Next Article in Journal
NRF2 Activation by Nitrogen Heterocycles: A Review
Next Article in Special Issue
Co-Immobilization of Laccase and Mediator into Fe-Doped ZIF-8 Significantly Enhances the Degradation of Organic Pollutants
Previous Article in Journal
Experimental Study on Electrochemical Desulfurization of Coal Liquefaction Residue
Previous Article in Special Issue
Enhanced Activity of Alkali-Treated ZSM-5 Zeolite-Supported Pt-Co Catalyst for Selective Hydrogenation of Cinnamaldehyde
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Ammonia Decomposition by Tuning the Support Properties of Ni/GdxCe1-xO2-δ at 600 °C

College of Pharmacy, Jinhua Polytechnic, Jinhua 321007, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2750; https://doi.org/10.3390/molecules28062750
Submission received: 20 February 2023 / Revised: 15 March 2023 / Accepted: 16 March 2023 / Published: 18 March 2023
(This article belongs to the Special Issue Catalysts: New Materials for Green Chemistry)

Abstract

:
Ammonia decomposition is a promising method to produce high-purity hydrogen. However, this process typically requires precious metals (such as Ru, Pt, etc.) as catalysts to ensure high efficiency at relatively low temperatures. In this study, we propose using several Ni/GdxCe1-xO2-δ catalysts to improve ammonia decomposition performance by adjusting the support properties. We also investigate the underlying mechanism for this enhanced performance. Our results show that Ni/Ce0.8Gd0.2O2-δ at 600 °C can achieve nearly complete ammonia decomposition, resulting in a hydrogen production rate of 2008.9 mmol.g−1.h−1 with minimal decrease over 150 h. Density functional theory calculations reveal that the recombinative desorption of nitrogen is the rate-limiting step of ammonia decomposition over Ni. Our characterizations indicate that Ni/Ce0.8Gd0.2O2-δ exhibits a high concentration of oxygen vacancies, highly dispersed Ni on the surface, and abundant strong basic sites. These properties significantly enhance the associative desorption of N and strengthen the metal support interactions, resulting in high catalytic activity and stability. We anticipate that the mechanism could be applied to designing additional catalysts with high ammonia decomposition performance at relatively low temperatures.

1. Introduction

Hydrogen is considered an important carrier of energy for the future due to its clean and renewable characteristics [1,2]. However, the low density and high flammability of hydrogen present significant challenges for its storage and transportation, severely hindering the scale-up of hydrogen fuel use [3,4]. To address the above-mentioned issues, researchers have proposed the use of liquid fuels with high energy density, such as methanol [5] and ammonia [6], as carriers for hydrogen [7]. Among all the candidates, ammonia represents a highly promising medium to store and produce hydrogen due to ammonia’s impressive energy density (3000 Wh/kg), ease of liquefaction (~0.8 MPa, 20 °C), low cost, safe storage and transport attributes, and well-established infrastructure [8,9,10]. Consequently, ammonia has gained increasing attention worldwide as a medium for hydrogen storage. An efficient ammonia decomposition reaction is a key factor in hydrogen storage. However, currently the ammonia decomposition reaction requires the use of noble metals (Ru, Pt, etc. [11]) as catalysts to ensure efficient reaction kinetics and high ammonia conversion at relatively low temperatures. The scalable utilization of ammonia as an intermediary to store hydrogen has been severely hindered by the high cost of these metals. Thus, it is crucially important to develop non-precious catalysts that can efficiently decompose ammonia at low temperatures.
Currently, the attention of numerous researchers has been directed toward transition-metal-based catalysts, such as Ni [12], Fe [13], and Co [14], for ammonia decomposition. Within these transition metals, Ni is the most active and shows high potential for practical application [15]. Nowadays, Ni has been widely recognized as an alternative to Ru-based catalysts, and it has been preliminary investigated [16]. However, Ni is not active enough, and its performance requires further improvement at medium or low temperatures. For example, in a recent study by Martinez and colleagues [17], Ru was incorporated into Ni-based catalysts to facilitate ammonia decomposition, resulting in simultaneously improved activity and stability. Meanwhile, Thomson and colleagues [18] investigated the effect of platinum (Pt) as an additive for Ni. Their findings suggest that the bimetallic PtNi catalyst can significantly alleviate the hydrogen poisoning effect, resulting in a much-improved performance. However, the use of precious metals as dopants would increase the cost and inevitably hinder the application of these catalysts. To overcome this problem, Xiao’s group [19] synthesized several bimetallic MNi/Al2O3 catalysts, using Co, Ni, and Fe as dopants to improve the catalytic activity and delved into the synergy between the dopants and Ni to disclose the mechanism for the enhanced performance. The results show that bimetallic CoNi/Al2O3 exhibits almost full ammonia decomposition at 700 °C with the stable trend in the temperature range of 650–750 °C. We can conclude from these results that the doping strategy can effectively enhance the performance of Ni. However, the catalytic performance is still unsatisfactory at temperatures <700 °C.
In addition to the doping strategy, the literature also suggests that the activity of Ni is contingent on the support, with the support properties exerting a significant influence on its ammonia decomposition performance [20,21]. Recent efforts have focused on depositing Ni on different supports [22,23]. Xiao’s group proposed the Ni/MgAl2O4 catalysts with different support morphologies and studied the effect of the support morphologies on performance [24]. The findings indicate that Ni/MgAl2O4-LDH exhibits superior ammonia decomposition performance due to its high surface area and well-dispersed Ni. Additionally, the porous alumina matrix’s structural features were observed by Wang et al. [25] to promote the production of finely dispersed and small-sized Ni particles, thereby resulting in improved catalytic performance. These studies imply that tuning the support properties can effectively facilitate the promotion of Ni’s catalytic activity. However, the synthesis of these supports requires sophisticated methods, unsuitable for the scale-up of catalysts’ synthesis. In addition, the origin of the enhanced performance by support modification is still unclear.
In this study, we have synthesized several Ni/GdxCe1-xO2-δ as catalysts for ammonia decomposition by tuning the properties of GdxCe1-xO2-δ support. The results show both high ammonia conversion and stability for Ni/Gd0.2Ce0.8O2-δ at 600 °C. The improved performance was studied through a combination of DFT calculation and different physicochemical characterizations, revealing the fundamental mechanism behind it.

2. Results

2.1. Characterizations

The X-ray diffraction (XRD) patterns in Figure 1 show that all catalysts exhibit a similar cubic fluorite structure of GDC (PDF: 00-046-0508) and Ni (PDF: 00-004-0850). The absence of gadolinium planes is attributed to its proper doping within the ceria lattice. Thus, it can be deduced that the GdxCe1-xO2-δ supports and the active compound are in a crystalline phase. Importantly, based on the XRD patterns, there is no evidence of new compound formation or chemical reaction occurring between the support and the active compound. Scherer’s formula (Equation (1)) was used to evaluate crystallite sizes of the supports based on the (111) peak of GdxCe1-xO2-δ:
D = 0.89 λ β cos θ
where λ, β, θ, and D represent the X-ray wavelength, the angle between the sample and the X-ray beam, the full width at half maximum position, and the crystallite size, respectively. The derived crystallite size and lattice constant values of the support are summarized in Table 1. Specifically, we have found that the lattice constants of Gd0.1Ce0.9O2-δ, Gd0.2Ce0.8O2-δ, Gd0.3Ce0.7O2-δ, Gd0.4Ce0.6O2-δ are 5.15 Å, 5.17 Å, 5.20 Å, and 5.23 Å. We can attribute this discrepancy to the larger ionic radii of Gd3+ (1.19Å) compared with Ce3+ (1.01Å), as previously reported in the literature [26].
The EPR profiles displayed in Figure 2 reveal the oxygen vacancy concentration in the bulk of the reduced catalysts. The EPR signal centered at approximately 2.003 is present in all catalysts, indicating the existence of oxygen vacancies. The catalysts’ intensity follows this ranking: Gd0.2Ce0.8O2-δ > Gd0.3Ce0.7O2-δ > Gd0.1Ce0.9O2-δ > Gd0.4Ce0.6O2-δ. This indicates that Gd0.2Ce0.8O2-δ shows the most oxygen vacancies among all the samples. Therefore, we can conclude that adjusting the molar ratio of Gd and Ce can increase the oxygen vacancy concentration, with Gd0.2Ce0.8O2-δ showing the highest concentration.
Figure S1 illustrates the SEM images of the Ni/GdxCe1-xO2-δ with different Gd: Ce molar ratios. The results show the catalysts featured in the gathered spherical structures. In addition, the morphologies of the catalysts were found to be similar across all samples. This suggests that varying the molar ratio of Gd:Ce did not significantly impact the surface morphologies. Therefore, the difference in the catalytic activity was not influenced by the surface morphologies of the catalysts.
Additional textural properties in terms of surface area, actual Ni loading, Ni dispersion, and Ni particle size are shown in Table 2. It can be seen that all catalysts show a similar surface area at around 120 m2/g. The actual Ni loading of all catalysts was estimated to be near 9.8%, which is only slightly lower than the nominal Ni amount. The Ni particle size of the catalysts was also not significantly influenced by the Gd: Ce molar ratio. However, comparing the Ni dispersion of the catalysts on the surface, we can find that Ni/Gd0.2Ce0.8O2-δ displays the highest Ni dispersion of 20.7%. In addition, the amount of Ni dispersion is positively related to that of the oxygen vacancy concentration. This observation is consistent with recent findings by Liu et al. that the oxygen vacancy on the surface of catalysts can promote the dispersion of active sites on the surface [27].
H2-TPR profiles of the catalysts were used to evaluate the metal–support interaction between the active sites (Ni) with the GdxCe1-xO2-δ support. As is shown in Figure 3, all catalysts show a distinct peak in the temperature range of 250–500 °C. NiO reduction typically occurs within the temperature range of 200–500 °C. However, if there is a strong metal–support interaction, the reduction of NiO requires a temperature above 300 °C. In addition, the GdxCe1-xO2-δ support was usually reduced at 300–600 °C. It is difficult to differentiate the reduction peak of NiO with the significant metal–support interaction from that of GdxCe1-xO2-δ. However, the metal–support interaction can be evaluated by the number of reduction peaks at temperatures lower than 300 °C. After comparing the H2-profiles of the catalysts, it is evident that the reduction peaks of NiO/Gd0.2Ce0.8O2-δ are mostly located at temperatures above 300 °C, whereas the remaining catalysts exhibit partial reduction peaks at temperatures below 300 °C. In conclusion, NiO/Gd0.2Ce0.8O2-δ displays the most significant metal–support interaction compared with the other catalysts. Previous studies [28,29] demonstrated that increased oxygen vacancy concentration facilitates enhanced interaction between metal and support. The high metal–support interaction observed in NiO/Gd0.2Ce0.8O2-δ results from the increased oxygen vacancy concentration, which is achieved through the manipulation of support properties.

2.2. Ammonia Decomposition Performance

The ammonia decomposition performance of Ni/GdxCe1-xO2-δ catalysts was evaluated in the temperature range of 300–700 °C with an interval of 50 °C. As we can see in Figure 4, increasing the decomposition temperature can significantly improve the ammonia conversion for all catalysts. This is because ammonia decomposition is an exothermic reaction, and increasing the reaction temperature enhances its reaction thermodynamics. From a kinetic point of view, an increase in reaction temperature also promotes the rate of ammonia decomposition on the catalyst surface, thus increasing its decomposition efficiency. Comparing the performance of Ni/GdxCe1-xO2-δ with different Gd: Ce ratios, we can observe that the support properties of the catalysts have a significant influence on their catalytic performance for ammonia decomposition. For Ni/Gd0.4Ce0.6O2-δ, the catalyst only started to react at a temperature of 400 °C. Its ammonia decomposition efficiency was only 80.8% at 600 °C. The catalyst’s ammonia decomposition efficiency only approached 100% when the temperature reached as high as 700 °C. For Ni/Gd0.2Ce0.8O2-δ, the decomposition reaction began at temperatures higher than 300 °C, and the reaction efficiency gradually increased to 29.9% at 450 °C, followed by a sharp increase to 100% at 600 °C and GHSV of 30,000 L.kg−1.h−1 (Figure 4a). At 600 °C, Ni/Gd0.2Ce0.8O2-δ demonstrated a hydrogen production rate of 2008.9 mmol.g−1.h−1, outperforming the other catalysts evaluated. The ammonia conversion ranking of the catalysts, from highest to lowest: Ni/Gd0.2Ce0.8O2-δ, Ni/Gd0.3Ce0.7O2-δ, Ni/Gd0.1Ce0.9O2-δ, and Ni/Gd0.4Ce0.6O2-δ. It is worth pointing out that the performance correlates with the oxygen vacancy concentration and Ni dispersion on the surface.
Arrhenius equation was used to estimate the activation energies:
ln r NH 3 = E a R T + n
r NH3 = r NH3,catalyst r NH3,blank
where rNH3, Ea, R, and T represent the ammonia decomposition rate, activation energy, molar gas constant, and thermodynamic temperature; r NH3,catalyst and r NH3,blank denote the NH3 decomposition rate with and without catalysts. The apparent activation energies (Ea) for the catalysts were evaluated as 81.65 kJ/mol, 58.45 kJ/mol, 65.52 kJ/mol, and 83.02 kJ/mol for Ni/Gd0.1Ce0.9O2-δ, Ni/Gd0.2Ce0.8O2-δ, Ni/Gd0.3Ce0.7O2-δ, and Ni/Gd0.4Ce0.6O2-δ, respectively (Figure 5). We can conclude from the results that Gd0.2Ce0.8O2-δ support can significantly reduce the apparent energy barrier for ammonia decomposition, resulting in high catalytic performance.
The stability was investigated by conducting the ammonia decomposition reaction over the course of 150 h (Figure 6a). Ni/Gd0.4Ce0.6O2-δ showed a gradual decrease in ammonia conversion from 80.79% to 59.8%, while Ni/Gd0.2Ce0.8O2-δ maintained almost constant conversion, indicating its good stability. It can also be found from the results in Figure 6b that the performance of Ni/Gd0.2Ce0.8O2-δ is reproducible at variable temperatures within the range of 550–600 °C. As we all know, the high stability of the catalysts is derived from the inhibition of aggregation by strong metal–support interaction. These results suggest that adjusting the support properties can boost the concentration of oxygen vacancies, improving the interaction between the metal and support, ultimately resulting in high stability for Ni/Gd0.2Ce0.8O2-δ.
To highlight the novelty of the catalysts in this study, we also compared the performance of Ni/GdXCe1-xO2-δ with those reported in the literature (Table 3). The results show that Ni/Gd0.2Ce0.8O2-δ exhibits complete ammonia deocmpositon at temperatures as low as 600 °C and high GHSV of 30,000 mL/(gcat.h), which is higher than the majority of the catalysts reported in the literature, demonstrating its high potential for practical ammonia decomposition applications.

2.3. Mechanical Investigation

2.3.1. DFT Calculation

A DFT calculation was conducted to investigate the reaction route of ammonia decomposition over Ni (110), calculating the potential energy curves for each, thus uncovering the rate-limiting step during this process. The ammonia decomposition reaction was found to consist of two steps, as shown in Figure 7: stepwise dehydrogenation and atom-coupled desorption. A comparison of the energy barriers for these steps revealed that the energy barrier for the associative desorption of N2 was estimated to be 1.74 eV, which is much higher than that of the stepwise dehydrogenation processes. Therefore, the rate-limiting step of the ammonia decomposition reaction is the associative desorption of N.

2.3.2. Surface Properties of the Catalysts

Previous studies [32,33] have demonstrated that strong basicity is an important indicator for the enhanced ammonia decomposition activity. Wei et al. [34] further explain the enhanced activity of ammonia decomposition catalysts at strong basic sites. According to Wei et al., the strong basic sites can donate their electrons to the surface of the active sites, benefiting the recombinative desorption of N atoms on the surface. This claim was also supported by many published papers [35,36]. The above analysis has demonstrated that the associative desorption of N is the rate-limiting step. Therefore, evaluating the influence of GdxCe1-xO2-δ support on the abundance of the strong basic sites can be used to explain the enhanced performance of the Ni/Gd0.2Ce0.8O2-δ at medium temperatures.
Figure 8 shows the CO2-TPD profiles of the Ni/ GdxCe1-xO2-δ profiles. The spectrum can be classified into two types according to the position of the CO2 desorption peaks. The peaks at temperatures <300 °C can be linked to the weak basic sites (Bronstone sites) and the peaks at high temperature (>450 °C) can be ascribed to the contribution of the strong basic sites (Lewis sites). Comparing the CO2-TPD profiles of the catalysts, we can find that all catalysts exhibit two CO2 desorption peaks at low and high temperatures. In addition, Ni/Gd0.2Ce0.8O2-δ had the highest area of the peak at a high temperature, followed by Ni/Gd0.3Ce0.7O2-δ, Ni/Gd0.1Ce0.9O2-δ, and Ni/Gd0.4Ce0.6O2-δ. This indicates that Gd0.2Ce0.8O2-δ support can significantly improve the abundance of the strong basic sites on the surface, benefiting the associative desorption of N on the surface, thus leading to enhanced catalytic performance toward the ammonia decomposition.

2.3.3. NH3 TPSR

NH3-TPSR can yield crucial insights into catalyst surface reactions. Our DFT calculations have demonstrated that N recombinative desorption is the rate-limiting step, and the energy barriers can be estimated from N2 formation profiles obtained from NH3-TPSR (Figure 9). All four catalysts exhibit two N2 formation peaks in the temperature range of 300–400 °C and 450–600 °C. A comparison of N2 formation peaks reveals that Ni/Gd0.2Ce0.8O2-δ and Ni/Gd0.3Ce0.7O2-δ show major peaks at lower temperatures, while Ni/Gd0.1Ce0.9O2-δ and Ni/Gd0.4Ce0.6O2-δ have their majority N2 formation peaks at higher temperatures. Furthermore, the catalysts can be ranked in order of N2 formation temperature as follows: Ni/Gd0.2Ce0.8O2-δ > Ni/Gd0.3Ce0.7O2-δ > Ni/Gd0.1Ce0.9O2-δ > Ni/Gd0.4Ce0.6O2-δ. This ranking indicates that associative desorption of N is easier on Ni/ Ni/Gd0.2Ce0.8O2-δ, which is consistent with its abundant strong basic sites, benefiting the electron conduction to the active sites. Therefore, we can conclude from the results that the Gd0.2Ce0.8O2-δ support can improve the abundance of the strong basic sites, benefiting the associative desorption of N and, thus, leading to enhanced catalytic performance resulting in ammonia decomposition. The H2 and N2 signal from the NH3-TPSR is shown in Figure S2.

3. Materials and Methods

3.1. Preparation of the Catalysts

Ni/GdxCe1-xO2-δ were synthesized using a sol-gel method. The support properties were modified by the adjustment of the relative molar concentrations of Gd and Ce. For the synthesis of the Ni/GdxCe1-xO2-δ, Ni(NO3)3·6H2O, Ce(NO3)3·6H2O, Gd(NO3)3·6H2O, and citric acid were dissolved in deionized water to form a solution with a metal ion/citric acid molar ratio of 1:1.1. After that, the solution was agitated and heated at 95 °C. The PEG-400 was employed as a gel promoter. The solution was heated at 95 °C overnight to remove all the water in the solution, thus forming the gel. After the evaporation of the water in the solution, the gel was dehydrated and annealed in air at 850 °C for 6 h to derive the final catalysts. The nominal mass ratio of metallic Ni was set as 10 wt%, while the nominal molar ratio of Ce:Gd varied between 9:1 and 6:4.

3.2. Characterizations

XRD patterns were obtained using the PANalytical X’Pert-Pro powder X-ray diffractometer with CuKα radiation (0.1541 nm), with a fixed scanning step size of 0.0334° in the 10−80° range, and operated at 40 kV and 40 mA. All catalysts were pretreated in 250 mL/min 10% H2/N2 for 1 h at 700 °C to reduce NiO before the characterization. The Scanning Electron Microscopy coupled with Energy Dispersive X-ray Spectroscopy (SEM-EDS), Quanta 200FEG with an Oxford EDS system was used for the morphological analysis of the catalysts. An ASAP 2010 instrument (Micromeritics, Norcross, GA, USA) was used to conduct the N2 physisorption characterization, which was used to determine the specific surface area of the catalysts. To quantify the actual content, an inductively coupled plasma optical emission spectrometry (ICP-OES) was carried out on an Optima 5300 DV instrument. The metal dispersion of Ni was assessed on an ASAP 2010C chemisorption analyzer (Micromeritics) using the static volumetric H2 chemisorption method. It was assumed that each nickel atom would adsorb a single hydrogen atom.
TPR Win v 1.50 (Quantanchrome Instruments Co., Boynton Beach, FL, USA) was used for H2-TPR analysis using 10% H2/Ar mixture (25 mL·min−1) as the reducing agent. The procedure for conducting the H2-TPR reaction was as follows: First, 30 mg of catalysts was placed into the equipment. After that, the catalysts were pretreated at 140 °C under 30 mL/min He for 30 min to remove the possible moisture on the surface. Then, the samples were cooled down to room temperature to initiate the reduction reaction. The temperature was ramped at a rate of 10 °C·min−1 from room temperature to 800 °C.
CO2-temperature-programmed desorption (TPD) measurement was performed using the same instrument as for the H2-TPR. The characterization was conducted to evaluate the distribution of the basic sites on the surface of catalysts. Typically, the catalysts were pre-reduced in 10% H2/Ar at 600 °C for 4 h to remove all reducible oxygen, followed by cooling the catalysts to room temperature. After that, the catalysts were exposed to CO2 atmosphere at 60 °C for 1 h to adsorb CO2. After the termination of the adsorption reaction, the reactor was heated to 100 °C and flushed with He for 1 h to remove the physisorbed CO2. Finally, the reactor temperature was heated from 100 °C to 800 °C with a ramp of 10 °C/min.
The NH3-temperature-programmed surface reaction (TPSR) was measured using a homemade fixed-bed reactor. To begin the experiment, the reactor was first evacuated and heated to the desired temperature. Ammonia was then introduced into the reactor at a constant flow rate, and the temperature was gradually increased at 10 K/min. As the temperature increased, the reaction between the ammonia and the catalyst surface occurred, resulting in the production of various products. The effluent gas was continuously monitored by a Pfeiffer Vacuum mass spectrometer.

3.3. Computational Details

The Vienna Ab-initio Simulation Package (VASP) was used to conduct density functional theory (DFT) calculations using a five slab-layer Ni(110) and a p(3 × 3) unit cell. The calculations incorporated the Perdew–Burke–Ernzerhof (PBE) functional and the generalized gradient approximation (GGA) to calculate the exchange-correlation energy. The valence electron properties were modeled through the projector augmented-wave (PAW) approach, with an energy cutoff of 400 eV and a Monkhorst-Pack grid of (3 × 3 × 1) used for k-point sampling. To prevent interaction between adjacent slabs, a vacuum spacing of 15 Å in the z-direction was applied. The transition states of ammonia decomposition were determined using the dimmer method, with geometry optimization carried out using the force-based conjugated gradient method. Finally, the adsorption energy (Eads) was calculated as:
E ads = E adsorbate + surface E adsorbate E surface
where Eadsorbate and Esurface denote the total energies of the gaseous species and the clean slab.

3.4. Experiment Setup

The ammonia decomposition reaction was conducted in a homemade, lab-scale fixed bed reactor. A quartz tube (300 mm in length, 8 mm in diameter) and a porous distributor were used to support the catalysts. An electric furnace was used to heat the reactor. The temperature of the reactor and the catalyst was controlled and monitored by two thermocouples to achieve the desired operating temperature. The details of the experimental setup used in this study are depicted in Figure S3. Prior to the ammonia decomposition reaction, calcined catalysts (200 mg) were subjected to pretreatment in 10% H2/N2. The reaction performance was evaluated using pure NH3 at 300–650 °C, at the gas hourly space velocity (GHSV) of 30,000 mL/(gcat.h). The residual NH3 in the effluent was absorbed in a 6 mol/L sulfuric acid solution. A mass flowmeter was used to monitor the flow rate of the outlet gas.
The NH3 conversion was calculated as follows:
NH 3 ,   conversion = F NH 3 , converted F NH 3 , inlet
Since ammonia is converted to N2 and H2 in the outlet gas, and a unit volume of ammonia is converted into two volumes of nitrogen and hydrogen mixture, the volume of ammonia converted is half the flow rate of the gas after the unconverted ammonia has been removed from the reactor outlet. That is, the converted ammonia flow rate was calculated as:
F NH 3 , converted = F out 2
Therefore, NH3 conversion can be calculated as follows:
NH 3 ,   conversion = F NH 3 , converted F NH 3 , inlet = F out 2 F NH 3 , inlet

4. Conclusions

Tuning the support properties can enhance the activity and stability of Ni/ GdxCe1-xO2-δ for ammonia decomposition to obtain high-purity hydrogen. Ni/Gd0.2Ce0.8O2-δ shows almost full ammonia decomposition, leading to a high hydrogen production rate (2008.9 mmol.g−1.h−1) at 600 °C and 30,000 mL/(gcat.h). The performance persisted over 150 h long run and at variable temperatures of 550–600 °C. The mechanism investigations have manifested that Gd0.2Ce0.8O2-δ exhibits high oxygen vacancy concentration, highly dispersed Ni, and abundant strong active sites. These properties of Ni/Gd0.2Ce0.8O2-δ significantly improve the associative desorption of N and strengthen the metal–support interaction. Therefore, Ni/Gd0.2Ce0.8O2-δ shows satisfactory activity and stability at 600 °C.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28062750/s1, Figure S1: SEM images of the as-prepared NiO/CexGd1-xO2-δ catalysts. (a) Ni/Gd0.1Ce0.9O2-δ, (b) Ni/Gd0.2Ce0.8O2-δ, (c) Ni/Gd0.3Ce0.7O2-δ, and (d) Ni/Gd0.4Ce0.6O2-δ, Figure S2: (a) NH3 and (b) H2 profiles derived from NH3-TPSR measurement over Ni/GdxCe1-xO2-δ catalysts, Figure S3: Schematic illustration of the ammonia decomposition setup.

Author Contributions

Investigation and conceptualization, H.H.; writing—original draft and software, C.C.; visualization and funding acquisition, C.B.; formal analysis, J.R.; writing—review and editing, J.L.; conceptualization and supervision, W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Jinhua Science and Technology Bureau Project, grant number 2021-4-334; National Nature Science Foundation of China, grant number 21902065.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nishiyama, H.; Yamada, T.; Nakabayashi, M.; Maehara, Y.; Yamaguchi, M.; Kuromiya, Y.; Nagatsuma, Y.; Tokudome, H.; Akiyama, S.; Watanabe, T.; et al. Photocatalytic solar hydrogen production from water on a 100-m2 scale. Nature 2021, 598, 304–307. [Google Scholar] [CrossRef]
  2. Scheffe, J.R.; Allendorf, M.D.; Coker, E.N.; Jacobs, B.W.; McDaniel, A.H.; Weimer, A.W. Hydrogen production via chemical looping redox cycles using atomic layer deposition-synthesized iron oxide and cobalt ferrites. Chem. Mater. 2011, 23, 2030–2038. [Google Scholar] [CrossRef]
  3. Fang, X.; Wang, X.; Ouyang, L.; Zhang, L.; Sun, S.; Liang, Y.; Luo, Y.; Zheng, D.; Kang, T.; Liu, Q.; et al. Amorphous Co-Mo-B Film: A High-Active Electrocatalyst for Hydrogen Generation in Alkaline Seawater. Molecules 2022, 27, 7617. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, J.; Chen, Y.; Yuan, L.; Zhang, M.; Zhang, C. Scandium Decoration of Boron Doped Porous Graphene for High-Capacity Hydrogen Storage. Molecules 2019, 24, 2382. [Google Scholar] [CrossRef] [Green Version]
  5. Shao, Z.; Li, Y.; Liu, C.; Ai, W.; Luo, S.-P.; Liu, Q. Reversible interconversion between methanol-diamine and diamide for hydrogen storage based on manganese catalyzed (de)hydrogenation. Nat. Commun. 2020, 11, 591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Klerke, A.; Christensen, C.H.; Nørskov, J.K.; Vegge, T. Ammonia for hydrogen storage: Challenges and opportunities. J. Mater. Chem. 2008, 18, 2304–2310. [Google Scholar] [CrossRef]
  7. Puszkiel, J.; Gasnier, A.; Amica, G.; Gennari, F. Tuning LiBH4 for Hydrogen Storage: Destabilization, Additive, and Nanoconfinement Approaches. Molecules 2020, 25, 163. [Google Scholar] [CrossRef] [Green Version]
  8. Li, D.-H.; Li, Q.-M.; Qi, S.-L.; Qin, H.-C.; Liang, X.-Q.; Li, L. Theoretical Study of Hydrogen Production from Ammonia Borane Catalyzed by Metal and Non-Metal Diatom-Doped Cobalt Phosphide. Molecules 2022, 27, 8206. [Google Scholar] [CrossRef]
  9. Akbayrak, S.; Özkar, S. Ammonia borane as hydrogen storage materials. Int. J. Hydrogen Energy 2018, 43, 18592–18606. [Google Scholar] [CrossRef]
  10. Lamb, K.E.; Dolan, M.D.; Kennedy, D.F. Ammonia for hydrogen storage; A review of catalytic ammonia decomposition and hydrogen separation and purification. Int. J. Hydrogen Energy 2019, 44, 3580–3593. [Google Scholar] [CrossRef]
  11. Lee, H.J.; Park, E.D. Ammonia Decomposition over Ru/SiO2 Catalysts. Catalysts 2022, 12, 1203. [Google Scholar] [CrossRef]
  12. Hu, Z.-P.; Weng, C.-C.; Yuan, G.-G.; Lv, X.-W.; Yuan, Z.-Y. Ni nanoparticles supported on mica for efficient decomposition of ammonia to COx-free hydrogen. Int. J. Hydrogen Energy 2018, 43, 9663–9676. [Google Scholar] [CrossRef]
  13. Lu, B.; Li, L.; Ren, M.; Liu, Y.; Zhang, Y.; Xu, X.; Wang, X.; Qiu, H. Ammonia decomposition over iron-based catalyst: Exploring the hidden active phase. Appl. Catal. B 2022, 314, 121475. [Google Scholar] [CrossRef]
  14. Pinzón, M.; Sánchez-Sánchez, A.; Romero, A.; de la Osa, A.R.; Sánchez, P. Self-combustion Ni and Co-based perovskites as catalyst precursors for ammonia decomposition. Effect of Ce and Mg doping. Fuel 2022, 323, 124384. [Google Scholar] [CrossRef]
  15. Okura, K.; Miyazaki, K.; Muroyama, H.; Matsui, T.; Eguchi, K. Ammonia decomposition over Ni catalysts supported on perovskite-type oxides for the on-site generation of hydrogen. RSC Adv. 2018, 8, 32102–32110. [Google Scholar] [CrossRef] [Green Version]
  16. Duan, X.; Qian, G.; Fan, C.; Zhu, Y.; Zhou, X.; Chen, D.; Yuan, W. First-principles calculations of ammonia decomposition on Ni (110) surface. Surf. Sci. 2012, 606, 549–553. [Google Scholar] [CrossRef]
  17. Lucentini, I.; García Colli, G.; Luzi, C.; Serrano, I.; Soler, L.; Divins, N.J.; Martínez, O.M.; Llorca, J. Modelling and simulation of catalytic ammonia decomposition over Ni-Ru deposited on 3D-printed CeO2. Chem. Eng. J. 2022, 427, 131756. [Google Scholar] [CrossRef]
  18. Chellappa, A.S.; Fischer, C.M.; Thomson, W.J. Ammonia decomposition kinetics over Ni-Pt/Al2O3 for PEM fuel cell applications. Appl. Catal. A 2002, 227, 231–240. [Google Scholar] [CrossRef]
  19. Fu, E.; Qiu, Y.; Lu, H.; Wang, S.; Liu, L.; Feng, H.; Yang, Y.; Wu, Z.; Xie, Y.; Gong, F.; et al. Enhanced NH3 decomposition for H2 production over bimetallic M(M=Co, Fe, Cu)Ni/Al2O3. Fuel Process. Technol. 2021, 221, 106945. [Google Scholar] [CrossRef]
  20. 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]
  21. Deng, Q.-F.; Zhang, H.; Hou, X.-X.; Ren, T.-Z.; Yuan, Z.-Y. High-surface-area Ce0.8Zr0.2O2 solid solutions supported Ni catalysts for ammonia decomposition to hydrogen. Int. J. Hydrogen Energy 2012, 37, 15901–15907. [Google Scholar] [CrossRef]
  22. Torrente-Murciano, L.; Hill, A.K.; Bell, T.E. Ammonia decomposition over cobalt/carbon catalysts—Effect of carbon support and electron donating promoter on activity. Catal. Today 2017, 286, 131–140. [Google Scholar] [CrossRef] [Green Version]
  23. Li, X.-K.; Ji, W.-J.; Zhao, J.; Wang, S.-J.; Au, C.-T. 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]
  24. Qiu, Y.; Fu, E.; Gong, F.; Xiao, R. Catalyst support effect on ammonia decomposition over Ni/MgAl2O4 towards hydrogen production. Int. J. Hydrogen Energy 2022, 47, 5044–5052. [Google Scholar] [CrossRef]
  25. 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]
  26. Zeng, D.; Kang, F.; Qiu, Y.; Cui, D.; Li, M.; Ma, L.; Zhang, S.; Xiao, R. Iron oxides with gadolinium-doped cerium oxides as active supports for chemical looping hydrogen production. Chem. Eng. J. 2020, 396, 125153. [Google Scholar] [CrossRef]
  27. Tang, H.; Wang, Y.; Zhang, W.; Liu, Z.; Li, L.; Han, W.; Li, Y. Catalytic activity of Ru supported on SmCeOx for ammonia decomposition: The effect of Sm doping. J. Solid State Chem. 2021, 295, 121946. [Google Scholar] [CrossRef]
  28. Lin, L.-Y.; Kavadiya, S.; He, X.; Wang, W.-N.; Karakocak, B.B.; Lin, Y.-C.; Berezin, M.Y.; Biswas, P. Engineering stable Pt nanoparticles and oxygen vacancies on defective TiO2 via introducing strong electronic metal–support interaction for efficient CO2 photoreduction. Chem. Eng. J. 2020, 389, 123450. [Google Scholar] [CrossRef]
  29. Xiao, Z.; Li, Y.; Hou, F.; Wu, C.; Pan, L.; Zou, J.; Wang, L.; Zhang, X.; Liu, G.; Li, G. Engineering oxygen vacancies and nickel dispersion on CeO2 by Pr doping for highly stable ethanol steam reforming. Appl. Catal. B 2019, 258, 117940. [Google Scholar] [CrossRef]
  30. Muroyama, H.; Saburi, C.; Matsui, T.; Eguchi, K. Ammonia decomposition over Ni/La2O3 catalyst for on-site generation of hydrogen. Appl. Catal. A 2012, 443–444, 119–124. [Google Scholar] [CrossRef]
  31. He, H.; Jiang, H.; Yang, F.; Liu, J.; Zhang, W.; Jin, M.; Li, Z. Bimetallic NixCo10-x/CeO2 as highly active catalysts to enhance mid-temperature ammonia decomposition: Kinetics and synergies. Int. J. Hydrogen Energy 2023, 48, 5030–5041. [Google Scholar] [CrossRef]
  32. 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] [Green Version]
  33. Podila, S.; Driss, H.; Zaman, S.F.; Ali, A.M.; Al-Zahrani, A.A.; Daous, M.A.; Petrov, L.A. Effect of preparation methods on the catalyst performance of Co/MgLa mixed oxide catalyst for COx-free hydrogen production by ammonia decomposition. Int. J. Hydrogen Energy 2017, 42, 24213–24221. [Google Scholar] [CrossRef]
  34. 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]
  35. Bajus, S.; Agel, F.; Kusche, M.; Ní Bhriain, N.; Wasserscheid, P. Alkali hydroxide-modified Ru/γ-Al2O3 catalysts for ammonia decomposition. Appl. Catal. A 2016, 510, 189–195. [Google Scholar] [CrossRef]
  36. Wang, B.; Kong, H.; Wang, H.; Wang, Y.; Hu, X. Kinetic and thermodynamic analyses of mid/low-temperature ammonia decomposition in solar-driven hydrogen permeation membrane reactor. Int. J. Hydrogen Energy 2019, 44, 26874–26887. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the reduced Ni/GdxCe1-xO2-δ catalysts. (a) Ni/Gd0.1Ce0.9O2-δ, (b) Ni/Gd0.2Ce0.8O2-δ, (c) Ni/Gd0.3Ce0.7O2-δ, and (d) Ni/Gd0.4Ce0.6O2-δ.
Figure 1. XRD patterns of the reduced Ni/GdxCe1-xO2-δ catalysts. (a) Ni/Gd0.1Ce0.9O2-δ, (b) Ni/Gd0.2Ce0.8O2-δ, (c) Ni/Gd0.3Ce0.7O2-δ, and (d) Ni/Gd0.4Ce0.6O2-δ.
Molecules 28 02750 g001
Figure 2. EPR profiles of the reduced Ni/GdxCe1-xO2-δ support.
Figure 2. EPR profiles of the reduced Ni/GdxCe1-xO2-δ support.
Molecules 28 02750 g002
Figure 3. H2-TPR of the as-prepared NiO/GdxCe1-xO2-δ catalysts. (a) NiO/Gd0.1Ce0.9O2-δ, (b) NiO/Gd0.2Ce0.8O2-δ, (c) NiO/Gd0.3Ce0.7O2-δ, and (d) NiO/Gd0.4Ce0.6O2-δ.
Figure 3. H2-TPR of the as-prepared NiO/GdxCe1-xO2-δ catalysts. (a) NiO/Gd0.1Ce0.9O2-δ, (b) NiO/Gd0.2Ce0.8O2-δ, (c) NiO/Gd0.3Ce0.7O2-δ, and (d) NiO/Gd0.4Ce0.6O2-δ.
Molecules 28 02750 g003
Figure 4. Ammonia decomposition performance over Ni/GdxCe1-xO2-δ catalysts. (a) Temperature-resolved ammonia conversion at 300−700 °C, and (b) the hydrogen production rate at 600 °C.
Figure 4. Ammonia decomposition performance over Ni/GdxCe1-xO2-δ catalysts. (a) Temperature-resolved ammonia conversion at 300−700 °C, and (b) the hydrogen production rate at 600 °C.
Molecules 28 02750 g004
Figure 5. The Arrhenius plots of ammonia decomposition over Ni/GdxCe1-xO2-δ.
Figure 5. The Arrhenius plots of ammonia decomposition over Ni/GdxCe1-xO2-δ.
Molecules 28 02750 g005
Figure 6. Ammonia conversion as a function of time (a) over the course of 150 h and (b) at 550–600 °C.
Figure 6. Ammonia conversion as a function of time (a) over the course of 150 h and (b) at 550–600 °C.
Molecules 28 02750 g006
Figure 7. The potential energy profiles of ammonia decomposition over Ni (110) surface.
Figure 7. The potential energy profiles of ammonia decomposition over Ni (110) surface.
Molecules 28 02750 g007
Figure 8. CO2-TPD profiles of the Ni/GdxCe1-xO2-δ catalysts.
Figure 8. CO2-TPD profiles of the Ni/GdxCe1-xO2-δ catalysts.
Molecules 28 02750 g008
Figure 9. N2 formation profiles derived from NH3-TPSR measurement over Ni/GdxCe1-xO2-δ catalysts.
Figure 9. N2 formation profiles derived from NH3-TPSR measurement over Ni/GdxCe1-xO2-δ catalysts.
Molecules 28 02750 g009
Table 1. The crystallite size and lattice constant of the Ni/GdxCe1-xO2-δ catalysts.
Table 1. The crystallite size and lattice constant of the Ni/GdxCe1-xO2-δ catalysts.
Crystallite Size (nm)Lattice Constant (Å)
Ni/Gd0.1Ce0.9O2-δ585.15
Ni/Gd0.2Ce0.8O2-δ605.17
Ni/Gd0.3Ce0.7O2-δ575.20
Ni/Gd0.4Ce0.6O2-δ595.23
Table 2. The textural properties of the catalysts.
Table 2. The textural properties of the catalysts.
CatalystsSurface Area aActual Ni Loading bNi Dispersion cNi Particle Size c
m2/gwt%%nm
Ni/Gd0.1Ce0.9O2-δ121.49.8515.88.7
Ni/Gd0.2Ce0.8O2-δ120.69.9320.78.3
Ni/Gd0.3Ce0.7O2-δ124.39.7917.18.1
Ni/Gd0.4Ce0.6O2-δ119.59.8514.88.9
a derived from BET measurement, b derived from ICP-OES, c derived from H2 chemisorption.
Table 3. Comparison of the catalytic activity of Ni/GdxCe1-xO2-δ with those in the literature.
Table 3. Comparison of the catalytic activity of Ni/GdxCe1-xO2-δ with those in the literature.
CatalystsTemperatureGHSVNH3 ConversionHydrogen Production RateReferences
°Cml/(gcat.h)%mmol/(gcat.h)
Ni/Al2O370030,00079.21591.1[19]
FeNi/Al2O370030,00015.7315.4[19]
CoNi/Al2O370030,0001002008.9[19]
CuNi/Al2O370030,00034.7697.1[19]
Ni/ZrO2650600034.8139.8[30]
Ni/TiO2650600083.1333.9[30]
Ni/CeO2650600089.7360.4[30]
Ni/La2O36506000100401.8[30]
Ni10/CeO265030,00031.0622.8[31]
Ni7.5Co2.5/CeO265030,00031.0622.8[31]
Ni2.5Co7.5/CeO265030,00031.0622.8[31]
Co10/CeO265030,00031.0622.8[31]
Ni/Gd0.1Ce0.9O2-δ60030,00084.61699.6This work
Ni/Gd0.2Ce0.8O2-δ60030,0001002008.9This work
Ni/Gd0.3Ce0.7O2-δ60030,00092.31854.2This work
Ni/Gd0.4Ce0.6O2-δ60030,00080.81623.2This work
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

He, H.; Chen, C.; Bian, C.; Ren, J.; Liu, J.; Huang, W. Enhanced Ammonia Decomposition by Tuning the Support Properties of Ni/GdxCe1-xO2-δ at 600 °C. Molecules 2023, 28, 2750. https://doi.org/10.3390/molecules28062750

AMA Style

He H, Chen C, Bian C, Ren J, Liu J, Huang W. Enhanced Ammonia Decomposition by Tuning the Support Properties of Ni/GdxCe1-xO2-δ at 600 °C. Molecules. 2023; 28(6):2750. https://doi.org/10.3390/molecules28062750

Chicago/Turabian Style

He, Haihua, Chonglai Chen, Chaoqun Bian, Junhua Ren, Jiajia Liu, and Wei Huang. 2023. "Enhanced Ammonia Decomposition by Tuning the Support Properties of Ni/GdxCe1-xO2-δ at 600 °C" Molecules 28, no. 6: 2750. https://doi.org/10.3390/molecules28062750

Article Metrics

Back to TopTop