The results show that the morphology of CeO2-t after calcination was basically similar to that of the Ce-BPDC precursor, while its crystal particle size decreases.
The Ru/CeO2-t catalysts were prepared by the impregnation method with RuCl3, and their PXRD patterns are shown in Figure 1. It was found that the face-centered cubic phase of CeO2 fluorite structure was retained, and no characteristic diffraction peak of Ru was observed, indicating that Ru was highly dispersed on the surface of the support [42]. The Ru content in these catalysts was characterized through Inductively Coupled Plasma Optical Emission Spectrometry (ICP-OES). The results show that the Ru loading of Ru/CeO2-t (t = 0, 0.5, 1, 1.5, 2, 3, 4 h) was 4.29, 4.42, 4.34, 4.46, 4.20, 4.32, and 4.19 wt%, respectively. In order to have a deeper understanding of the catalyst microstructure and the dispersion of Ru on the CeO2 support, detailed observations were made using HRTEM. The lattice stripes of Ru and CeO2 can be observed in Figure S8. The spacing of 0.31 nm corresponds to the (111) face of CeO2. After doping Ru, a small number of Ru nanoparticles was observed on the surface of CeO2, and its lattice spacing was about 0.21 nm, corresponding to the (101) face of Ru substance. The EDS (Figure S8) results show that Ru was evenly dispersed on the surface of CeO2-t. The particle size of Ru in the catalyst Ru/CeO2-t (t = 0, 0.5, 1, 1.5, 2, 3, 4 h) was mainly concentrated between 1.5 and 4 nm (Figure S9).

2.2. Comparison of Catalytic Performance of Ammonia Decomposition to Hydrogen

The ammonia decomposition performance of the catalyst was tested in a self-built fixed-bed reactor (Figure S10), where 0.05 g catalyst was mixed with 2.95 g quartz sand and then transferred to a reactor equipped with quartz cotton. By injecting 50% NH3/Ar gas, the system temperature was adjusted to the corresponding reaction temperature. The catalyst was evaluated at a certain gas hourly space velocity (GHSV) in the temperature range of 375–500 °C. The detailed procedures are listed in the Supporting Information.
Under the GHSV condition of 12,000 mL gcat−1 h−1, the catalytic activities of Ru/CeO2-t and Ru/CeO2-C (Ru supported on the commercial CeO2) catalysts in NH3 decomposition were compared and analyzed, as presented in Table 1. It was found that, compared with Ru/CeO2-C, the catalytic performance of the MOF-derived CeO2-supported Ru catalyst significantly improved. In addition, the ammonia decomposition performance increased significantly with the increase in the reaction temperature. Meanwhile, the catalytic activity also increased on increasing the reduction time, during which the catalytic activity of Ru/CeO2-4 h is the highest. The ammonia conversion rate of Ru/CeO2-4 h reached 98.38% at 475 °C and 100% at 500 °C, respectively. In order to evaluate the practical application potential of Ru/CeO2-4 h catalyst, the ammonia decomposition performance at high GHSV (36,000 mL gcat−1 h−1) was evaluated. And the results are shown in Figure S11. It can be seen that Ru/CeO2-4 h still maintained a high catalytic activity, and the ammonia conversion rate was up to 97.04% at 475 °C and 98.9% at 500 °C, respectively. The hydrogen production rate was calculated according to the ammonia conversion of the catalyst (described in Section 3.3). Under the condition of GHSV = 36,000 mL gcat−1 h−1, the hydrogen production rate of Ru/CeO2-4 h was 38.99 mmol gcat−1 min−1 at 475 °C and 39.74 mmol gcat−1 min−1 at 500 °C, respectively. Under similar conditions, the synthesized Ru/CeO2-t catalyst was compared with the reported Ru catalysts, as shown in Table S2. It is worth noting that the catalytic activity of Ru/CeO2-4 h remained in the first echelon.
In order to evaluate the long-term stability of the catalyst, the Ru/CeO2-4 h catalyst was subjected to a 50 h stability test under GHSV = 36,000 mL gcat−1 h−1. The results are shown in Figure S12. No catalytic performance attenuation was observed after the 50 h test, which demonstrated that Ru/CeO2-4 h had excellent catalytic stability in the ammonia decomposition reaction.

2.3. Surface Chemical State of the Carrier and the Catalyst

It has been reported that hydrogen can react with lattice oxygen (OL) in CeO2 to form H2O, while inducing the reduction of Ce4+ to Ce3+ and generating oxygen vacancies (OV) [43]. The existence of OV can enhance the SMSI between Ru and CeO2 support, and perhaps, this is the reason that Ru/CeO2-4 h exhibits excellent catalytic activity and stability. To systematically investigate the existence of OV on the surface of CeO2-t, Raman spectroscopy and the XPS test were performed.
OV can cause lattice distortions that create new characteristic peaks in the Raman spectrum or change the position and intensity of existing peaks [44]. Thus, Raman spectroscopy was used to reveal the defect location of the samples in this work, and the results are shown in Figure 2. The Raman peaks of CeO2-t and Ru/CeO2-t at 456 and 605 cm−1 can be attributed to the octahedral symmetric tensile vibration mode (F2g) and the defect induction mode (D) [45]. Since the D-mode peak is caused by the presence of Ce3+, the intensity ratio of the D-peak to the F2g peak (ID/IF2g) can be used to reflect the relative concentration of OV in CeO2 and Ru/CeO2 [46]. The results are given in Table S3. The ID/IF2g value increased on increasing the reduction time, indicating that more OV were formed on the surface of CeO2. The Raman spectrum of CeO2-C is shown in Figure S13. It was found that CeO2 prepared from the Ce-BPDC precursor had greater reducibility in a H2/Ar atmosphere than commercial CeO2, resulting in more OV generation during the reduction process.
In addition, the surface electronic states and chemical compositions of Ru/CeO2-t and the carrier CeO2 were studied by XPS. Ru 3p XPS measurements were performed to study the Ru valence states (Figure S14), which exhibited characteristic peaks of Ru0 (462.2 eV) and Ru4+ (465.2 eV). The different chemical valence states of Ru on these samples may be due to the charge transfer between CeO2 carriers and Ru nanoparticles. In order to verify this, the Ce 3d spectra were recorded. The Ce 3d spectra of CeO2-t are shown in Figure S15, and the Ce 3d spectra of Ru/CeO2-t are shown in Figure 3a, which were divided into 10 groups due to the hybridization of Ce 4f orbitals with O 2p valence bands [47]. The six peaks at 882.4, 889.4, 898.4, 901, 907.3, and 916.9 eV are attributed to the Ce4+ species, while the other four peaks at 881.4, 885.4, 899, and 903.6 eV are attributed to the Ce3+ species [48]. Each cerium cation is coordinated by eight oxygen anions, and due to the electronic structure of cerium, charge is reversibly transferred between Ce4+ and Ce3+ [49]. Thus, the emergence of Ce3+ species is usually accompanied by the formation of OV on the CeO2 surface [50]. The concentration of OV can be inferred from the relative atomic ratio of Ce3+/(Ce4+ + Ce3+), as shown in Table S4. It can be observed that the concentration of OV on the surface of CeO2-t and Ru/CeO2-t increased with an increase in the reduction time, which is consistent with the Raman results. Among these catalysts, the Ce3+/(Ce4+ + Ce3+) ratio of Ru/CeO2-4 h was the highest, indicating that the most OV was generated in Ru/CeO2-4 h, which accorded with the test results of the catalyst performance.
To further confirm the variation law of OV, the O 1s XPS spectra of Ru/CeO2-t catalysts were collected (Figure 3b). The peak at 529.1–529.6 eV corresponds to the OL, the peak at 531.1 eV corresponds to OV, and the peak observed at 533.4 eV is attributed to chemisorbed oxygen (OC) in CeO2 [51]. The number of OV can be quantified according to the ratio of OV/(OL + OC + OV) [52]. The results are shown in Table S4. It decreased in the order of Ru/CeO2-4 h > Ru/CeO2-3 h > Ru/CeO2-2 h > Ru/CeO2-1.5 h > Ru/CeO2-1 h > Ru/CeO2-0.5 h, indicating that the number of OV at Ru/CeO2-4 h was the highest. The concentration of OV on the CeO2-t surface showed the same trend. The results were consistent with the Raman results and the Ce 3d XPS spectra.
Figure S16 presents the Ce 3d and O 1s XPS spectra of hydrogen-reduced commercial CeO2, with the quantified Ce3+/(Ce3+ + Ce4+) and OV/(OV + OL + OC) ratios summarized in Table S5. The results (Tables S4 and S5) demonstrate that CeO2 derived from Ce-BPDC displayed a superior reducibility. This result is consistent with the Raman result.
According to the literature, the surface basicity of the catalyst is conducive to ammonia decomposition; usually, the stronger the surface basicity, the higher the activity [53]. CO2-TPD was used to characterize the distribution of the surface basicity of the catalyst, and the results are shown in Figure S17. The number and intensity of basic sites can be determined according to the area and location of the desorption peaks. A certain number of weakly basic sites, moderately strong basic sites, and strong basic sites appeared in these catalysts. The area above 500 °C was a strong basicity site; the basicity strength increased with the increase in the reduction time. The desorption amount of CO2 on the surface of Ru/CeO2-4 h was the largest. From the above analysis, it can be concluded that the density of strong basic sites of Ru/CeO2-4 h was the largest, which is also in good agreement with the ammonia decomposition activity of the catalyst. As an electron donor, OV increases the electron density of the adjacent metal site and enhances the Lewis basic site. On the surface of CeO2, oxygen ions near the OV can act as Lewis base sites. The increase in OV could increase the basicity sites on the surface of CeO2 to a certain extent [54]. Therefore, the concentration of OV on the surface of Ru/CeO2-4 h catalyst is the highest. This was consistent with the Raman and XPS results.
To further clarify the electronic states and coordination environment evolution of Ru species, we conducted Ru K-edge X-ray absorption fine structure (XAFS) tests on the Ru/CeO2-t (t = 0, 0.5, 1, 2, 3, 4 h) catalysts and used Ru foil (Ru0) and RuO2 (Ru4+) as reference standards. As shown in Figure S18a, the absorption edge positions of all the Ru/CeO2-t samples were similar to those of the RuO2 standard samples, confirming that Ru mainly existed in the +4 valence state (Ru4+). Figure S18b shows the Fourier transform Ru K-edge extended X-ray absorption fine structure (EXAFS) curves of all the samples. The figure shows two different peaks, 1.41 Å (the first shell) and 2.43 Å (the second shell), corresponding to Ru-O and Ru-Ru coordination, respectively. Then, a wavelet transform (WT) analysis was conducted on the EXAFS data, thereby further exploring in detail the contributions of different coordination shells to the Ru/CeO2-t EXAFS signal. In Figure 4, two main maximum intensities can be observed, which belong to the first coordination shell of Ru-O-Ce and the second coordination shell of Ru-Ru, respectively. The results show that Ru was stably anchored on the carrier surface through the Ru-O-Ce bond, and it was found that, with the increase in the reduction time, Ru-Ru gradually weakened, while the strength of the Ru-O-Ce bond increased. This further indicates that the interaction between Ru and the carrier increases with an increase in the reduction time.

2.4. Reaction Mechanism

To gain a deeper understanding of the OV impact on ammonia decomposition, we conducted density functional theory (DFT) calculations and developed three reaction models: CeO2, Ru/CeO2, and Ru/CeO2-OV (Figure S19). Figures S21 and S22 illustrate the reaction process of ammonia on the catalyst surface. The NH3 decomposition reaction follows a well-defined pathway: initially, NH3 is adsorbed onto the catalyst surface; subsequently, it undergoes a gradual dehydrogenation process; finally, N2 and H2 are generated and released from the surface.
As shown in Figure 5, the dehydrogenation potential energy of Ru/CeO2-OV was lower than that of CeO2 and Ru/CeO2, which is due to the stronger adsorption of NH3 by this catalyst. It can be seen from the figure that the breaking of the N-H bond is the rate-determining step of the reaction. The reaction energy barrier required for the N-H cleavage of the Ru/CeO2-OV system is lower than that of Ru/CeO2 and CeO2, thereby significantly improving the catalytic activity. Furthermore, Figure S20 compares the adsorption energy data of NH3 on the three models. The results show that the Ru/CeO2-OV model had the highest adsorption intensity for NH3, reaching 1.021 eV, while the adsorption energies of the Ru/CeO2 and CeO2 models were 0.813 eV and 0.605 eV, respectively. This discovery clearly indicates that the presence of oxygen vacancies can reduce the reaction energy barrier for N-H bond breaking and significantly enhance the adsorption capacity of the catalyst for NH3. This phenomenon is due to the strong interaction between the metal and the carrier, which helps to improve the catalytic performance.

3. Experimental Section

3.1. Materials

Ammonium cerium nitrate (Ce(NH4)2(NO3)6, ≥98%), 4-4’ diphenyl dicarboxylic acid (H2BPDC, 99%), N, N-dimethylformamide (C3H7NO, DMF, ≥99.5%), acetone (C3H6O, ≥99.5%), anhydrous ethanol (C2H5OH, ≥99.7%), ruthenium trichloride (RuCl3, ≥99.9%) purchased from Aladdin Co., Ltd., and deionized water for laboratory use.

3.2. Catalyst Preparation

Ce-BPDC was synthesized by an improved hydrothermal method. The detailed preparation process is in the Supporting Information. After the successful synthesis of Ce-BPDC, it was calcined at 500 °C for 5 h to obtain CeO2. The CeO2 was reduced in the H2/Ar atmosphere for different durations (t = 0, 0.5, 1, 1.5, 2, 3, 4 h; 0 means the CeO2 sample was not reduced) at 500 °C, and the obtained samples were recorded as CeO2-t. The detailed preparation process of the carrier can be found in the Supporting Information. Briefly, 1 g of CeO2-t (t = 0, 0.5, 1, 1.5, 2, 3, 4 h) and 0.108 g (0.52 mmol) of RuCl3, respectively, were weighed and mixed in different glass bottles. Subsequently, 15 mL of deionized water was added to each vial and stirred on a mixing table for 8 h to obtain a gray solution. The solution was washed three times with deionized water, and the resulting precipitate was dried overnight in an 80 °C oven and finally allowed to yield the Ru catalyst. The experimental amount of Ru for each catalyst was 5 wt%. The resulting sample was named as the Ru/CeO2-t (t = 0, 0.5, 1, 1.5, 2, 3, 4 h) catalyst.

3.3. Catalyst Activity Evaluation

The hydrogen production evaluation process of ammonia decomposition catalyst was carried out in a fixed-bed reactor. Typically, 0.05 g of catalyst (40–60 mesh) was fully mixed with 2.95 g of quartz sand (40–60 mesh) and then transferred to a reactor equipped with quartz cotton. Before the activity test, the catalyst was heated to 200 °C with a ramp rate of 5 °C/min in 5% H2/Ar (50 mL/min) and reduced at 200 °C for 2 h. After switching to 50% NH3/Ar, the system temperature was adjusted to the corresponding reaction temperature. The catalyst was evaluated at a certain gas hourly space velocity (GHSV) in the temperature range of 325–600 °C at atmospheric pressure. The feed gas and the yield were analyzed by online gas chromatography (GC) equipped with a thermal conductivity detector (TCD). The conversion rate of NH3 (XNH3) and the generation rate of H2 (rH2) were calculated using the following formula:
X N H 3 ( % ) = [ V N 2 ] o u t / [ V N H 3 ] o u t [ V N 2 ] o u t / [ V N H 3 ] o u t + 0.5 × 100 %
r H 2 ( m m o l / g c a t / m i n ) = V N H 3 22.4 × X N H 3 × 1.5 m c a t
where [VN2]out, and [VNH3]out are the volume percentage of N2 and NH3 in the effluent, respectively. VNH3 is the NH3 flow rate (mL/min), and mcat is the mass of the catalyst (g).

4. Conclusions

In this study, Ce-MOF-derived CeO2 carriers were prepared by calcining Ce-BPDC, and then the obtained CeO2 was reduced in a H2/Ar atmosphere for different times (t = 0, 0.5, 1, 1.5, 2, 3, 4 h). Finally, a series of Ru-based catalysts were prepared by the impregnation method. The ammonia decomposition results showed that the performance of the Ru/CeO2-t was superior to that of the commercial CeO2-supported Ru catalyst. Under the condition of GHSV = 36,000 mL gcat−1 h−1, the ammonia conversion rate and hydrogen production rate of Ru/CeO2-4 h at 500 °C were 98.9% and 39.74 mmol gcat−1 min−1, respectively. Furthermore, the catalytic activity remained stable after continuous testing for 50 h. The XRD, SEM, and TEM results showed that Ru was well dispersed on the surface of CeO2-t. The Raman and XPS characterization results showed that, with the extension of the reduction time, the number of oxygen vacancies in these samples increased. This phenomenon can be attributed to the valence state transformation from Ce4+ to Ce3+ on the surface of CeO2 during the hydrogen reduction process, thereby inducing the generation of more OV. The results of the synchrotron radiation showed that an increase in oxygen vacancy concentration can enhance the interaction between the metal and the carrier. The DFT calculation determined that the rate-determining step of ammonia decomposition was the cleavage of the N-H bond, and the existence of oxygen vacancies can significantly reduce the reaction energy barrier of N-H cleavage, thereby improving the ammonia decomposition performance. These findings highlight a new idea for the design of ammonia decomposition catalysts and might open up new possibilities for the development of MOF-based catalysts in industrial application.

Supplementary Materials

The following Supporting Information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30112301/s1: Table S1: NH3 conversion of Ru/CeO2-C catalyst at GHSV = 12,000 mL gcat−1 h−1. Table S2: Comparison of the synthesized catalysts with Ru-based catalysts reported in the literature. Table S3: The Raman quantification results. Table S4: XPS quantitative results of CeO2-t and Ru/CeO2-t. Table S5: XPS quantitative results of CeO2-C and Ru/CeO2-C. Figure S1: The XRD profile of the Ce-BPDC (a) and the SEM images of the Ce-BPDC (b). Figure S2: N2 adsorption/desorption isotherms (a) and aperture distribution curve (b) of Ce-BPDC. Figure S3: The TGA curve of the Ce-BPDC. Figure S4: The XRD profile of the CeO2-t. Figure S5: The N2 adsorption isotherm of the derivative cerium oxide (a) and its corresponding catalyst (b). Figure S6: The SEM of the CeO2-0 h (a), CeO2-0.5 h (b), CeO2-1 h (c), CeO2-1.5 h (d), CeO2-2 h (e), CeO2-3 h (f). Figure S7: The TEM of the CeO2-t. Figure S8: The TEM and the corresponding EDS of Ru/CeO2-t. Figure S9: The AC-STEM of Ru/CeO2-0 h (a), Ru/CeO2-0.5 h (b), Ru/CeO2-1 h (c), Ru/CeO2-1.5 h (d), Ru/CeO2-2 h (e), Ru/CeO2-3 h(f), Ru/CeO2-4 h(g). Figure S10: Schematic diagram of a fixed-bed reactor. Figure S11: NH3 conversion diagram of Ru/CeO2-4 h catalyst, GHSV = 36,000 mL gcat−1 h−1. Figure S12. Stability test of Ru/CeO2-4 h catalyst, GHSV = 36,000 mL gcat−1 h−1. Figure S13: Raman spectra of CeO2-C (a) and Ru/CeO2-C (b). Figure S14: The Ru 3p3/2 XPS of Ru/CeO2-t. Figure S15: The (a) Ce 3d and (b) O 1s XPS of CeO2-t. Figure S16: The (a) Ce 3d and (b) O 1s XPS of CeO2-C, (c) Ce 3d and (d) O 1s XPS of Ru/CeO2-C. Figure S17: CO2-TPD curves for catalyst. Figure S18: XAS characterization results: (a) Ru K-edge XANES spectra of Ru/CeO2-t and (b) Fourier transform K3-weighted EXAFS spectra. Figure S19: Adsorption energy of NH3 on surfaces of CeO2, Ru/CeO2, and Ru/CeO2-Ov models. Figure S20: The reaction process of ammonia on Ru/CeO2 catalyst surface. Figure S21: The reaction process of ammonia on Ru/CeO2 catalyst surface. Figure S22: The reaction process of ammonia on Ru/CeO2-OV catalyst surface. Figure S23: Ce 3d (a) and O 1s (b) spectra of catalyst Ru/CeO2-2 h before and after pretreatment. References [55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76] are cited in the supplementary materials.

Author Contributions

Literature search, writing, and creation of figures and tables, W.W., W.Y., and Y.L.; review, editing, and supervision, S.X. and T.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, Grant No. 22102008.

Data Availability Statement

The data presented in this study are available upon request from the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sun, S.; Jiang, Q.; Zhao, D.; Cao, T.; Sha, H.; Zhang, C.; Song, H.; Da, Z. Ammonia as hydrogen carrier: Advances in ammonia decomposition catalysts for promising hydrogen production. Renew. Sustain. Energy Rev. 2022, 169, 112918. [Google Scholar] [CrossRef]
  2. Mukherjee, S.; Devaguptapu, S.V.; Sviripa, A.; Lund, C.R.F.; Wu, G. Low-temperature ammonia decomposition catalysts for hydrogen generation. Appl. Catal. B Environ. 2018, 226, 162–181. [Google Scholar] [CrossRef]
  3. Andersson, J.; Grönkvist, S. Large-scale storage of hydrogen. Int. J. Hydrogen Energy 2019, 44, 11901–11919. [Google Scholar] [CrossRef]
  4. Feng, J.; Zhang, X.; Wang, J.; Ju, X.; Chen, P. Applications of rare earth oxides in catalytic ammonia synthesis and decomposition. Catal. Sci. Technol. 2021, 11, 6330–6343. [Google Scholar] [CrossRef]
  5. Morlane, N.; Katikaneni, S.P.; Paglieri, S.N.; Harale, A.; Solami, B.; Sarathy, S.M.; Gascon, J. A technological roadmap to the ammonia energy economy: Current state and missing technologies. Chem. Eng. J. 2021, 408, 127310. [Google Scholar] [CrossRef]
  6. Wan, Z.; Tao, Y.; Shao, J.; Zhang, Y.; You, H. Ammonia as an effective hydrogen carrier and a clean fuel for solid oxide fuel cells. Energy Convers. Manag. 2021, 228, 113729. [Google Scholar] [CrossRef]
  7. Bell, T.E.; Torrente-Murciano, L. H2 production via ammonia decomposition using non-noble metal catalysts: A review. Top. Catal. 2016, 59, 1438–1457. [Google Scholar] [CrossRef]
  8. Dongil, A.B. Recent progress on transition metal nitrides nanoparticles as heterogeneous catalysts. Nanomaterials 2019, 9, 1111. [Google Scholar] [CrossRef]
  9. Cha, J.; Lee, T.; Lee, Y.-J.; Jeong, H.; Jo, Y.S.; Kim, Y.; Nam, S.W.; Han, J.; Lee, K.B.; Yoon, C.W.; et al. Highly monodisperse sub-nanometer and nanometer Ru particles confined in alkali-exchanged zeolite Y for ammonia decomposition. Appl. Catal. B Environ. 2021, 283, 119627. [Google Scholar] [CrossRef]
  10. Wu, Z.-W.; Li, X.; Qin, Y.-H.; Deng, L.; Wang, C.-W.; Jiang, X. Ammonia decomposition over SiO2-supported Ni-Co bimetallic catalyst for COx-free hydrogen generation. Int. J. Hydrogen Energy 2020, 45, 15263–15269. [Google Scholar] [CrossRef]
  11. Hu, Z.P.; Chen, L.; Chen, C. Fe/ZSM-5 catalysts for ammonia decomposition to COx-free hydrogen: Effect of SiO2/Al2O3 ratio. Mol. Catal. 2018, 455, 14–22. [Google Scholar] [CrossRef]
  12. Cao, C.F.; Wu, K.; Zhou, C.; Yao, Y.H. Electronic metal-support interaction enhanced ammonia decomposition efficiency of perovskite oxide supported ruthenium. Chem. Eng. Sci. 2022, 257, 117719–117726. [Google Scholar] [CrossRef]
  13. Antunes, R.; Steiner, R.; Marot, L. Decomposition studies of NH3 and ND3 in presence of H2 and D2 with Pt/Al2O3 and Ru/Al2O3 catalysts. Int. J. Hydrogen Energy 2022, 47, 14130–14140. [Google Scholar] [CrossRef]
  14. Pinzón, M.; Avilés-García, O.; Osa, A.R. New catalysts based on reduced graphene oxide for hydrogen production from ammonia decomposition. Sustain. Chem. Pharm. 2022, 25, 100615–100625. [Google Scholar] [CrossRef]
  15. Mazzone, S.; Goklany, T.; Zhang, G. Ruthenium-based catalysts supported on carbon xerogels for hydrogen production via ammonia decomposition. Appl. Catal. A Gen. 2022, 632, 118484–118499. [Google Scholar] [CrossRef]
  16. Yin, S.F.; Xu, B.Q.; Wang, S.J.; Ng, C.F.; Au, C.T. Magnesia-Carbon Nanotubes (MgO-CNTs) Nanocomposite: Novel Support of Ru Catalyst for the Generation of COx-Free Hydrogen from Ammonia. Catal. Lett. 2004, 96, 113–116. [Google Scholar] [CrossRef]
  17. Choudhary, T.V.; Sivadinarayana, C.; Goodman, D.W. Catalytic ammonia decomposition: COX-free hydrogen production for fuel cell applications. Catal. Lett. 2001, 72, 197–201. [Google Scholar] [CrossRef]
  18. Fang, H.; Wu, S.; Ayvali, T.; Zheng, J.; Fellowes, J.; Ho, P.-L.; Leung, K.C.; Large, A.; Held, G.; Kato, R.; et al. Dispersed surface Ru ensembles on MgO(111) for catalytic ammonia decomposition. Nat. Commun. 2023, 14, 647. [Google Scholar] [CrossRef]
  19. Yin, S.F.; Zhang, Q.H.; Xu, B.Q.; Zhu, W.X.; Ng, C.F.; Zhou, X.P.; Au, C.T. Carbon nanotubes-supported Ru catalyst for the generationof COx-free hydrogen from ammonia. Catal. Today 2004, 93, 27–38. [Google Scholar] [CrossRef]
  20. Ju, X.; Liu, L.; Xu, X.; Wang, J.; He, T.; Chen, P. Strong metal-support interaction modulatescatalytic activity of Ru nanoparticles on Gd2O3 forefficient ammonia decomposition. iScience 2024, 27, 110931. [Google Scholar] [CrossRef]
  21. Wang, Z.; Luo, H.; Wang, L.; Li, T.; Li, S.; Liu, Y.Q. Promotion of Low-Temperature Catalytic Activity of Ru-Based Catalysts for Ammonia Decomposition via Lanthanum and Cesium Codoping. ACS Sustainable Chem. Eng. 2024, 12, 5620–5631. [Google Scholar] [CrossRef]
  22. Im, Y.; Muroyama, H.; Matsui, T.; Eguchi, K. Investigation on catalytic performance and desorption behaviors of ruthenium catalysts supported on rare-earth oxides for NH3 decomposition. Int. J. Hydrogen Energy 2022, 47, 32543–32551. [Google Scholar] [CrossRef]
  23. Furusawa, T.; Kuribara, H.; Kimura, K.; Sato, T.; Itoh, N. Development of a Cs-Ru/CeO2 Spherical Catalyst Prepared by Impregnation and Washing Processes for Low-Temperature Decomposition of NH3: Characterization and Kinetic Analysis Results. Ind. Eng. Chem. Res. 2020, 59, 18460–18470. [Google Scholar] [CrossRef]
  24. Huang, C.; Yu, Y.; Tang, X.; Liu, Z.; Zhang, J.; Ye, C.; Ye, Y.; Zhang, R. Hydrogen generation by ammonia decomposition over Co/CeO2 catalyst: Influence of support morphologies. Appl. Surf. Sci. 2020, 532, 147335. [Google Scholar] [CrossRef]
  25. Lucentini, I.; Serrano, I.; Soler, L.; Divins, N.J.; Llorca, J. Ammonia decomposition over 3D-printed CeO2 structures loaded with Ni. Appl. Catal. A Gen. 2020, 591, 117382. [Google Scholar] [CrossRef]
  26. Ju, X.; Liu, L.; Yu, P.; Guo, J.; Zhang, X.; He, T.; Wu, G.; Chen, P. Mesoporous Ru/MgO prepared by a deposition-precipitation method as highly active catalyst for producing COx-free hydrogen from ammonia decomposition. Appl. Catal. B Environ. 2017, 211, 167–175. [Google Scholar] [CrossRef]
  27. Ju, X.; Liu, L.; Zhang, X.; Feng, J.; He, T.; Chen, P. Highly Efficient Ru/MgO Catalyst with Surface-Enriched Basic Sites for Production of Hydrogen from Ammonia Decomposition. ChemCatChem 2019, 11, 4161–4170. [Google Scholar] [CrossRef]
  28. Muroyama, H.; Saburi, C.; Matsui, T.; Eguchi, K. Ammonia decomposition over Ni/La2O3 catalyst for on-site generation of hydrogen. Appl. Catal. A Gen. 2012, 443–444, 119–124. [Google Scholar] [CrossRef]
  29. Gu, Y.; Ma, Y.; Long, Z.; Zhao, S.; Wang, Y.; Zhang, W. One-pot synthesis of supported Ni@Al2O3 catalysts with uniform small-sized Ni for hydrogen generation via ammonia decomposition. Int. J. Hydrogen Energy 2021, 46, 4045–4054. [Google Scholar] [CrossRef]
  30. Bell, T.E.; Ménard, H.; González Carballo, J.M.; Tooze, R.; Torrente-Murciano, L. Hydrogen production from ammonia decomposition using Co/γ-Al2O3 catalysts—Insights into the effect of synthetic method. Int. J. Hydrogen Energy 2020, 45, 27210–27220. [Google Scholar] [CrossRef]
  31. Zhang, Z.S.; Fu, X.P.; Wang, W.W.; Jin, Z.; Song, Q.S.; Jia, C.J. Promoted porous Co3O4-Al2O3 catalysts for ammonia decomposition. Sci. China Chem. 2018, 61, 1389–1398. [Google Scholar] [CrossRef]
  32. Yin, S.F.; Zhang, Q.H.; Xu, B.Q.; Zhu, W.X.; Ng, C.F.; Au, C.T. Investigation on the catalysis of COx-free hydrogen generation from ammonia. J. Catal. 2004, 224, 384–396. [Google Scholar] [CrossRef]
  33. Lorenzut, B.; Montini, T.; Pavel, C.C.; Comotti, M.; Vizza, F.; Bianchini, C.; Fornasiero, P. Embedded Ru@ZrO2 Catalysts for H2 Production by Ammonia Decomposition. ChemCatChem 2010, 2, 1096–1106. [Google Scholar] [CrossRef]
  34. Thien, A.L.; Youngmin, K.; Hyun, W.K. Ru-supported lanthania-ceria composite as an efficient catalyst for COx-free H2 production from ammonia decomposition. Appl. Catal. B Environ. 2021, 285, 119831. [Google Scholar]
  35. Hu, X.-C.; Fu, X.-P.; Wang, W.-W.; Wang, X.; Wu, K.; Si, R.; Ma, C.; Jia, C.-J.; Yan, C.-H. Ceria-supported ruthenium clusters transforming from isolated single atoms for hydrogen production via decomposition of ammonia-ScienceDirect. Appl. Catal. B Environ. 2020, 268, 118424. [Google Scholar] [CrossRef]
  36. Chaouiki, A.; Fatimah, S.; Chafiq, M.; Ryu, J.; Ko, Y.G. State-of-the-art advancements in metal-organic framework nanoarchitectures for catalytic applications. Appl. Mater. Today 2024, 38, 102224. [Google Scholar] [CrossRef]
  37. Khalil, I.E.; Fonseca, J.; Reithofer, M.R. Tackling orientation of metal-organic frameworks (MOFs): The quest to enhance MOF performance. Coordin. Chem. Rev. 2023, 481, 21503. [Google Scholar] [CrossRef]
  38. Sivan, S.E.; Kang, K.H.; Han, S.J. Facile MOF-derived one-pot synthetic approach toward Ru single atoms, nanoclusters, and nanoparticles dispersed on CeO2 supports for enhanced ammonia synthesis. J. Catal. 2022, 408, 316–328. [Google Scholar] [CrossRef]
  39. He, Y.; Zheng, X.; Mao, D.; Meng, T.; Mao, H.; Yu, J. Promoting catalytic CO2 methanation using Ru catalyst supported on Ce-MOF-derived CeO2. Renew. Energy 2025, 245, 122834. [Google Scholar] [CrossRef]
  40. Chen, X.; Yu, E.; Cai, S.; Jia, H.; Chen, J.; Liang, P. In situ pyrolysis of Ce-MOF to prepare CeO2 catalyst with obviously improved catalytic performance for toluene combustion. Chem. Eng. J. 2018, 344, 469–479. [Google Scholar] [CrossRef]
  41. Khan, M.E.; Khan, M.M.; Cho, M.H. Ce3+-ion, Surface Oxygen Vacancy, and Visible Light-induced Photocatalytic Dye Degradation and Photocapacitive Performance of CeO2-Graphene Nanostructures. Sci. Rep. 2017, 7, 5928. [Google Scholar] [CrossRef] [PubMed]
  42. Li, J.; Liu, Z.; Cullen, D.A. Distribution and Valence State of Ru Species on CeO2 Supports: Support Shape Effect and Its Influence on CO Oxidation. ACS Catal. 2019, 9, 11088–11103. [Google Scholar] [CrossRef]
  43. 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]
  44. Wang, X.; Yin, Y.; Wang, H. Precision tailoring strategy of oxygen vacancies for electromagnetic pollution regulation. Appl. Surf. Sci. 2025, 681, 20–25. [Google Scholar] [CrossRef]
  45. Pu, Z.Y.; Lu, J.Q.; Luo, M.F. Study of oxygen vacancies in Ce0.9Pr0.1O2-δ solid solution by in situ X-ray diffraction and in situ Raman spectro-scopy. J. Phys. Chem. C 2007, 111, 18695–18702. [Google Scholar] [CrossRef]
  46. Wang, F.; He, S.; Chen, H. Active site dependent reaction mechanism over Ru/CeO2 catalyst toward CO2 methanation. J. Am. Chem. Soc. 2016, 138, 6298–6305. [Google Scholar] [CrossRef]
  47. Yoo, C.J.; Lee, D.W.; Kim, M.S. The synthesis of methanol from CO/CO2/H2 gas over Cu/Ce1-xZrxO2 catalysts. J. Mol. Catal. A Chem. 2013, 378, 255–262. [Google Scholar] [CrossRef]
  48. Sakpal, T.L.L. Structure-dependent activity of CeO2 supported Ru catalysts for CO2 methanation. J. Catal. 2018, 367, 171–180. [Google Scholar] [CrossRef]
  49. Huang, X.; Zhang, K.; Peng, B. Ceria-Based Materials for Thermocatalytic and Photocatalytic Organic Synthesis. ACS Catal. 2021, 11, 9618–9678. [Google Scholar] [CrossRef]
  50. Wang, Z.; Huang, Z.; Brosnahan, J.T. Ru/CeO2 Catalyst with Optimized CeO2 Support Morphology and Surface Facet for Propane Combustion. Environ. Sci. Technol. 2019, 53, 5349–5358. [Google Scholar] [CrossRef]
  51. Huang, H.; Dai, Q.; Wang, X. Morphology effect of Ru/CeO2 catalysts for the catalytic combustion of chlorobenzene. Appl. Catal. B Environ. 2014, 158–159, 96–105. [Google Scholar] [CrossRef]
  52. Hu, Z.; Liu, X.; Meng, D. Effect of Ceria Crystal Plane on the Physicochemical and Catalytic Properties of Pd/Ceria for CO and Propane Oxidation. ACS Catal. 2016, 6, 2265–2279. [Google Scholar] [CrossRef]
  53. Le, T.A.; Kim, Y.; Han, S.J.; Do, Q.C.; Kim, G.J.; Im, Y.; Chae, H.-J. Ru dispersed on CeO2{100} facets boosting the catalytic NH3 decomposition for green H2 generation. Chem. Eng. J. 2024, 493, 152503. [Google Scholar] [CrossRef]
  54. Ren, X.; Zhang, Z.; Wang, Y. Capping experiments reveal multiple surface activesites in CeO2 and their cooperative catalysis. RSC Adv. 2019, 9, 15229–15237. [Google Scholar] [CrossRef]
  55. Luo, L.; Huang, L.; Liu, X. Mixed-Valence Ce-BPyDC Metal-Organic Framework with Dual Enzyme-like Activities for Colorimetric Biosensing. Inorg. Chem. 2019, 58, 11382–11388. [Google Scholar] [CrossRef]
  56. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  57. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef]
  58. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  59. Montini, T.; Melchionna, M.; Monai, M. Fundamentals and Catalytic Applications of CeO2-Based Materials. Chem. Rev. 2016, 116, 5987–6041. [Google Scholar] [CrossRef]
  60. Li, X.; Ji, W.; Zhao, J.; Wang, S.; Au, C. Ammonia decomposition over Ru and Ni catalysts supported on fumed SiO2, MCM-41, and SBA-15. J. Catal. 2005, 236, 181–189. [Google Scholar] [CrossRef]
  61. Wang, Z.; Qu, Y.; Shen, X.; Cai, Z. Ruthenium catalyst supported on Ba modified ZrO2 for ammonia decomposition to COx-free hydrogen. Int. J. Hydrogen Energy 2019, 44, 7300–7307. [Google Scholar] [CrossRef]
  62. Feng, J.; Liu, L.; Ju, X.; Wang, J.; Zhang, X.; He, T. Highly dispersed ruthenium nanoparticles on Y2O3 as superior catalyst for ammonia decomposition. ChemCatChem 2021, 13, 1552–1558. [Google Scholar] [CrossRef]
  63. Jeon, N.; Kim, S.; Tayal, A.; Oh, J.; Yoon, W.; Kim, W.B. Y-doped BaCeO3 perovskite-supported Ru catalysts for COx-free hydrogen production from ammonia: Effect of strong metal–support interactions. ACS Sustain. Chem. Eng. 2022, 10, 15564–15573. [Google Scholar] [CrossRef]
  64. Li, Y.; Yao, L.; Song, Y.; Liu, S.; Zhao, J.; Ji, W.; Au, C.T. Core-Shell Structured Microcapsular-like Ru@SiO2 Reactor for Efficient Generation of COx-Free Hydrogen through Ammonia Decomposition. Chem. Commun. 2010, 46, 5298–5300. [Google Scholar] [CrossRef]
  65. Wang, Z.; Cai, Z.; Wei, Z. Highly active ruthenium catalyst supported on barium hexaaluminate for ammonia decomposition to COx-free hydrogen. ACS Sustain. Chem. Eng. 2019, 7, 8226–8235. [Google Scholar] [CrossRef]
  66. Meng, Q.; Liu, H.; Xu, K.; Wang, W.; Jia, C. CeO2 modified Ru/γ-Al2O3 catalysts for ammonia decomposition reaction. J. Rare Earths 2023, 41, 801–809. [Google Scholar] [CrossRef]
  67. Li, L.; Wang, Y.; Xu, Z.P.; Zhu, Z. Catalytic ammonia decomposition for CO-free hydrogen generation over Ru/Cr2O3 catalysts. Appl. Catal. A 2013, 467, 246–252. [Google Scholar] [CrossRef]
  68. Lucentini, I.; Casanovas, A.; Llorca, J. Catalytic ammonia decomposition for hydrogen production on Ni, Ru and NiRu supported on CeO2. Int. J. Hydrogen Energy 2019, 44, 12693–12707. [Google Scholar] [CrossRef]
  69. Lucentini, I.; Colli, G.G.; Luzi, C.D.; Serrano, I.; Martínez, O.M.; Llorca, J. Catalytic ammonia decomposition over Ni-Ru supported on CeO2 for hydrogen production: Effect of metal loading and kinetic analysis. Appl. Catal. B Environ. 2021, 286, 119896. [Google Scholar] [CrossRef]
  70. Kim, A.; Cha, J.; Kim, J.S.; Ahn, C.; Kim, Y.; Jeong, H. Hydrogen production from ammonia decomposition over Ru-rich surface on La2O2CO3-Al2O3 catalyst beads. Catal. Today 2023, 411, 113867. [Google Scholar] [CrossRef]
  71. Huang, C.; Yu, Y.; Yang, J.; Yan, Y.; Wang, D.; Hu, F. Ru/La2O3 catalyst for ammonia decomposition to hydrogen. Appl. Surf. Sci. 2019, 476, 928–936. [Google Scholar] [CrossRef]
  72. Yin, S.; Xu, B.; Ng, C.; Au, C. Nano Ru/CNTs: A highly active and stable catalyst for the generation of COx-free hydrogen in ammonia decomposition. Appl. Catal. B Environ. 2004, 48, 237–241. [Google Scholar] [CrossRef]
  73. Chung, D.B.; Kim, H.Y.; Jeon, M.; Lee, D.H.; Park, H.S.; Choi, S.H. Enhanced ammonia dehydrogenation over Ru/LaxAl2O3 (x = 0-50 mol%): Structural and electronic effects of La doping. Int. J. Hydrogen Energy 2017, 42, 1639–1647. [Google Scholar] [CrossRef]
  74. Hu, Z.; Mahin, J.; Datta, S.; Bell, T.E.; Torrente-Murciano, L. Ru-based catalysts for H2 production from ammonia: Effect of 1D support. Top. Catal. 2019, 62, 1169–1177. [Google Scholar] [CrossRef]
  75. Li, L.; Zhu, Z.H.; Yan, Z.F.; Lu, G.Q.; Rintoul, L. Catalytic ammonia decomposition over Ru/carbon catalysts: The importance of the structure of carbon support. Appl. Catal. A 2007, 320, 166–172. [Google Scholar] [CrossRef]
  76. Kocer, T.; Oztuna, F.E.S.; Kurtoğlu, S.F.; Unal, U.; Uzun, A. Graphene aerogel-supported ruthenium nanoparticles for COx-free hydrogen production from ammonia. Appl. Catal. A 2021, 610, 117969. [Google Scholar] [CrossRef]
Figure 1. The PXRD profile of the Ru/CeO2-t.
Figure 1. The PXRD profile of the Ru/CeO2-t.
Molecules 30 02301 g001
Figure 2. Raman spectra of various supports (a) and catalysts (b).
Figure 2. Raman spectra of various supports (a) and catalysts (b).
Molecules 30 02301 g002
Figure 3. The (a) Ce 3d and (b) O 1s XPS profiles of Ru/CeO2-t.
Figure 3. The (a) Ce 3d and (b) O 1s XPS profiles of Ru/CeO2-t.
Molecules 30 02301 g003
Figure 4. The WT-EXAFS of the Ru K-edge of Ru/CeO2-t, (a) Ru foil, (b) RuO2, (ch) Ru/CeO2-t (t = 0, 0.5, 1, 2, 3, 4 h).
Figure 4. The WT-EXAFS of the Ru K-edge of Ru/CeO2-t, (a) Ru foil, (b) RuO2, (ch) Ru/CeO2-t (t = 0, 0.5, 1, 2, 3, 4 h).
Molecules 30 02301 g004
Figure 5. Potential energy diagram of NH3 dehydrogenation on the surface of CeO2, Ru/CeO2, and Ru/CeO2-OV models.
Figure 5. Potential energy diagram of NH3 dehydrogenation on the surface of CeO2, Ru/CeO2, and Ru/CeO2-OV models.
Molecules 30 02301 g005
Table 1. NH3 conversion of Ru/CeO2-t catalyst at GHSV = 12,000 mL gcat−1 h−1.
Table 1. NH3 conversion of Ru/CeO2-t catalyst at GHSV = 12,000 mL gcat−1 h−1.
T °C375400425450475500
NH3 Conv./% a
Ru/CeO2-C7.8716.8633.4752.6673.7585.16
Ru/CeO2-0 h23.4145.2664.4478.9591.8895.95
Ru/CeO2-0.5 h30.2455.7973.5386.2492.9596.99
Ru/CeO2-1 h34.9257.9475.5687.7995.797.72
Ru/CeO2-1.5 h38.8561.3778.5789.7797.1199.01
Ru/CeO2-2 h37.9858.977687.8496.6398.13
Ru/CeO2-3 h42.8664.9181.0492.1998.2299.90
Ru/CeO2-4 h43.5865.1081.7492.3998.38100.00
a The calculation formula of NH3 conv./% is   X N H 3 ( % ) = [ N H 3 ] i n [ N H 3 ] o u t [ N H 3 ] o u t × 100 % , which is described in detail in Section 3.3.
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

Wu, W.; Yao, W.; Liu, Y.; Xi, S.; Zhang, T. Efficient Hydrogen Production from Ammonia Using Ru Nanoparticles on Ce-Based Metal–Organic Framework (MOF)-Derived CeO2 with Oxygen Vacancies. Molecules 2025, 30, 2301. https://doi.org/10.3390/molecules30112301

AMA Style

Wu W, Yao W, Liu Y, Xi S, Zhang T. Efficient Hydrogen Production from Ammonia Using Ru Nanoparticles on Ce-Based Metal–Organic Framework (MOF)-Derived CeO2 with Oxygen Vacancies. Molecules. 2025; 30(11):2301. https://doi.org/10.3390/molecules30112301

Chicago/Turabian Style

Wu, Wenying, Wenhao Yao, Yitong Liu, Senliang Xi, and Teng Zhang. 2025. "Efficient Hydrogen Production from Ammonia Using Ru Nanoparticles on Ce-Based Metal–Organic Framework (MOF)-Derived CeO2 with Oxygen Vacancies" Molecules 30, no. 11: 2301. https://doi.org/10.3390/molecules30112301

APA Style

Wu, W., Yao, W., Liu, Y., Xi, S., & Zhang, T. (2025). Efficient Hydrogen Production from Ammonia Using Ru Nanoparticles on Ce-Based Metal–Organic Framework (MOF)-Derived CeO2 with Oxygen Vacancies. Molecules, 30(11), 2301. https://doi.org/10.3390/molecules30112301

Article Metrics

Back to TopTop