NC State
BioResources
Wang, R., Chen, X., Hao, H., Wang, B., Yu, H., Wang, M., Xie, Y., Wang, J., and Si, H. (2024). “Enhanced activity of Ru-based catalysts for ammonia decomposition through nitrogen doping of hierarchical porous carbon carriers,” BioResources 19(3), 4313-4334.

Abstract

Activated carbon (AC) materials, renowned for their high specific surface area, excellent conductivity, and customizable functional groups, are widely employed as catalyst carriers. However, enhancing the activity of Ru-based catalysts supported on AC (Ru/AC) for ammonia decomposition remains a challenge. In this study, commercial AC was utilized as a substrate, with glucose and urea employed as modifiers. Specifically, the surface of the AC was modified via a hydrothermal pyrolysis method, resulting in the successful post-treatment in situ co-doping of nitrogen (AC-GN). Experimental results revealed that Ru/AC-GN exhibited a hydrogen production rate 46% higher than that of Ru/AC at 475 °C, indicating improved activity and stability. The characterization of AC-GN demonstrated that nitrogen doping primarily occurred on the external surface and macropores of the AC, increasing the nitrogen content in the carrier, particularly pyrrolic nitrogen content, while preserving the original structural and morphological integrity of the AC. The enhanced dispersion of Ru, combined with the improved electronic transmission capabilities and strengthened interactions between the metal and the modified carrier, were identified as pivotal factors contributing to the enhanced low-temperature efficacy of Ru/AC-GN. This paper presents a novel direction for the large-scale preparation of efficient catalysts for ammonia decomposition.


Download PDF

Full Article

Enhanced Activity of Ru-based Catalysts for Ammonia Decomposition through Nitrogen Doping of Hierarchical Porous Carbon Carriers

Rui Wang,a Xiuxiu Chen,a Han Hao,b Bing Wang,a Hewei Yu,a,* Meng Wang,c Yongjun Xie,d Jianmei Wang,a,* and Hongyu Si a,*

Activated carbon (AC) materials, renowned for their high specific surface area, excellent conductivity, and customizable functional groups, are widely employed as catalyst carriers. However, enhancing the activity of Ru-based catalysts supported on AC (Ru/AC) for ammonia decomposition remains a challenge. In this study, commercial AC was utilized as a substrate, with glucose and urea employed as modifiers. Specifically, the surface of the AC was modified via a hydrothermal pyrolysis method, resulting in the successful post-treatment in situ co-doping of nitrogen (AC-GN). Experimental results revealed that Ru/AC-GN exhibited a hydrogen production rate 46% higher than that of Ru/AC at 475 °C, indicating improved activity and stability. The characterization of AC-GN demonstrated that nitrogen doping primarily occurred on the external surface and macropores of the AC, increasing the nitrogen content in the carrier, particularly pyrrolic nitrogen content, while preserving the original structural and morphological integrity of the AC. The enhanced dispersion of Ru, combined with the improved electronic transmission capabilities and strengthened interactions between the metal and the modified carrier, were identified as pivotal factors contributing to the enhanced low-temperature efficacy of Ru/AC-GN. This paper presents a novel direction for the large-scale preparation of efficient catalysts for ammonia decomposition.

DOI: 10.15376/biores.19.3.4313-4334

Keywords: Activated carbon; Nitrogen doping; Glucose; Ammonia decomposition

Contact information: a: Shandong Provincial Key Laboratory of Biomass Gasification Technology, Energy Research Institute, Qilu University of Technology (Shandong Academy of Sciences), Jinan, 250014, China; b: Jinan Xinhang Experimental Foreign Language School, Jinan, 250014, China; c: The Li Xia Squadron of the Ecological Environment Protection Comprehensive Administrative Law Enforcement Detachment in Jinan City, Shandong Province, China; d: 101 Shanjian Road, Economic Development Zone Linyi, Shandong, China; *Corresponding authors: yhw@qlu.edu.cn (H. YU), wangjm@sderi.cn (J. Wang), sihy@sderi.cn (H. Si).

GRAPHICAL ABSTRACT

INTRODUCTION

Hydrogen is widely recognized as an efficient and clean energy source, offering potential solutions to environmental pollution and the energy crisis (Mehrpooya and Habibi 2020; Wan et al. 2021). However, its low volumetric density and challenges associated with its storage and transportation, particularly liquefaction, pose significant barriers to its widespread application (Abe et al. 2019; Luo et al. 2020). In recent years, ammonia has garnered considerable attention as an effective hydrogen carrier owing to its high hydrogen content of 17.7 wt% alongside its liquid state at room temperature (25 °C) and low pressure (6 atm). Additionally, its COX-free decomposition process and the relative maturity of ammonia storage and transport technologies make it a highly promising medium for H2 storage and transportation (Chang et al. 2021; Sun et al. 2022).

The decomposition of ammonia for hydrogen production is considered a promising pathway for transitioning to clean energy and developing a hydrogen economy, despite existing technical challenges, particularly in enhancing decomposition efficiency and reducing costs (Lee et al. 2021). A key strategy involves the development of highly active catalysts that can operate effectively at temperatures below 500 °C (Mukherjee et al. 2018). Commonly used active metals in ammonia decomposition catalysts include Ru, Ni, and Co. Among them, Ru-based catalysts are the most frequently utilized, owing to their exceptional activity at low temperatures (Cao et al. 2022; Nakamura et al. 2022). Notably, catalytic activity is greatly influenced by the dispersibility, morphology, and size of Ru nanoparticles (Yao et al. 2011). The B5-type surface of Ru contains high-activity sites, and nanoparticles approximately 2 nm in size are considered to provide an increased number of B5-type sites (Zheng et al. 2007; Su et al. 2024). Karim et al. (2009) have demonstrated that the catalytic activity of spherical Ru nanoparticles ranging from 1.8 to 3 nm in size was the highest. Furthermore, extensive research has revealed a close relationship between the efficiency of NH3 decomposition and factors such as alkaline environments and electron transfer capabilities (Tan et al. 2012; Nagaoka et al. 2014), which can be modulated by employing different carriers.

Porous carbon-based materials, known for their expansive specific surface area, adaptable functional groups, and exceptional conductivity, are widely employed as catalyst carriers (Yin et al. 2004a). The primary focus of research has centered on activated carbon (AC) and carbon nanotubes (CNTs) (Chen et al. 2021). For instance, Yin et al. (2004b) synthesized K-Ru/CNTs at 450 °C with an NH3 flow rate of 60000 mLgcat-1 h-1, achieving a decomposition efficiency of 97.3% and hydrogen formation rate 32.6 mmol gcat-1min-1, which significantly outperformed the 9.6 mmol gcat-1min-1 hydrogen formation rate of Ru/AC. This enhancement is attributed to the high degree of graphitization in CNTs, facilitating electron transfer between the carrier and metal. There is a notable increase in the H2 production rate by doping nitrogen into CNT carriers, primarily owing to enhanced electron interactions between the carrier and the metal, as well as the increased anchoring sites provided by nitrogen, which improve Ru particle dispersion (Hien et al. 2015; Ren et al. 2017). However, research on modifying AC via nitrogen doping for ammonia decomposition remains relatively limited.

Biomass carbon materials are considered excellent carrier materials from both environmental and cost perspectives. However, current research indicates that heteroatoms such as S, Cl, and N on the surface of AC can inhibit ammonia decomposition owing to low graphitization levels, impeding effective electron transfer between the metal and the carrier (Yin et al. 2004b). While heteroatoms can be removed by introducing H2 at high temperatures (Raróg-Pilecka et al. 2003), the presence of defects benefits Ru anchoring, thereby enhancing dispersion (Rodríguez-reinoso 1998). Nevertheless, oxygen-containing functional groups on the surface can diminish the activity of Ru owing to their electron-withdrawing effect (Rarogpilecka et al. 2005). Consequently, a suitable nitrogen doping method must be chosen to strike an optimal balance. In this context, a pioneering approach uniformly anchors micrometer-sized carbon spheres onto the surface and within the large pores of commercial AC using glucose via a hydrothermal method, revealing a new direction for post-treatment nitrogen doping (Liu et al. 2010).

To enhance the performance of the AC carrier, a nitrogen-doping method based on hydrothermal pyrolysis was employed. Concurrently, nitrogen-doped hydrothermal carbon was deposited on the surface of the AC, achieving the post-treatment in situ co-doping of nitrogen. This method aimed to preserve the original pore structure and degree of graphitization to the greatest possible extent, while the introduction of nitrogen provides additional sites for Ru loading. Consequently, the performance of AC as an ammonia decomposition catalyst carrier is substantially enhanced, thereby offering a novel approach for preparing high-activity Ru/AC catalysts.

EXPERIMENTAL

Materials

AC, derived from coconut shells, was procured from Shanghai Ron Chemical Technology Co., Ltd., with the item number R019806. Urea (> 99%), glucose (C6H12O6 > 99%), nitric acid (65 to 69%), and ruthenium (31.3%) were sourced from commercial suppliers and employed without additional purification.

Fig. 1. Experimental process and detailed schematic diagram

Sample synthesis

The carrier was prepared by grinding a suitable amount of commercial AC to 20–40 mesh, referred to as AC henceforth in this paper. Subsequently, 5 g of AC was mixed with 200 mL of nitric acid for oxidative pretreatment, followed by multiple washes with deionized water and drying at 80 °C for 12 h, resulting in AC-O. Then, 1 g of AC-O and 1 g of urea were added to 20 mL of distilled water and stirred at room temperature for 30 min, followed by transferring the mixture to an autoclave. The autoclave was placed in an oven and maintained at 180 °C for 6 h. After cooling to room temperature, the sample was filtered and dried at 80 °C for 12 h (Figure 1a). The obtained sample was then placed in a tube furnace and calcined under a N2 flow of 100 mL min-1. The pyrolysis protocol involved heating the sample to 800 °C at a rate of 10 °C min-1, holding the sample at this temperature for 1 h, and finally cooling it to room temperature. The resultant sample was denoted as AC-N (Fig. 1b). Furthermore, AC-G and AC-GN were prepared using the same procedure, with the only difference being the substitution of 1 g of urea with 1 g of glucose and a combination of 1 g urea and 1 g glucose, respectively.

The catalysts were prepared using an equal volume impregnation method to achieve a 3 wt% Ru loading. Specifically, AC, AC-N, AC-G, and AC-G-N were ultrasonically mixed for 20 min, allowed to stand for 2 h, and then dried at 80 °C for 12 h to obtain catalysts with 3 wt% Ru loading, designated as Ru/AC, Ru/AC-N, Ru/AC-G, and Ru/AC-GN, respectively (Fig. 1c). The prepared catalysts were stored in sealed conditions.

Methods

Sample characterization

The samples underwent degassing under an N2 atmosphere at 120 °C for 12 h. Specific surface area and pore volume were determined using an automatic surface area and porosity analyzer (BET, Beijing Jingwei Gaobo Company JW-BK132F, China). Surface composition analyses and semi-quantitative assessments of the samples were conducted using an X-ray photoelectron spectroscope (Thermo Scientific K-Alpha, USA). X-ray diffraction (XRD) scans were performed from 10° to 80° at a rate of 2° min-1 using a Rigaku D/max-2200pc instrument (Japan). Raman spectra were recorded with a Horiba LabRAM HR Evolution Raman spectrometer (Japan) using an excitation wavelength of 532 nm. Elemental composition was analyzed with an element analyzer (Elementar UNICUBE, Germany). Surface characteristics were explored via scanning electron microscopy (SEM, TESCAN MIRA LMS, Czech Republic), and the morphology and dispersion of catalysts, along with elemental types and contents, were investigated using a transmission electron microscope (JEOL JEM-F200). Crystal structure and elemental analyses were performed using a high-resolution transmission electron microscope combined with an energy-dispersive spectrometer. Thermogravimetric analysis (TG, Netzsch STA449F3, Germany) was performed under an N2 atmosphere from 30 to 800 °C with a heating rate of 10 °C min-1. H2 temperature-programmed reduction (H2-TPR) and N2 temperature-programmed desorption (N2-TPD) analyses were carried out using a temperature-programmed chemical adsorption instrument (Micromeritics Auto Chem II 2920, USA), with detailed procedures provided in the supplementary materials.

Catalyst Activity Tests

The catalytic activity of the various samples for ammonia decomposition was assessed at atmospheric pressure and 400 to 550 °C in a quartz tube reactor with an 8 mm inner diameter. Specifically, a 200 mg catalyst sample was activated under a 50% H2 and 50% Ar mixed gas flow at 300 °C for 2 h. Following reduction, the system was purged with NH3 for 30 min before initiating ammonia decomposition tests. Furthermore, the effluent gases (H2, N2, and NH3) were analyzed using a gas chromatograph equipped with a thermal conductivity detector and a Porapak-Q column (Zhejiang Fuli GC-9720P, China). The corresponding methodological and calculation details are provided in Fig. S1.

Experimental and detection data were compiled using Excel 2021, while linear fitting and graph plotting were performed using Origin 2021. XPS peak deconvolution and fitting were conducted using Avantage 6.6. Phase analysis was performed using MDI Jade 6, while particle size statistics and lattice fringe measurements were performed using Nano Measurer 1.2 and Digital Micrograph 3.5.

RESULTS AND DISCUSSION

Structural Characteristics of Carriers and Catalysts

Figure 2 presents the adsorption–desorption isotherms and pore size distributions of the various samples. The specific surface area and pore structure of a carrier play a crucial role in determining the dispersion and size of active metal particles in supported catalysts (Ren et al. 2023; Du et al. 2024). As shown in Fig 2a, AC, AC-N, AC-G, and AC-GN exhibited similar N2 adsorption–desorption isotherms, namely Type I isotherms with H1-type hysteresis loops (Calzaferri et al. 2023). The pore size distribution, depicted in Fig. 2b, indicates that the various modification techniques did not substantially alter the structure of AC, thereby preserving its hierarchical porous carbon architecture (micropore–mesopore–macropore), which facilitates the dispersion of active metals.

Fig. 2. N2 sorption isotherms, (a) and (c), and pore size distribution curves, (b) and (d), of various supports and catalysts

Further analysis of specific surface area and pore structure data from Table 1 revealed that AC-N exhibited the highest specific surface area (750 m²g-1) and pore volume (0.49 cm³g-1), representing increases of 27% and 8%, respectively, compared to AC, while the average pore diameter decreased by 13%. This decrease can be attributed to the etching effect on the carbon structure caused by radicals generated during urea decomposition, leading to enhanced porosity (Stöhr et al. 1991). By contrast, AC-G showed a slight decrease in average pore diameter, specific surface area, and pore volume relative to AC. This decrease is attributed to the deposition of new carbon structures formed by glucose on the surface and within the pores of AC. The specific surface area and pore volume of AC-GN were greater than those of AC-G, likely owing to urea decomposition. Given that the carbon microspheres formed during the hydrothermal reaction of glucose ranged between 50 to 200 nm in size (Liu et al. 2010), it is inferred that the carbon structures derived from glucose adhered to the external surface and large pores of AC. These findings suggest that the introduction of glucose did not significantly block the pores, thereby minimally impacting the micro and mesopores of AC. Consequently, this modification method preserved the original micro and mesoporous structures of the carrier alongside facilitating the formation of novel nitrogen-doped carbon structures on the outer surface and in the large pores. Figures 2c and 2d illustrate the post-catalyst-loading adsorption–desorption isotherms and pore size distribution curves. The most significant changes in average pore size and pore volume were observed in AC after loading, with reductions of 19% and 22%, respectively. These changes are likely owing to the agglomeration of Ru active metal atoms within the AC pores, which blocked some micropores. Similarly, the extent of changes observed in Ru/AC-G surpassed those observed in AC-N and AC-GN. Thus, consistent with initial observations, the carriers AC-N and AC-GN were more conducive to the dispersion of the Ru active metal atoms.

Table 1. Textural Properties of Various Carriers and Catalysts

Note: SBET : Brunauer-Emmett-Teller (BET) specific surface area

The SEM and TEM analyses were performed to characterize the morphologies of the carriers and catalysts. Large pores and mesopores were observed, as depicted in Fig. 3 parts a and b. The introduction of urea for nitrogen doping resulted in the generation of radicals at elevated temperatures, thereby etching the carbon carrier, consistent with the BET results. Carbon deposition on the surface and within the large pores of AC-GN is revealed in Fig. 3c. This occurred owing to the dehydration of glucose during the hydrothermal process at 180 °C, gradually separating from the solution to form an aqueous emulsion. Further dehydration of the sugar resulted in the formation of oligomeric nuclei within the sugar micelles, which subsequently grew into nanoscale spheres (Sakaki et al. 1996; Xu et al. 2008). These carbon microspheres, deposited on the carbon surface and further carbonized at 800 °C, were not observed in AC-GN. As depicted in Fig. 3d, the TEM analysis of AC-GN clearly revealed micro and mesoporous structures, confirming that the carbonization of glucose did not lead to the blockage of the micro and mesopores. The energy-dispersive spectroscopy (EDS) elemental distribution map (Fig. 3e) indicated the doping of nitrogen atoms into the aromatic carbon framework, with Ru nanoparticles being uniformly dispersed on the AC-GN carrier. Furthermore, graphitic carbon lattice fringes and Ru nanoparticles with a lattice spacing of 0.206 nm, corresponding to the Ru (101) plane, are evident in Fig. 3f-h (Agarwal and Ganguli 2014). In addition, the electron diffraction patterns revealed two distinct diffraction rings, corresponding to the (002) and (101) planes of graphitic carbon and Ru, respectively (Wang et al. 2023a).

Fig. 3. SEM images of AC (a), AC-N (b), and AC-GN (c). TEM images of AC-GN (d). EDS elemental mapping of Ru/AC-GN (e). High-resolution transmission electron microscopy image of Ru/AC-GN (f-h). SAED pattern of Ru/AC-GN.

The XRD was used to analyze the carbon crystalline structures and the metallic Ru crystalline structures of the four Ru-based catalysts. A prominent diffraction peak near 26° was observed for all the catalysts, as shown in Fig. 4a (Kulkarni et al. 2017). This peak is attributed to the (002) diffraction of graphitic carbon. There was no considerable difference in the diffraction peak at 26° between Ru/AC and Ru -N, indicating that the hydrothermal post-treatment nitrogen doping process had a minimal impact on the crystalline structure of the carbon. The diffraction peaks observed for Ru/AC-G and Ru/AC-GN were less intense and broader than those for AC and AC-N, signifying a slight variation in the crystallinity of the carbon carrier. This suggests that the nitrogen-doped carbon generated from the glucose–urea mixture was successfully deposited on the AC surface and underwent carbonization at 800 °C. Notably, distinct absorbance peaks corresponding to the Ru (101) plane were observed for Ru/AC and Ru/AC-G near 43°, while similar peaks were not observed for Ru/AC-G and Ru/AC-GN, in contrast with the TEM results(Tee et al. 2015). This discrepancy was attributed to the uniform dispersion of Ru nanoparticles on the AC-N and AC-GN carriers (Wen et al. 2018). Raman spectroscopy was employed to further analyze the graphitization degree of the four carbon carriers, as illustrated in Fig. 4b. Peaks were noted near 1345 cm-1 (D band) and 1575 cm-1 (G band), with a lower D to G band intensity ratio (ID: IG) indicating a higher level of graphitization (Odedairo et al. 2014). AC exhibited the lowest ID: IG ratio of 1.05, signifying the highest degree of graphitization. In the case of AC-N, the ID: IG ratio was 1.12, suggesting a reduction in graphitization owing to the disruption of carbon chains by nitrogen doping. Furthermore, AC-G exhibited an ID: IG value of 1.14, revealing that the hydrothermal carbon was graphitized at 800 °C, consistent with the XRD results. AC-GN displayed the highest ID: IG ratio of 1.23 among the carriers, indicating the lowest degree of graphitization and an increased number of defect structures. This is a consequence of the combined effects of the disruption of carbon chains by nitrogen atom doping and the deposition of nitrogen-doped carbon on the AC surface.

Fig. 4. XRD patterns of the catalyst samples (a) and Raman patterns of the support samples (b)

Table 2 presents the elemental analysis results of the different carbon carriers. The results indicate that the respective nitrogen contents in AC and AC-G, which were not doped with urea, were only 0.5 and 0.4 wt%. Upon the addition of urea alone, the nitrogen content increased to 1.8 wt%. This is because the generation of ammonia at high temperatures led to the formation of NH2, NH, and N radicals, thereby doping N atoms into the aromatic carbon framework. In AC-GN, the nitrogen content rose to 4.7 wt%, indicating that glucose addition not only facilitated the doping of N atoms into the highly aromatic carbon framework but also generated new nitrogen-doped carbon deposits on the AC surface, achieving in situ nitrogen doping. Notably, compared to traditional post-treatment nitrogen doping methods, this method resulted in increased nitrogen doping. The thermal stability of the carriers, a critical factor affecting ammonia decomposition catalyst activity at reaction temperatures between 400 and 550 °C, was determined through TG. The TG curves of the four carbon carriers (Fig. 5a) exhibited similar weight loss profiles, with varying degrees of loss before 200 °C, primarily due to the vaporization of water. For AC-G and AC-GN, slight mass loss was observed between 500 and 600 °C. Ultimately, the mass loss for AC and AC-G was within 1%, while that for AC-G and AC-GN was approximately 2%, presumed to be caused by the decomposition of certain unstable structures generated by glucose carbonization at high temperatures(Liu et al. 2020). Given that the optimal reaction temperature for Ru/AC-GN in ammonia decomposition is around 500 °C, AC-GN possesses adequate thermal stability.

Table 2. Summary of the Elemental Analysis Results of the Various Carriers

Fig. 5. Mass loss of various carrier composites as a function of temperature between 50 and 600 °C, measured with a heating rate of 10 °C/min under a nitrogen atmosphere (a). XPS survey spectra of AC-N, AC-N, AC-G, and AC-GN (b). XPS N1s spectra of AC-N and AC-GN (c)

X-ray photoelectron spectroscopy (XPS) was employed to investigate the elemental composition and chemical bonds of the carriers. The XPS spectra of the Ru/AC-N and Ru/AC-GN samples revealed deconvoluted peaks attributed to C 1s, N 1s, and O 1s (Fig. 5b)(Yu et al. 2014). By contrast, Ru/AC and Ru/AC-G lacked a peak corresponding to N 1s, and the N1S peak was substantially higher in AC-GN, consistent with the elemental analysis results. The high-resolution N1s peak in Fig. 5c revealed the presence of pyridinic-N, pyrrolic-N, graphitic-N, and oxide-N (Han et al. 2019). Nitrogen species play a crucial role in the performance of the carriers, as shown in Table S1, which displays the content of nitrogen species in the catalysts obtained from the two nitrogen doping methods. Notably, the highest pyridinic-N and pyrrolic-N contents were observed in Ru/AC-N and Ru/AC-GN, respectively. Furthermore, Ru/AC-GN exhibited a higher content of pyrrolic-N compared to Ru/AC-N, while Ru/AC-N had a higher pyridinic-N content. Studies have indicated that pyridinic-N and pyrrolic-N can serve as effective sites for active metals, enhancing the dispersion of active metal components and facilitating the binding between metallic Ru and nitrogen-doped carbon carriers (Zhong and Aika 1998). However, the mechanism of action of these two nitrogen species remains unclear (Miao et al. 2017; Wang et al. 2023b). With increasing pyrolysis temperature, the pyridinic-N content decreased gradually, transforming into pyrrolic and graphitic nitrogen. Excessive pyridinic-N appears unsuitable for high-temperature ammonia decomposition systems. Therefore, the AC-GN carrier, with high pyrrolic-N and graphitic-N contents, is deemed more suitable.

Structure–Activity Relationship between Carriers and Active Metals

Currently, there exists a consensus regarding the mechanism of NH3 decomposition on active metal surfaces, which can be delineated into eight steps (Fig. 6). These steps encompass ammonia adsorption, dissociation, and the formation and desorption of decomposition products (Chen et al. 2021; Su et al. 2024). Initially, gaseous NH3(g) adsorbs on the active sites of the catalyst(ad), generating adsorbed NH3(ad) (step 1). Subsequently, NH3(ad) dehydrogenates to produce adsorbed N(ad) and H(ad) (steps 2 to 4). Eventually, N(ad) and H(ad) combine to form adsorbed N2(ad) and H2(ad) (steps 5 to 6), respectively, which subsequently desorb to yield gaseous N2(g) and H2(g) (steps 7 and 8). However, the rate-limiting step remains a subject of debate. Enhanced Ru nanoparticle dispersion facilitates the conversion of reaction gases between gaseous and adsorbed states (step 1) or from adsorbed states back to gaseous states (steps 7 and 8). Moreover, the efficiency of gas transport also influences the aforementioned reaction process, necessitating the provision of more anchoring sites and a porous structure in the carrier.

Fig. 6. Schematic of NH3 decomposition reaction in the presence of Ru catalyst

As previously mentioned, the nitrogen-doped carbon structures deposited in AC-GN predominantly reside on the outer surface and within large pores, preserving a hierarchical porous structure conducive to gas transport. Additionally, the outer surface and large pores of AC-GN offer more defect sites, conducive to anchoring Ru, thereby enhancing its dispersion and stability. Wu et al. (2023) found that active metals loaded on the outer surface of the carrier exhibit superior ammonia decomposition performance, attributed to the ease of gas adsorption–desorption on the outer surface. Consequently, Ru situated on the outer surface and large pores of AC-GN would theoretically enhance the overall catalytic performance. The gradual dissociation of NH3(ad) (3-5) also plays a pivotal role in influencing the catalytic reaction. Wang et al. (2023b) observed that the electron density on the outer surface of CNTs surpasses that on the inner surface, resulting in heightened activity of the active metals loaded on the outer surface. The introduction of N alters the electron distribution of C, inducing electron-rich states and enhancing the electron transfer capability. Furthermore, AC-GN retains its original degree of graphitization, accompanied by a notable improvement in the electrical conductivity of its outer surface, thereby facilitating the dissociation of NH3(ad).

To determine the dispersion of Ru nanoparticles on the carrier, the diameters of at least 50 Ru particles were analyzed statistically. Figure S2 presents TEM images of three catalysts, namely Ru/AC, Ru/AC-N, and Ru/AC-GN, with Ru nanoparticle diameters of 3.39, 2.98, and 2.79 nm, respectively. The Ru nanoparticles in Ru/AC-GN exhibited the smallest diameter and the most optimal dispersion. To further elucidate the presence of the active metal Ru on AC-N and AC-GN compared to AC, Ru 3p spectra were deconvoluted and fitted. Figure 7a-c illustrates two main peaks corresponding to Ru3p3/2 and Ru3p1/2. The peaks near 462.4 and 484.7 eV correspond to Ru0, whereas those at 464.4 and 486.7 eV correspond to Run+ (Tang et al. 2015; Wang et al. 2018). The relative contents of Ru0 and Run+ were determined based on the peak areas (Fig. 7d). The Ru0 contents of Ru/AC, Ru/AC-N, and Ru/AC-GN were found to be 50%, 70%, and 74%, respectively, indicating that the Ru0 content increases with nitrogen content. This demonstrates that nitrogen doping can regulate the electron cloud density of the carbon carrier, thereby enhancing electron transfer between the metal and carrier. Consequently, electron transfer occurs between N atoms and metallic Ru in Ru/AC-N and Ru/AC-GN, resulting in higher Ru0 content. Notably, among the four carriers, AC-GN exhibited the greatest performance.

Fig. 7. Ru 3p XPS spectra of Ru/AC (a), Ru/AC-N (b), Ru/AC-GN (c), and Ru0 and Run+ contents of various catalysts (d)

The H2-TPR curves of the unreduced catalysts indicated that 300 °C is a suitable reduction temperature for Ru (Fig. 8a). Excessively high reduction temperatures might result in Ru particle agglomeration, while temperatures below 300 °C might not effectively reduce Ru oxides to Ru. In the H2-TPR spectra, the peaks observed for Ru/AC-GN showed an increase of approximately 10 °C compared to Ru/AC-N, a consequence of the interactions between the nitrogen-doped carriers and Ru oxides, which intensified with increased nitrogen content (Chen et al. 2017). Such metal–carrier interactions prevent Ru nanoparticle agglomeration. Furthermore, the issue of catalyst poisoning is crucial to the study of catalyst performance. It has been reported that Ru-based catalysts are primarily affected by hydrogen poisoning in ammonia synthesis reactions (Lin et al. 2021), while activated carbon carriers may generate gases such as COX (CO, CO2) at high temperatures. Due to the porous nature of activated carbon, these COX and H2 molecules excessively adsorb on the catalyst surface, occupying active sites and thereby inhibiting or hindering the catalytic reaction (Chen et al. 2024). Since the treatment temperature of activated carbon (800 °C) is much higher than the temperature of ammonia decomposition reactions, unstable oxygen-containing functional groups are decomposed in advance, making the impact of CO on the catalyst negligible. Ren and others have studied the desorption behavior of hydrogen on catalysts through H2-TPD, finding that hydrogen begins to desorb at approximately 100 °C and is essentially fully desorbed by around 200 °C (Ren et al. 2017). Under the conditions of ammonia decomposition reactions at 400 to 500 °C, hydrogen desorbs more readily. This suggests that nitrogen may have a more significant impact on catalyst performance in ammonia decomposition reactions. Current research considers nitrogen desorption as the rate-controlling step in ammonia decomposition on Ru catalysts (Zhang et al. 2014). The N2-TPD curves exhibited three peaks (Fig. 8b) near 140, 450, and 720 °C. The desorption peak at 140 °C, attributed to the physical desorption of N2 from the carbon surface and pores, suggests that the other two peaks at higher temperatures correspond to deep pore desorption and chemical adsorption. Ru/AC-GN desorbed N2 primarily between 400 and 500 °C, whereas Ru/AC-N displayed a minor peak after 700 °C. This is possibly owing to the larger surface area and smaller average pore size of AC-N, leading to greater N2 desorption resistance in the pores of Ru/AC-N. It is anticipated that Ru loaded on AC-GN will demonstrate superior NH3 decomposition performance.

Fig. 8. H2-TPR (a) and N2-TPD (b) profiles of the catalysts fabricated in this study

Assessment of NH3 Decomposition Performance

Figure 9a demonstrates that the NH3 conversion rates of the four catalysts exhibit a direct correlation with temperature. NH3 decomposition is an endothermic reaction. Therefore, the NH3 conversion rate rises with increasing reaction temperature (Wu et al. 2023). Among the four catalysts, Ru/AC-GN displayed the highest activity, followed by Ru/AC-N, whereas Ru/AC and Ru/AC-G showed no significant difference in activity. Notably, within the low-temperature range of 425 to 500 °C, Ru/AC-GN exhibited a markedly superior conversion efficiency compared to the other catalysts. At a gas hourly space velocity (GHSV) of 15000 mL gcat-1 h-1, Ru/AC-GN achieved 69% NH3 conversion at 475 °C, corresponding to an H2 generation rate of 11.5 mmol g cat -1 min-1 (Table S2). Moreover, at 525 °C, it achieved 94% NH3 conversion, equivalent to an H2 generation rate of 15.8 mmol g cat -1 min-1. Figure 9b presents Arrhenius plots illustrating the logarithm of the H2 formation rate versus the reciprocal of the absolute temperature. The determined apparent activation energies (Ea) for Ru/AC, Ru/AC-N, and Ru/AC-GN were78.13, 61.80 and 60.28 kJ mol-1, respectively. As previously discussed, the dispersion of Ru on the outer surface and large pores of AC-GN facilitated higher gas adsorption–desorption rates, resulting in the lowest Ea and high activity at low temperatures in the case of Ru/AC-GN. Compared with previously reported Ru-based carbon-supported NH3 decomposition catalysts (Table S3), the performance of Ru/AC-GN is significantly superior, and it is comparable to the high-performing MgO, offering a cost-effective alternative through the valorization of biomass resources (Wang et al. 2004; Lucentini et al. 2019).

Fig. 9. NH3 conversion in the presence of various catalysts at a GHSV of 15000 mL gcat-1 h-1 (a) and the Arrhenius plots obtained between 400 and 475 ℃ (b)

The catalytic stability of Ru/AC-GN for NH3 decomposition was evaluated at 500 °C with a GHSV of 15000 mL gcat-1 h-1. Over a 24 h reaction period, Ru/AC-N and Ru/AC-GN demonstrated substantially better stability than Ru/AC (Fig. 10a), with Ru/AC-GN exhibiting particularly notable effects owing to the interactions between the nitrogen-doped structures and Ru within the outer surface and large pores of the AC-GN carrier. The XRD patterns in Fig. 10b show no significant change in Ru size after 24 h of durability testing. However, the stability of the Ru/AC-GN catalyst must be further improved. The addition of a second metal or promoters could further enhance the low-temperature activity of the catalyst and reduce the reaction temperature, thereby extending the catalyst lifespan.

Fig. 10. Stability test results of various catalysts for NH3 decomposition at 500 ℃ with a GHSV of 15000 mL gcat-1 h-1 (a), and XRD patterns of the aged catalyst samples (b)

CONCLUSIONS

  1. Nitrogen-doped composite materials were successfully synthesized via a hydrothermal pyrolysis method. Unlike conventional nitrogen doping techniques, this approach not only introduces N atoms into carbon chains but also deposits nitrogen-containing carbon materials on the surface and within the large pores of AC.
  2. By incorporating glucose, in situ nitrogen doping was achieved, boosting the nitrogen content of the carbon carrier while preserving its hierarchical pore structure and degree of graphitization.
  3. This methodology improved the effective dispersion of Ru nanoparticles on the external surface and within the large pores of the carrier, enhancing electron transfer capabilities. Leveraging of the inherent advantage of easier gas adsorption–desorption on the external surface of the catalyst, the low-temperature activity of the catalyst was increased in this study.
  4. Examination of nitrogen species revealed that nitrogen doping enhanced metal–carrier interactions, effectively improving the thermal stability of the catalyst. The relatively stable pyrrolic and graphitic nitrogen were more suitable for the ammonia decomposition reaction.
  5. Finally, catalytic performance tests demonstrated that the Ru/AC-GN catalyst exhibited outstanding activity and catalytic stability for NH3 decomposition, notably surpassing other carbon-supported Ru catalysts at low temperatures. This nitrogen doping method in AC, characterized by relatively high nitrogen doping levels, excellent hierarchical pore structures, and superior catalytic performance, provides valuable insights into the design and preparation of catalyst carriers for ammonia decomposition reactions.

ACKNOWLEDGEMENTS

This study was supported by the National Natural Science Foundation of China (No. 32202606); Youth Foundation Natural Science Foundation of Shandong Province(No. ZR2022QC056); The central government guides local Science and Technology Development Fund Projects-Yellow River Basin-Xinjiang (No. YDZX2023007); The central government guides local Science and Technology Development Fund Projects-Yellow River Basin-Shanxi (No. YDZX2023024); Jinan City science and technology plan social livelihood special (No. 202317009); 20 New Items of Universities” funding project of Jinan (No. 202228043); Science, education and industry integration pilot training fund (No. 2023PY045); Shandong Province Science and Technology-based Small and Medium-sized Enterprises Innovation Capacity Enhancement Project (No. 2023TSGC0235); R&D and Demonstration of Key Technologies for Green and Efficient Hydrogen Production (No. 22JBZ02-03); Talent Research Project of Qilu University of Technology (Shandong Academy of Sciences) (2023RCKY170); Shandong Province building and transportation double carbon innovation joint venture research and development plan (STGTT0101202302) and (STGTT0101202303).

REFERENCES CITED

Abe, J. O., Popoola, A. P. I., Ajenifuja, E., and Popoola, O. M. (2019). “Hydrogen energy, economy and storage: Review and recommendation,” International Journal of Hydrogen Energy 44(29), 15072-15086. DOI: 10.1016/j.ijhydene.2019.04.068

Agarwal, S., and Ganguli, J. N. (2014). “Hydrogenation by nanoscale ruthenium embedded into the nanopores of K-10 clay,” RSC Advances 4(23), article 11893. DOI: 10.1039/c3ra47162d

Calzaferri, G., Gallagher, S. H., Lustenberger, S., Walther, F., and Brühwiler, D. (2023). “Multiple equilibria description of type H1 hysteresis in gas sorption isotherms of mesoporous materials,” Materials Chemistry and Physics 296, article 127121. DOI: 10.1016/j.matchemphys.2022.127121

Cao, C.-F., Wu, K., Zhou, C., Yao, Y.-H., Luo, Y., Chen, C.-Q., Lin, L., and Jiang, L. (2022). “Electronic metal-support interaction enhanced ammonia decomposition efficiency of perovskite oxide supported ruthenium,” Chemical Engineering Science 257, article 117719. DOI: 10.1016/j.ces.2022.117719

Chang, F., Gao, W., Guo, J., and Chen, P. (2021). “Emerging materials and methods toward ammonia‐based energy storage and conversion,” Advanced Materials 33(50), article 2005721. DOI: 10.1002/adma.202005721

Chen, C., Wu, K., Ren, H., Zhou, C., Luo, Y., Lin, L., Au, C., and Jiang, L. (2021). “Ru-based catalysts for ammonia decomposition: A mini-review,” Energy & Fuels 35(15), 11693-11706. DOI: 10.1021/acs.energyfuels.1c01261

Chen, L., Li, Y., Zhang, X., Zhang, Q., Wang, T., and Ma, L. (2017). “Effect of Ru particle size on hydrogenation/decarbonylation of propanoic acid over supported Ru catalysts in aqueous phase,” Catalysis Letters 147(1), 29-38. DOI: 10.1007/s10562-016-1877-4

Chen, S.-Y., Wang, L.-Y., Chen, K.-C., Yeh, C.-H., Hsiao, W.-C., Chen, H.-Y., Nishi, M., Keller, M., Chang, C.-L., Liao, C.-N., Mochizuki, T., Chen, H.-Y. T., Chou, H.-H., and Yang, C.-M. (2024). “Ammonia synthesis over cesium-promoted mesoporous-carbon-supported ruthenium catalysts: Impact of graphitization degree of the carbon support,” Applied Catalysis B: Environment and Energy 346, article 123725. DOI: 10.1016/j.apcatb.2024.123725

Du, H.-G., Zhang, X.-F., Ding, L.-W., Liu, J.-L., Yu, L.-H., Zhang, X.-H., Dou, Y., Cao, L.-M., Zhang, J., and He, C.-T. (2024). “Engineering pore-size distribution of metal-loaded carbon catalysts by in situ cavitation for boosting electrochemical mass transfer,” Applied Catalysis B: Environmental 342, article 123396. DOI: 10.1016/j.apcatb.2023.123396

Han, F., Liu, Z., Jia, J., Ai, J., Liu, L., Liu, J., and Wang, Q.-D. (2019). “Influences of N species in N-doped carbon carriers on the catalytic performance of supported Pt,” Materials Chemistry and Physics 237, article 121881. DOI: 10.1016/j.matchemphys.2019.121881

Hien, N., Kim, H., Jeon, M., Lee, J., Ridwan, M., Tamarany, R., and Yoon, C. (2015). “Ru-N-C Hybrid nanocomposite for ammonia dehydrogenation: Influence of N-doping on catalytic activity,” Materials 8(6), 3442-3455. DOI: 10.3390/ma8063442

Karim, A. M., Prasad, V., Mpourmpakis, G., Lonergan, W. W., Frenkel, A. I., Chen, J. G., and Vlachos, D. G. (2009). “Correlating particle size and shape of supported Ru/γ-Al2O3 catalysts with NH3 decomposition activity,” Journal of the American Chemical Society 131(34), 12230-12239. DOI: 10.1021/ja902587k

Kulkarni, G., Velhal, N., Phadtare, V., and Puri, V. (2017). “Enhanced electromagnetic interference shielding effectiveness of chemical vapor deposited MWCNTs in X-band region,” Journal of Materials Science: Materials in Electronics 28(10), 7212-7220. DOI: 10.1007/s10854-017-6402-z

Lee, Y.-J., Cha, J., Kwak, Y., Park, Y., Jo, Y. S., Jeong, H., Sohn, H., Yoon, C. W., Kim, Y., Kim, K.-B., and Nam, S. W. (2021). “Top-down syntheses of nickel-based structured catalysts for hydrogen production from ammonia,” ACS Applied Materials & Interfaces, 13(1), 597-607. DOI: 10.1021/acsami.0c18454

Lin, B., Wu, Y., Fang, B., Li, C., Ni, J., Wang, X., Lin, J., and Jiang, L. (2021). “Ru surface density effect on ammonia synthesis activity and hydrogen poisoning of ceria-supported Ru catalysts,” Chinese Journal of Catalysis 42(10), 1712-1723. DOI: 10.1016/S1872-2067(20)63787-1

Liu, H., Chen, B., and Wang, C. (2020). “Pyrolysis kinetics study of biomass waste using shuffled complex evolution algorithm,” Fuel Processing Technology 208, article 106509. DOI: 10.1016/j.fuproc.2020.106509

Liu, S., Sun, J., and Huang, Z. (2010). “Carbon spheres/activated carbon composite materials with high Cr(VI) adsorption capacity prepared by a hydrothermal method,” Journal of Hazardous Materials 173(1), 377-383. DOI: 10.1016/j.jhazmat.2009.08.086

Lucentini, I., Casanovas, A., and Llorca, J. (2019). “Catalytic ammonia decomposition for hydrogen production on Ni, Ru and Ni Ru supported on CeO2,” International Journal of Hydrogen Energy 44(25), 12693-12707. DOI: 10.1016/j.ijhydene.2019.01.154

Luo, Y., Yang, Q., Nie, W., Yao, Q., Zhang, Z., and Lu, Z.-H. (2020). “Anchoring IrPdAu nanoparticles on NH2 -SBA-15 for fast hydrogen production from formic acid at room temperature,” ACS Applied Materials & Interfaces 12(7), 8082-8090. DOI: 10.1021/acsami.9b16981

Mehrpooya, M., and Habibi, R. (2020). “A review on hydrogen production thermochemical water-splitting cycles,” Journal of Cleaner Production 275, artcile 123836. DOI: 10.1016/j.jclepro.2020.123836

Miao, H., Li, S., Wang, Z., Sun, S., Kuang, M., Liu, Z., and Yuan, J. (2017). “Enhancing the pyridinic N content of nitrogen-doped graphene and improving its catalytic activity for oxygen reduction reaction,” International Journal of Hydrogen Energy 42(47), 28298-28308. DOI: 10.1016/j.ijhydene.2017.09.138

Mukherjee, S., Devaguptapu, S. V., Sviripa, A., Lund, C. R. F., and Wu, G. (2018). “Low-temperature ammonia decomposition catalysts for hydrogen generation,” Applied Catalysis B: Environmental 226, 162-181. DOI: 10.1016/j.apcatb.2017.12.039

Nagaoka, K., Eboshi, T., Abe, N., Miyahara, S., Honda, K., and Sato, K. (2014). “Influence of basic dopants on the activity of Ru/Pr6O11 for hydrogen production by ammonia decomposition,” International Journal of Hydrogen Energy 39(35), 20731-20735. DOI: 10.1016/j.ijhydene.2014.07.142

Nakamura, I., Kubo, H., and Fujitani, T. (2022). “Critical role of Cs doping in the structure and NH3 decomposition performance of Ru/MgO catalysts,” Applied Catalysis A: General 644, article 118806. DOI: 10.1016/j.apcata.2022.118806

Odedairo, T., Ma, J., Gu, Y., Chen, J., Zhao, X. S., and Zhu, Z. (2014). “One-pot synthesis of carbon nanotube–graphene hybrids via syngas production,” J. Mater. Chem. A 2(5), 1418-1428. DOI: 10.1039/C3TA13871B

Rarogpilecka, W., Miskiewicz, E., Szmigiel, D., and Kowalczyk, Z. (2005). “Structure sensitivity of ammonia synthesis over promoted ruthenium catalysts supported on graphitised carbon,” Journal of Catalysis 231(1), 11-19. DOI: 10.1016/j.jcat.2004.12.005

Raróg-Pilecka, W., Szmigiel, D., Komornicki, A., Zieliński, J., and Kowalczyk, Z. (2003). “Catalytic properties of small ruthenium particles deposited on carbon,” Carbon 41(3), 589-591. DOI: 10.1016/S0008-6223(02)00393-7

Ren, R., Dou, B., Zhang, H., Wu, K., Wang, Y., Chen, H., and Xu, Y. (2023). “Syngas production from CO2 reforming of glycerol by mesoporous Ni/CeO2 catalysts,” Fuel 341, article 127717. DOI: 10.1016/j.fuel.2023.127717

Ren, S., Huang, F., Zheng, J., Chen, S., and Zhang, H. (2017). “Ruthenium supported on nitrogen-doped ordered mesoporous carbon as highly active catalyst for NH3 decomposition to H2,” International Journal of Hydrogen Energy 42(8), 5105-5113. DOI: 10.1016/j.ijhydene.2016.11.010

Rodríguez-reinoso, F. (1998). “The role of carbon materials in heterogeneous catalysis,” Carbon, 36(3), 159-175. DOI: 10.1016/S0008-6223(97)00173-5

Sakaki, T., Shibata, M., Miki, T., Hirosue, H., and Hayashi, N. (1996). “Decomposition of cellulose in near-critical water and fermentability of the products,” Energy & Fuels 10(3), 684-688. DOI: 10.1021/ef950160+

Stöhr, B., Boehm, H. P., and Schlögl, R. (1991). “Enhancement of the catalytic activity of activated carbons in oxidation reactions by thermal treatment with ammonia or hydrogen cyanide and observation of a superoxide species as a possible intermediate,” Carbon 29(6), 707-720. DOI: 10.1016/0008-6223(91)90006-5

Su, Z., Guan, J., Liu, Y., Shi, D., Wu, Q., Chen, K., Zhang, Y., and Li, H. (2024). “Research progress of ruthenium-based catalysts for hydrogen production from ammonia decomposition,” International Journal of Hydrogen Energy 51, 1019-1043. DOI: 10.1016/j.ijhydene.2023.09.107

Sun, S., Jiang, Q., Zhao, D., Cao, T., Sha, H., Zhang, C., Song, H., and Da, Z. (2022). “Ammonia as hydrogen carrier: Advances in ammonia decomposition catalysts for promising hydrogen production,” Renewable and Sustainable Energy Reviews 169, article 112918. DOI: 10.1016/j.rser.2022.112918

Tan, H., Li, K., Sioud, S., Cha, D., Amad, M. H., Hedhili, M. N., and Al-Talla, Z. A. (2012). “Synthesis of Ru nanoparticles confined in magnesium oxide-modified mesoporous alumina and their enhanced catalytic performance during ammonia decomposition,” Catalysis Communications 26, 248-252. DOI: 10.1016/j.catcom.2012.06.007

Tang, M., Mao, S., Li, M., Wei, Z., Xu, F., Li, H., and Wang, Y. (2015). “RuPd alloy nanoparticles supported on N-doped carbon as an efficient and stable catalyst for benzoic acid hydrogenation,” ACS Catalysis 5(5), 3100-3107. DOI: 10.1021/acscatal.5b00037

Tee, S. Y., Lee, C. J. J., Dinachali, S. S., Lai, S. C., Williams, E. L., Luo, H.-K., Chi, D., Andy Hor, T. S., and Han, M.-Y. (2015). “Amorphous ruthenium nanoparticles for enhanced electrochemical water splitting,” Nanotechnology 26(41), article 415401. DOI: 10.1088/0957-4484/26/41/415401

Wan, Z., Tao, Y., Shao, J., Zhang, Y., and You, H. (2021). “Ammonia as an effective hydrogen carrier and a clean fuel for solid oxide fuel cells,” Energy Conversion and Management 228, article 113729. DOI: 10.1016/j.enconman.2020.113729

Wang, J., Wei, Z., Mao, S., Li, H., and Wang, Y. (2018). “Highly uniform Ru nanoparticles over N-doped carbon: pH and temperature-universal hydrogen release from water reduction,” Energy & Environmental Science 11(4), 800-806. DOI: 10.1039/C7EE03345A

Wang, S. J., Yin, S. F., Li, L., Xu, B. Q., Ng, C. F., and Au, C. T. (2004). “Investigation on modification of Ru/CNTs catalyst for the generation of COx-free hydrogen from ammonia,” Applied Catalysis B: Environmental 52(4), 287-299. DOI: 10.1016/j.apcatb.2004.05.002

Wang, Z., Chen, X., Sun, Y., Hua, D., Yang, S., Sun, L., Li, T., and Chen, L. (2023a). “Co-pyrolysis induced strong metal-support interaction in N-doped carbon supported Ni catalyst for the hydrogenolysis of lignin,” Chemical Engineering Journal 473, 145182. DOI: 10.1016/j.cej.2023.145182

Wang, Z., Chen, X., Xie, X., Yang, S., Sun, L., Li, T., Chen, L., and Hua, D. (2023b). “Synthesis of aromatic monomers via hydrogenolysis of lignin over nickel catalyst supported on nitrogen-doped carbon nanotubes,” Fuel Processing Technology 248, article 107810. DOI: 10.1016/j.fuproc.2023.107810

Wen, G., Gu, Q., Liu, Y., Schlögl, R., Wang, C., Tian, Z., and Su, D. S. (2018). “Biomass‐derived graphene‐like carbon: Efficient metal‐free carbocatalysts for epoxidation,” Angewandte Chemie International Edition 57(51), 16898-16902. DOI: 10.1002/anie.201809970

Wu, Z.-W., Xiong, J., Wang, C.-W., and Qin, Y.-H. (2023). “Supporting high-loading Ni on SBA-15 as highly active and durable catalyst for ammonia decomposition reaction,” International Journal of Hydrogen Energy 48(12), 4728-4737. DOI: 10.1016/j.ijhydene.2022.11.050

Xu, Y., Weinberg, G., Liu, X., Timpe, O., Schlögl, R., and Su, D. S. (2008). “Nanoarchitecturing of activated carbon: Facile strategy for chemical functionalization of the surface of activated carbon,” Advanced Functional Materials 18(22), 3613-3619. DOI: 10.1002/adfm.200800726

Yao, L., Shi, T., Li, Y., Zhao, J., Ji, W., and Au, C.-T. (2011). “Core–shell structured nickel and ruthenium nanoparticles: Very active and stable catalysts for the generation of COx-free hydrogen via ammonia decomposition,” Catalysis Today 164(1), 112-118. DOI: 10.1016/j.cattod.2010.10.056

Yin, S. F., Xu, B. Q., Zhou, X. P., and Au, C. T. (2004a). “A mini-review on ammonia decomposition catalysts for on-site generation of hydrogen for fuel cell applications,” Applied Catalysis A: General 277(1–2), 1-9. DOI: 10.1016/j.apcata.2004.09.020

Yin, S.-F., Xu, B.-Q., Ng, C.-F., and Au, C.-T. (2004b). “Nano Ru/CNTs: A highly active and stable catalyst for the generation of COx-free hydrogen in ammonia decomposition,” Applied Catalysis B: Environmental 48(4), 237-241. DOI: 10.1016/j.apcatb.2003.10.013

Yu, J., Guo, M., Muhammad, F., Wang, A., Zhang, F., Li, Q., and Zhu, G. (2014). “One-pot synthesis of highly ordered nitrogen-containing mesoporous carbon with resorcinol–urea–formaldehyde resin for CO2 capture,” Carbon 69, 502-514. DOI: 10.1016/j.carbon.2013.12.058

Zhang, H., Alhamed, Y. A., Al-Zahrani, A., Daous, M., Inokawa, H., Kojima, Y., and Petrov, L. A. (2014). “Tuning catalytic performances of cobalt catalysts for clean hydrogen generation via variation of the type of carbon support and catalyst post-treatment temperature,” International Journal of Hydrogen Energy 39(31), 17573-17582. DOI: 10.1016/j.ijhydene.2014.07.183

Zheng, W., Zhang, J., Xu, H., and Li, W. (2007). “NH3 decomposition kinetics on supported Ru clusters: Morphology and particle size effect,” Catalysis Letters 119(3–4), 311-318. DOI: 10.1007/s10562-007-9237-z

Zhong, Z., and Aika, K. (1998). “The effect of hydrogen treatment of active carbon on Ru catalysts for ammonia synthesis,” Journal of Catalysis 173(2), 535-539. DOI: 10.1006/jcat.1997.1943

Article submitted: March 7, 2024; Peer review completed: April 24, 2024; Revised version received and accepted: April 27, 2024; Published: May 15, 2024.

DOI: 10.15376/biores.19.3.4313-4334

 

APPENDIX

Supplementary

Appendix contains:

Section S1: Sample characterization (H2-TPR and N2-TPD)

Section S2: Catalyst Activity Tests

Table S1: N1s XPS analysis of various carbon materials.

Figure. S2: TEM images and Ru particle size distribution histograms of Ru/AC (a, d), Ru/AC-N (b, e) and Ru/AC-GN (c, f).

Table S2: H2 formation rate (mmol min-1 g-1cat) for various catalysts.

Table S3 Catalytic activity of the supported Ru catalysts for NH3 decomposition.

Section S1 Sample characterization (H2-TPR and N2-TPD)

For the H2-TPR test, 50 mg of unreduced catalyst loaded in a U-shaped quartz tube was initially heated from room temperature to 100 ℃ at a heating rate of 10 ℃ min-1 and held at 100 ℃ for 1 h in an Ar gas flow (50 mL min-1) to remove impurities such as water. Subsequently, it was heated from 50 to 800 ℃ at a rate of 10 ℃ min-1 in a 10% H2/Ar gas flow (50 mL min-1), and the amount of H2 consumed during this heating process was recorded using a thermal conductivity detector (TCD).

For the N2-TPD test, 50 mg of reduced catalyst loaded in a U-shaped quartz tube was first heated from room temperature to 150 ℃ at a rate of 10 ℃ min-1 and maintained at 150 ℃ for 1 h in a He gas flow (50 mL min-1) to remove impurities such as water. After cooling to 50 °C, it was purged with nitrogen for 1 h to saturation, and finally, under a helium atmosphere, ramped to 800 °C at a rate of 10 °C min-1 for desorption. The desorbed gases were detected using a TCD.

Section S2 Catalyst Activity Tests

Fig. S1. Schematic of the experimental setup used for the NH3 decomposition test

A pure ammonia (NH3) flow rate of 50 mL min-1, corresponding to a GHSV of 15,000 mL gcat-1 h-1, was maintained. Temperature-dependent NH3 decomposition tests were conducted in the temperature range of 400 to 550 °C in 25 °C increments (60 min at each temperature). Measurements were taken three times at each temperature point, and the average was calculated. The NH3 conversion rate () and H2 formation rate (r) were calculated according to Eqs. (1) and (2):

where VN2 and VNH3 are the molar flow rates of N2 and NH3 in the outlet streams, respectively.

The apparent activation energy of the catalyst was determined based on the Arrhenius equation (3) in the temperature range of 400 to 475 °C. Subsequently, a linear regression was conducted between ln r and 1000/T, yielding a linear relationship from which the apparent activation energy was derived.

where r represents the rate constant of the reaction, A represents the pre-exponential factor, Ea represents the activation energy, R represents the universal gas constant, and T represents the thermodynamic temperature.

Table S1. N1s XPS Analysis of Various Carbon Materials

Fig. S2. TEM images and Ru particle size distribution histograms of Ru/AC (a,d), Ru/AC-N (b,e) and Ru/AC-GN (c,f)

Table S2. H2 Formation Rate (mmol min-1 g-1cat) for Various Catalysts

Table S3. Catalytic Activity of the Supported Ru Catalysts for NH3 Decomposition

Note: [a] represents: wt%; [b]: °C; [c]: mL h-1 g -1cat; [d]: %; [e]: mmol min-1 gcat-1

References Cited

Bell, T. E., Zhan, G., Wu, K., Zeng, H. C., and Torrente-Murciano, L. (2017). “Modification of ammonia decomposition activity of ruthenium nanoparticles by N-doping of CNT Supports,” Topics in Catalysis 60(15–16), 1251-1259. DOI: 10.1007/s11244-017-0806-0

El-kholany, M. R., Kishimoto, N., Tanaka, K., Takamura, H., and Kadota, I. (2023). “Efficient method for the preparation of ozonides under dry conditions,” Bulletin of the Chemical Society of Japan 96(12), 1316-1318. DOI: 10.1246/bcsj.20230195

Li, L., Zhu, Z. H., Yan, Z. F., Lu, G. Q., and Rintoul, L. (2007). “Catalytic ammonia decomposition over Ru/carbon catalysts: The importance of the structure of carbon support,” Applied Catalysis A: General, 320, 166-172. DOI: 10.1016/j.apcata.2007.01.029

Yin, S. F., Xu, B. Q., Zhu, W. X., Ng, C. F., Zhou, X. P., and Au, C. T. (2004a). “Carbon nanotubes-supported Ru catalyst for the generation of COx-free hydrogen from ammonia,” Catalysis Today, Selections from the presentations of the 3rd Asia-Pacific Congress on Catalysis, 93–95, 27-38. DOI: 10.1016/j.cattod.2004.05.011

Yin, S.-F., Xu, B.-Q., Ng, C.-F., and Au, C.-T. (2004b). “Nano Ru/CNTs: A highly active and stable catalyst for the generation of COx-free hydrogen in ammonia decomposition,” Applied Catalysis B: Environmental 48(4), 237-241. DOI: 10.1016/j.apcatb.2003.10.013

Yin, S.-F., Zhang, Q.-H., Xu, B.-Q., Zhu, W.-X., Ng, C.-F., and Au, C.-T. (2004c). “Investigation on the catalysis of COx-free hydrogen generation from ammonia,” Journal of Catalysis, 224(2), 384–396. DOI: 10.1016/j.jcat.2004.03.008