Next Article in Journal
Investigating the Catalytic Deactivation of a Pd Catalyst during the Continuous Hydrogenation of CO2 into Formate Using a Trickle-Bed Reactor
Previous Article in Journal
A Closed-Loop Biorefinery Approach for the Valorization of Winery Waste: The Production of Iron-Sulfonated Magnetic Biochar Catalysts and 5-Hydroxymethyl Furfural from Grape Pomace and Stalks
Previous Article in Special Issue
Heterogeneous Catalysts for Carbon Dioxide Methanation: A View on Catalytic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Perovskite-Derivative Ni-Based Catalysts for Hydrogen Production via Steam Reforming of Long-Chain Hydrocarbon Fuel

State Key Laboratory of Metastable Materials Science and Technology (MMST), Hebei Key Laboratory of Applied Chemistry, Yanshan University, Qinhuangdao 066004, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2024, 14(3), 186; https://doi.org/10.3390/catal14030186
Submission received: 21 February 2024 / Revised: 4 March 2024 / Accepted: 7 March 2024 / Published: 8 March 2024

Abstract

:
Large-scale hydrogen production by the steam reforming of long-chain hydrocarbon fuel is highly desirable for fuel-cell application. In this work, LaNiO3 perovskite materials doped with different rare earth elements (Ce, Pr, Tb and Sm) were prepared by a sol-gel method, and the derivatives supported Ni-based catalysts which were successfully synthesized by hydrogen reduction. The physicochemical properties of the as-prepared catalysts were characterized by powder X-ray diffraction, high-resolution transmission electron microscopy, N2 adsorption–desorption isotherms, H2 temperature-programmed reduction, and X-ray photoelectron spectroscopy. The catalytic performance of the as-prepared catalysts for hydrogen production was investigated via the steam reforming of n-dodecane. The results showed that the catalyst forms perovskite oxides after calcination with abundant mesopores and macropores. After reduction, Ni particles were uniformly distributed on perovskite derivatives, and can effectively reduce the particles’ sizes by doping with rare earth elements (Ce, Pr, Tb and Sm). Compared with the un-doped catalyst, the activity and hydrogen-production rate of the catalysts are greatly improved with rare earth element (Ce, Pr, Tb and Sm)-doped catalysts, as well as the anti-carbon deposition performance. This is due to the strong interaction between the uniformly distributed Ni particles and the support, as well as the abundant oxygen defects on the catalyst surface.

Graphical Abstract

1. Introduction

A fuel cell is a device that converts the chemical energy of hydrogen fuel into electrical energy, which has the characteristics of no combustion or Carnot cycles, a high energy conversion efficiency, low noise levels and zero emission [1,2]. It is regarded as a clean and efficient energy source for the 21st century. Therefore, the development of hydrogen energy has attracted great attention from researchers from all over the world, among which, hydrogen production via the steam reforming of fuel is a very practical method [3]. About 70% of the hydrogen in industry comes from hydrogen production via methane steam reforming [4]. Because methane is a gas at atmospheric pressure, and its transportation and compression costs are relatively high, hydrogen production via methane steam reforming is not suitable for distributed or mobile small-scale automobiles and ships. Therefore, hydrogen production via methane steam reforming technology is more suitable for hydrogen production in fixed areas with large methane reserves, and is not suitable for field and mobile hydrogen production. Diesel and kerosene (long-chain hydrocarbon fuels) have high hydrogen contents per unit volume and are present in the form of a liquid, which is convenient for transportation and storage [5,6,7]. So, the reforming hydrogen-production technology of liquid fuels such as diesel and kerosene are particularly suitable for hydrogen production in outdoor automobiles, ships and military submarines [8]. The development of hydrogen production technology for the steam reforming of liquid hydrocarbon fuels has significant implications for both mobile civilian and military equipment [9].
The steam-reforming process of liquid hydrocarbon fuels is relatively complex [10], with the main reaction being the reaction of fuel hydrocarbon molecules with steam, as shown in Equation (1), to form CO and H2. In addition, there are also water–gas shift reactions (such as Equation (2)) and some reactions of carbon deposition (Equations (3)–(5)).
CnH2n+2 + nH2O = nCO + (2n+1)H2
CO + H2O = CO2 + H2
CnHm → C2H4/C3H6→C
CH4 → 2H2 + C
2CO → CO2 + C
At present, the development of highly active and stable catalysts to improve the efficiency of hydrogen production while inhibiting carbon deposition is the key to achieving this process [11,12]. Among them, noble metal-based catalysts such as Ru, Pt and Rh have attracted the attention of scientists due to their high activity and stability [13,14,15,16,17,18,19]. However, because the relative scarcity and high prices of noble metals, large-scale use will inevitably increase costs. Therefore, the development of non-noble metal-based catalysts for fuel steam reforming has received increasing attention in recent years [20,21]. Ni-based catalysts have become a hot research topic due to their high activity and relatively low operating costs [22]. But, the biggest drawback of Ni-based catalysts currently is their relatively low selectivity for reforming, which can easily be accompanied by side reactions such as carbon deposition. In addition, due to the relatively low Taman temperature of Ni metal (591 °C), and that the temperature of the reforming reaction is generally higher than its Taman temperature, the active component Ni is prone to sintering and agglomeration, greatly reducing the service life of the catalyst [23,24,25].
There are various common strategies to improve catalyst stability, such as reducing the acidic sites of the support [26], enhancing metal–support interaction and constructing surface oxygen defects [27,28]. Masanori et al. [29] studied the use of La2O3-modified Ni/Al2O3 catalysts for kerosene steam reforming and found that the activity, stability and hydrogen yield of the catalysts were increased. This is attributed to the fact that La2O3 modification enhances the dispersion of Ni and reduces the acidic sites of the support, thereby reducing carbon deposition. Our group [30,31,32,33] studied the use of ceria-modified alumina or silica-supported Ni-Co bimetallic catalysts for the steam reforming of n-dodecane. It was found that ceria modification increased the dispersion of Ni-Co, enhanced the interaction between metal–support, and reduced the acidic sites on the support surface, thereby increasing the activity and stability of the catalyst. In recent years, the activity and stability of catalysts have been greatly improved by regulating the oxygen vacancies and metal–support interactions on the support [34,35,36]. Owing to the extremely important role of the support in the catalyst system, it can not only promote the dispersion of active components, but also suppress the sintering of active metals when there is a strong interaction between the support and metal. Among them, perovskite (ABO3) supports are a widely used type of support [37,38,39]. Its A and B sites can be finely regulated and substituted, and it has very high sintering resistance, showing good application prospects in many fields [40]. More importantly, perovskite oxides have abundant surface oxygen vacancies, which facilitate the activation of reactant water molecules and accelerate the elimination of carbon-deposition intermediates, thus enhancing the activity and anti-carbon-deposition performance of catalysts [41,42]. However, how to finely regulate the metal sites and oxygen-defect sites of perovskites remains a big challenge at present.
Based on this, LaNiO3 perovskite materials doped with different rare earth elements (Ce, Pr, Tb and Sm) were synthesized by the sol-gel method, and perovskite-derivative Ni-based catalysts were successfully synthesized by hydrogen reduction. The as-prepared catalysts were systematically analyzed and characterized using wide-angle X-ray diffraction, high-resolution transmission electron microscopy, temperature-programmed reduction, N2 adsorption–desorption, and X-ray transmission photoelectron spectroscopy. A study was conducted on the catalytic hydrogen production via the steam reforming of n-dodecane. The results showed that by controlling the doping of rare earth elements (Ce, Pr, Tb, and Sm), the size of nickel metal nanoparticles and the concentration of oxygen defects on the catalyst surface can be controlled, thereby exhibiting a different conversion of n-dodecane and hydrogen production rates. Moreover, doping with rare earth elements (Ce, Pr, Tb, and Sm) significantly improves the activity and anti-carbon-deposition performance of the catalyst.

2. Experimental Section

2.1. Chemicals and Materials

Nickel nitrate hexahydrate (Ni(NO3)2∙6H2O, Tianjin Guangfu Fine Chemical Research Institute, AR), lanthanum nitrate hexahydrate (La(NO3)3·6H2O, Tianjin Guangfu Fine Chemical Research Institute, AR), cerium nitrate hexahydrate (Ce(NO3)3∙6H2O), samarium nitrate hexahydrate (Sm(NO3)3∙6H2O), protactinium nitrate hexahydrate (Pr(NO3)3∙6H2O), terbium nitrate hexahydrate (Tb(NO3)3∙6H2O, Huawei Ruike, AR, Beijing, China), dodecane (C12H26, Tianjin China Kemio Chemical Reagent Co., Ltd., AR, Tijanjin, China), citric acid (CA) and ethylene glycol (EG) were purchased from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China, as were nitrogen/helium (high-purity, hexagonal gas company) and deionized water.

2.2. Preparation of Catalysts

The catalysts were prepared by sol-gel method [41]. Taking LaNiO3 synthesis as an example, firstly, 4.33 g (10 mmol) La(NO3)3·6H2O and 2.91 g (10 mmol) Ni(NO3)2·6H2O were added to 100 mL deionized water and stirred evenly at room temperature. Then, we added 4.20 g (20 mmol) of citric acid and 5.0 mL of ethylene glycol and stirred for 4 h until the mixture was uniform. We transferred the mixed solution to a heater and heated it at a heating temperature of 120 °C. The aqueous solution was removed by evaporation to obtain gel. We transferred the obtained gel to an oven for drying for 12 h at a temperature of 120 °C. After drying, a solid substance was obtained. The solid substance was ground into a solid powder with a mortar, and then calcined at a muffle furnace with 800 °C for 3 h to obtain LaNiO3. Then, we placed the calcined LaNiO3 in an atmosphere furnace, and reduced it at 650 °C under a 5% Vol H2/Ar mixture for 1 h to prepare the reduced catalyst. Catalysts doped with different rare earth elements were synthesized by using the same method, where the atomic ratio of La:Ce:Ni in LaCeNiO3 was 0.8:0.2:1, the atomic ratio of La:Ce:Pr:Ni in LaCePrNiO3 was 0.8:0.1:0.1:1, the atomic ratio of La:Ce:Sm:Ni in LaCeSmNiO3 was 0.8:0.1:0.1:1 and the atomic ratio of La:Ce:Tb:Ni in LaCeTbNiO3 was 0.8:0.1:0.1:1.

2.3. Characterization

The phase of structure analysis for the prepared catalyst, precursor and reduced catalyst was carried out using Rigaku D8-Focus X-ray diffractometer from Nippon Institute of Science Saitama, Japan. The testing conditions for a wide-angle X-ray diffractometer are Cu target, K α Ray (λ = 0.154 nm), tube voltage of 40 kV, tube current of 200 mA, scanning speed of 8 º/min, and scanning range of 10° to 80°. The morphology and nanoparticles of the catalyst were observed using the Tecnai G2 F20 transmission electron microscopy (TEM) from FEI in Eindhoven, The Netherlands, with an operating voltage of 150–300 kV. The distribution of elements was analyzed using surface scanning. The pore structure of the material was analyzed using Micromeritics Tristar 3000 (Norcross, GA, USA), with parameters of nitrogen atmosphere at −196 °C. Before sample testing, degassing and dehydration were required at 300 °C. Temperature-programmed reduction was tested on the AMI-300 chemical-adsorption instrument (Tuggerah, Australia). Firstly, we weighed a certain mass of a powder sample and pretreated it in an Ar atmosphere at 400 °C for 1 h. After cooling to 50 °C, a 10% Vol H2–Ar mixture gas was introduced, and the temperature was raised from 50 °C to 800 °C at a heating rate of 10 °C/min. The carbon deposition of the catalyst after the reaction was characterized using a TQ-500 thermal analyzer (Tokyo, Japan). Approximately 10 mg of the sample was weighed and placed in a ceramic crucible. The temperature was raised from room temperature to 800 °C in an air atmosphere at a heating rate of 10 °C/min to obtain the weight loss curve of the sample. The DTG data can be obtained by first-order differentiation of the weightlessness curve. Raman spectra were measured using a DXR Microscope system (Waltham, MA, USA). The sample (50 mg) was loaded and excited under an argon laser (532 nm).

2.4. Catalytic Performance Evaluation

The hydrogen production via the steam reforming of n-dodecane was carried out in a stainless-steel tubular fixed bed reactor, as shown in Figure S1. Before the reaction, the catalyst was compressed to a pressure of 10 MPa, and then sieved to obtain a catalyst with a mesh size of 20–40. We mixed 0.25 g of catalyst with 1.2 g of quartz sand and filled it evenly in a stainless-steel tube with a diameter of 6 mm. We sealed it, then introduced nitrogen gas and raised the temperature to 650 °C. It reduced for 1 h in a 5% Vol H2/Ar mixture, then was switched to nitrogen to stabilize the system temperature to reaction temperature (600 °C) at atmospheric pressure. We input a constant amount of water using a high-pressure metering pump and gasified it at 300 °C for 15 min. Then, liquid n-dodecane was introduced, with a feed rate of 15 mL/gcat∙h. After gasification at 300 °C and mixing with steam, it was uniformly introduced into the catalyst bed. After the reaction, the product underwent condensation and drying. Then, the composition and content of the gas phase products were analyzed online using portable micro-gas chromatography (Micro GC-490, Agilent, Santa Clara, CA, USA). The detector for chromatography was a TCD detector, equipped with three chromatography columns. Firstly, activated alumina columns were used to detect hydrocarbon molecules of C3 and above. Secondly, PPU columns were used to detect CO2, ethane and ethylene. Finally, 5 Å MS columns were used to detect H2, N2, CH4 and CO. We calculated the conversion rate of n-dodecane based on the analysis results of gas-phase products (Equation (6)). We calculated the yield and composition of gas-phase products based on the analysis results (Equations (7) and (8)).
X 12   ( wt % ) = F CO out + F CO 2 out + F CH 4 out 12 F c 12 in × 100
Gas   formation   rate   ( mmol / min )   R H 2 = A H 2 · f H 2 A N 2 · f N 2 × R N 2
V % H 2 = R ( H 2 ) R ( H 2 + CO + CO 2 + CH 4 ) × 100
In Equation (6), the conversion of X12—n-dodecane, %. R(N2) = 60 mL/min, F C 12 in —in total amount of carbon entering the reactor, mmol/min. F CO out , F CO 2 out , F CH 4 out —the total number of C1 products in the product leaving the reactor, mmol/min. In Equation (7), R(N2) = 60 mL/min, which is the nitrogen flow rate. AH2 and fH2 represent the chromatographic peak area and response factor of H2, respectively. AN2 and fN2 represent the peak area and response factor of N2 obtained in the chromatography, respectively. In Equation (8), V%(H2)—the percentage content of hydrogen gas in dry gas (excluding nitrogen).

3. Results and Discussion

3.1. Structural and Morphological Characteristics of Catalysts

The XRD patterns of the catalysts after calcination are shown in Figure 1a. After calcination, the catalysts mainly exist in the crystal structure of LaNiO3 (PDF-34-1181) perovskite [37]. With the doping of rare earth elements, the diffraction peak intensity of the crystal structure of perovskite significantly decreases, which may be due to the heteroatom doping hindering crystal growth, resulting in a decrease in the crystallinity of perovskite [42]. In addition, the La2O3 structure appeared after doping with rare earth elements, and without the structure of doped oxide and nickel oxide alone. This indicates that rare earth elements can partially replace La elements and enter the interior of perovskite crystals and are highly dispersed in perovskite oxides.
The XRD patterns of the reduced catalysts are shown in Figure 1b. Due to the reduction temperature of 650 °C, perovskite is easily reduced to Ni metal and La2O3 [37]. Therefore, new diffraction peaks dominated by La2O3 appeared in the reduced catalysts, and a characteristic peak of Ni metal appeared at 44.5 °C [43]. For LaCeNiO3 catalyst, in addition to Ni and La2O3, there are also characteristic peaks of CeO2. The CeO2 peak that appeared after reduction only for the LaCeNiO3 catalyst may be attributed to the content of Ce in LaCeNiO3 being higher than that of other catalysts. The above results indicate that perovskite is transformed into oxide after reduction, which can effectively load Ni metal.
The TEM results showed that the surface of the reduced perovskite-derived material is rich in nickel nanoparticles, and some pores can be observed (Figure 2a,d and Figure S2a). HRTEM showed that the lattice spacing of nickel nanoparticles was 0.20 nm, and La2O3 with a lattice spacing of 0.29 nm was observed (Figure 2b,e and Figure S2b,d), indicating that LaNiO3 perovskite was transformed into Ni metal and La2O3 after hydrogen reduction [44]. In addition, elemental surface mapping showed that particles size of nickel is approximately 25 nm (Figure 2c). After doping with rare earth elements Ce (Figure S2e) and Pr (Figure S2f), the particle size of nickel decreases to 20 nm (Figure 2f), indicating that rare earth doping can reduce the particle size of nickel [45]. The above results suggested that perovskite-derivative Ni-based catalysts can be successfully constructed.
The N2 adsorption–desorption isotherms of the catalysts after calcination are shown in Figure 3a. The hysteresis loop of the adsorption–desorption isotherms can be clearly observed from Figure 3a, indicating that the as-prepared catalysts are mesoporous materials. In addition, as the relative pressure increases, the adsorption capacity also increases, indicating the presence of a certain amount of macropores [46]. The pore-size distribution curve is shown in Figure 3b, and the prepared materials exhibit a dual pore-size distribution. Among them, the mesoporous pore size is around 36 nm, and the macroporous pore size is around 60 nm. The specific surface area, pore volume and pore size data of the catalysts after calcination are summarized in Table 1. In Table 1, the specific surface area of the catalysts is relatively small, and the pore size is relatively large, with an average pore size of 14–23 nm.
The H2 temperature-programmed reduction of the catalysts is shown in Figure 4. There are three reduction peaks distributed in the low temperature, medium temperature and high temperature regions (Figure 4). The reduction peaks belong to the reduction of surface oxygen, LaNiO3 to LaNiO2.5, and LaNiO2.5 to Ni metal, respectively [43,44]. The shoulder peak between 300 and 350 °C belongs to the reduction of surface oxygen, as confirmed by previous work [39]. The peak intensity increases with rare earth-element doping, suggesting increased surface oxygen vacancy. It was found that after doping with rare earth elements, the reduction temperature of the catalysts shifted towards the high-temperature region, indicating an enhanced interaction strength between NiO and the support. The two main reduction peaks of LaNiO3 are at 350 °C and 500 °C. After doping with rare earth elements, these two reduction peaks both shift towards the high-temperature region. The two main reduction peaks of LaCePrNiO3 are concentrated at temperatures of 425 °C and 513 °C, and the peak areas in the high-temperature region are relatively large, indicating that the catalyst has stronger metal–support interactions. Previous research has shown that stronger metal–support interactions contribute to the dispersion and sintering resistance of reduced metals [47,48]. As a result, the catalyst may exhibit good activity, stability and anti-carbon deposition.
Figure 5 shows the XPS spectra of La 3d in the samples. The peaks of 835 eV, 838.55 eV and 852.03 eV, 855.39 eV correspond to La 3d5/2 and La 3d3/2 in LaNiO3, respectively (Table 2). After doping with rare earth elements, the electron-binding energy of La shifts towards the direction of low binding energy, indicating that La has obtained electrons. The Ce-3d and Pr-3d are showed in Figure S3a and S3b, respectively. Four peaks can be observed in the O 1s spectra of LaNiO3, LaCeNiO3 and LaCePrNiO3 (Figure 5). They belong to lattice oxygen (OL), oxygen vacancies (OV), surface oxygen (Osurf), and adventitious oxygen (Oadv), respectively [42]. According to XPS peak-fitting calculations, the proportion of lattice oxygen in the sample has increased. However, the catalyst surface is mainly composed of oxygen vacancies, surface oxygen, and adventitious oxygen, accounting for more than 75.2% (Table 3). In addition, the XPS results of Ni 2p were observed. The three peaks at 852 eV, 855.5 eV and 865 eV belong to Ni0, Ni2+ and satellite peaks (Ni2+), respectively. The binding energies of Ni0 in LaNiO3, LaCeNiO3 and LaCePrNiO3 are 852 eV, 852.1 eV and 852.16 eV, indicating that the electron-binding energy of Ni0 in the catalysts doped with rare earth elements shifts towards lower binding energies. This indicates that the increase in electron density of Ni0 may be explained by the enhanced strength of the metal–support interaction, which is consistent with the results of H2-TPR. Moreover, previous studies have indicated that the electron density of nickel enhances the activation and dissociation of C–H bonds [49]. The ratio of Ni0/(Ni0 + Ni2+) was also calculated, and the Ni0 ratios for LaNiO3, LaCeNiO3 and LaCePrNiO3 were about 54.1%, 55.6% and 56.8%, respectively. This result suggests that the proportion of nickel metal in the three catalysts is similar, excluding the influence of different nickel metal contents on catalytic performance.

3.2. Catalytic Performance for H2 Production

Figure 6 shows the conversion rate and gas production rate of n-dodecane steam reforming catalyzed by the catalysts. As shown in Figure 6a, the initial conversion of n-dodecane on LaNiO3 catalyst was 53.5%, and the yield of hydrogen was 4.99 mmol/min. After 150 min of reaction, the conversion rate of n-dodecane decreased to 43.2%, and the yield of hydrogen was 3.87 mmol/min. The conversion of n-dodecane and the yield of hydrogen decreased by 19.2% and 22.44%, respectively. After the adding of Ce, the stability of the catalysts was significantly improved, with a conversion rate of 51.3% for n-dodecane and a hydrogen yield of 5.35 mmol/min for LaCeNiO3. And, the conversion rates of n-dodecane on LaCePrNiO3, LaCeTbNiO3 and LaCeSmNiO3 catalysts were 70%, 55% and 65% (Figure 6f), respectively. Especially for the LaCePrNiO3 catalyst, the conversion was as high as 70%, and the yield of hydrogen reached 5.9 mmol/min. After 300 min of reaction, there was no significant decrease in the conversion and hydrogen-production rate. This result indicates that the activity, stability, and hydrogen-production rate of the catalyst were significantly improved after doping with rare earth elements.
The composition of gas-phase products is shown in Figure 7, where the composition of H2 in dry gas was about 70%, and the proportion of CO and CO2 reached about 10–20% [15]. This result is consistent with the previously reported results of hydrogen production via the steam reforming of n-dodecane. The proportion of CO in LaNiO3 gradually decreases, indicating that its catalytic activity is also gradually decreasing, which is consistent with the results in Figure 6a. The catalyst doped with rare earth elements showed almost no significant change in gas-phase composition, indicating the good stability of the catalyst.
We also tested the performance of the catalysts under the condition of H2O:C12H26 mole ratio of 48, as shown in Figure 8a,b. The initial conversion rate of n-dodecane using the LaNiO3 catalyst was 58%, and the yield of hydrogen gas was 5.1 mmol/min. After 270 min of reaction, the conversion rate increased to 70% and the hydrogen production rate increased to 5.88 mmol/min, indicating that increasing the amount of water in the reaction can greatly improve the conversion, hydrogen production rate and stability. The LaCePrNiO3 catalyst has a conversion rate of up to 88% and a hydrogen gas yield of 6.6 mmol/min. After 300 min of reaction, there was no significant decrease in conversion rate and hydrogen-production rate. This result indicates that the activity, stability and hydrogen-production rate of the catalyst have been significantly improved after doping with rare earth elements.

3.3. Analysis of Carbon Deposition on Catalyst after Reaction

The carbon deposition on the surface of the catalysts after the reaction was analyzed by using thermogravimetric techniques, and the results are shown in Figure 9 (solid line). The weight-loss rate of the LaNiO3 catalyst was 33%, indicating the presence of a large amount of carbon deposition on the surface of the catalyst after the reaction [50]. After the introduction of rare earth element Ce, the weight-loss rate of the LaCeNiO3 catalyst was significantly reduced. Its weight-loss rate was 20%, indicating that the introduction of rare earth Ce doping reduces carbon deposition. Further introduction of other rare earth elements reduces the weight-loss rate of the catalyst to a certain extent, indicating a further reduction in carbon deposition. Especially when Ce and Pr were co-doped, the weight-loss rate was only 10%. The above results indicate that the introduction of rare earth dopants significantly reduces the carbon deposition of the catalyst, suggesting a significant improvement in the catalyst’s anti-carbon deposition performance. This can be attributed to the points as follows: firstly, the addition of rare earth elements leads to a stronger metal–support interaction, resulting in smaller nanoparticles size of nickel metal. The presence of smaller nickel nanoparticles helps to suppress the formation of carbon deposition. Secondly, the addition of rare earth elements results in abundant oxygen defects on the surface of the catalyst (H2-TPR and O 1s XPS), which helps to eliminate carbon deposition.
The differential thermal analysis of the catalyst after the reaction is shown by the dashed line in Figure 9 (dashed line). The temperature for carbon deposition elimination was concentrated at 610 °C for LaNiO3 catalyst, and its peak intensity was the highest, indicating the highest amount of carbon deposition. After the introduction of rare earth-element doping, the peak intensity of carbon deposition significantly decreases, and the elimination temperature of carbon deposition moved towards the low-temperature range [51]. The temperature for the LaCePrNiO3 catalyst eliminates carbon deposition was 596 °C. This result indicates that introducing rare earth-element doping makes it easier to eliminate carbon deposition. Therefore, catalysts with low carbon deposition display good catalytic performance, as shown in Figure 6 and Figure 8.
In addition, the SEM images of the spent catalysts of LaNiO3 and LaCePrNiO3 are shown in Figure S4. Obviously, LaNiO3 possessed much more carbon deposition than that of LaCePrNiO3, suggesting that rare earth-element doping can effectively eliminate carbon deposition, which was also verified by the above TG/DTG result.
The Raman spectra of the catalysts after the reaction is shown in Figure 10. The range of 1000 cm−1 to 1800 cm−1 is the characteristic Raman peak of carbon deposition. At 1350 cm−1, it represents an amorphous carbon abbreviation (D band) with many defects [52]. And at 1598 cm−1, it represents a planar C-C (sp2) bond-stretching vibration, which is a graphitized carbon (G band) [53]. The ID and IG of the LaNiO3 catalyst have very strong peak intensities, indicating a severe carbon deposition on the catalyst surface, with ID: IG = 1.51. The intensity of ID and IG in rare earth Ce-doped catalysts is significantly reduced, especially in Ce- and Pr-co-doped catalysts (LaCePrNiO3). The intensity of its peak is weak, and ID: IG = 1.23. The decrease in this ratio also indicates that the elimination of carbon deposition is easier. This result is consistent with the TG and DTG analysis results mentioned above, indicating that the catalyst doped with rare earth elements has abundant oxygen vacancies, and the presence of oxygen vacancies is very effective in eliminating carbon deposition. Therefore, the catalyst doped with rare earth elements significantly enhances its ability for anti-carbon deposition.

4. Conclusions

In a word, LaNiO3 perovskite materials doped with different rare earth elements (Ce, Pr, Tb and Sm) were prepared by a sol-gel method, and perovskite-derivative-supported Ni-based catalysts were successfully synthesized by hydrogen reduction. The physical and chemical properties of the prepared catalysts were characterized using various techniques. The results showed that the calcined catalyst existed in a perovskite oxide structure, with abundant mesopores and macropores. By controlling the doping of rare earth elements (Ce, Pr, Tb and Sm), the strength of the metal–support interaction, the size of the nickel metal particles, and the ratio of the lattice oxygen to surface oxygen on the catalyst surface can be controlled. After reduction, Ni metal nanoparticles are uniformly distributed on perovskite derivatives, and doping with rare earth elements (Ce, Pr, Tb and Sm) can effectively reduce the particle size of Ni metal. The prepared catalysts were used for hydrogen production via the steam reforming of n-dodecane. The results showed that compared to the un-doped catalyst, the activity and hydrogen production rate of the catalysts doped with rare earth elements (Ce, Pr, Tb and Sm) were significantly improved. The LaCePrNiO3 catalyst has the highest activity. At 600 °C, H2O/C = 4 and the liquid hour space velocity of 15 mL/gcat∙h, the conversion rate of n-dodecane reached 88%, and the hydrogen-production rate reached 6500 umol/min. After 270 min of reaction, the catalyst performance did not significantly decrease. The catalyst doped with rare earth elements significantly enhanced its ability for anti-carbon deposition, as verified by TG/DTG and Raman spectra. The reason for the high catalytic performance of LaCePrNiO3 was attributed to the strong metal–support interaction between uniformly distributed Ni particles and the support, as well as the rich surface oxygen concentration of the catalyst. This work provides a simple method for preparing Ni-based catalysts with high activity and anti-carbon deposition performance for hydrogen production via the steam reforming of fuel.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14030186/s1, Figure S1. The schematic diagram of experimental apparatus; Figure S2. (a) TEM images and (b) HRTEM images of LaNiO3. (c) TEM images and (d) HRTEM images of LaCePrNiO3, (e) The element mapping of Ce and (f) Pr of LaCePrNiO3 after hydrogen reduction; Figure S3. XPS spectra of the catalysts after reduction for (a) Ce-3d, and (b) Pr-3d.; Figure S4. SEM images of spent catalysts for (a) LaNiO3, and (b) LaCePrNiO3. Reaction condition: 600 °C, 0.1 MPa, LHSV of C12H26 15 mL/gcat∙h, with mole ratio of H2O:C12H26 at 24.

Author Contributions

Conceptualization, K.G. and Z.X.; methodology, K.G.; software, H.Z.; validation, K.G., H.Z. and Z.X.; formal analysis, C.Z., X.G. and H.L.; investigation, Z.X.; resources, K.G.; data curation, K.G., H.Z. and Z.X.; writing—original draft preparation, K.G., H.Z. and Z.X.; writing—review and editing, K.G., H.Z. and Z.X.; supervision, K.G.; project administration, Z.X.; funding acquisition, K.G. and Z.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Science and Technology Project of Hebei Education Department (No. QN2023094), the Key Research and Design Program of Qinhuangdao (No. 202101A005), the Cultivation Project for Basic Research and Innovation of Yanshan University (No. 2021LGQN028), the Project for Research and Development of Metal Catalysts for Photo-thermal Decomposition of Waste Plastics to Prepare Value-added Chemicals (No. x2023322), and the Subsidy for Hebei Key Laboratory of Applied Chemistry after Operation Performance (No. 22567616H).

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Materials, further inquiries can be directed to the corresponding author/s.

Acknowledgments

The authors gratefully acknowledge all of their co-workers for their valuable research contributions that are summarized in the present review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kreuer, K.D.; Paddison, S.J.; Spohr, E.; Schuster, M. Transport in proton conductors for fuel-cell applications: Simulations, elementary reactions, and phenomenology. Chem. Rev. 2004, 104, 4637–4678. [Google Scholar] [CrossRef]
  2. Su, H.R.; Hu, Y.H. Progress in low-temperature solid oxide fuel cells with hydrocarbon fuels. Chem. Eng. J. 2020, 402, 126235. [Google Scholar] [CrossRef]
  3. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. [Google Scholar] [CrossRef]
  4. Park, H.G.; Han, S.Y.; Jun, K.W.; Woo, Y.; Park, M.J.; Kim, S.K. Bench-scale steam reforming of methane for hydrogen production. Catalysts 2019, 9, 615. [Google Scholar] [CrossRef]
  5. Zhan, Z.; Barnett, S.A. An octane-fueled solid oxide fuel cell. Science 2005, 308, 844–848. [Google Scholar] [CrossRef]
  6. Xiao, Z.; Zhang, C.; Li, P.; Wang, D.; Zhang, X.; Wang, L.; Zou, J.-J.; Li, G. Engineering oxygen vacancies on Tb-doped ceria supported Pt catalyst for hydrogen production through steam reforming of long-chain hydrocarbon fuels. Chin. J. Chem. Eng. 2024, in press. [Google Scholar] [CrossRef]
  7. Do, L.T.; Nguyen-Phu, H.; Pham, N.N.; Jeong, D.H.; Shin, E.W. Highly dispersed nickel nanoparticles on hierarchically ordered macroporous Al2O3 and its catalytic performance for steam reforming of 1-methyl naphthalene. Catalysts 2022, 12, 1542. [Google Scholar] [CrossRef]
  8. Xie, C.; Chen, Y.S.; Engelhard, M.H.; Song, C.S. Comparative study on the sulfur tolerance and carbon resistance of supported noble metal catalysts in steam reforming of liquid hydrocarbon fuel. ACS Catal. 2012, 2, 1127–1137. [Google Scholar] [CrossRef]
  9. Zheng, Q.C.; Xiao, Z.R.; Xu, J.S.; Pan, L.; Zhang, X.W.; Zou, J.J. Catalytic steam reforming and heat sink of high-energy-density fuels: Correlation of reaction behaviors with molecular structures. Fuel 2021, 286, 119371. [Google Scholar] [CrossRef]
  10. Xue, Q.Q.; Li, Z.W.; Chen, M.; Wang, Y.J.; Yan, B.H.; Luo, G.S. Co-precipitation continuous synthesis of the Ni-Rh-Ce0.75Zr0.25O2-δ catalyst in the membrane dispersion microreactor system for n-dodecane steam reforming to hydrogen. Fuel 2021, 297, 120785. [Google Scholar] [CrossRef]
  11. Li, D.; Tamura, M.; Nakagawa, Y.; Tomishige, K. Metal catalysts for steam reforming of tar derived from the gasification of lignocellulosic biomass. Bioresour. Technol. 2015, 178, 53–64. [Google Scholar] [CrossRef]
  12. Li, S.; Gong, J. Strategies for improving the performance and stability of Ni-based catalysts for reforming reactions. Chem. Soc. Rev. 2014, 43, 7245–7256. [Google Scholar] [CrossRef] [PubMed]
  13. Shoynkhorova, T.B.; Simonov, P.A.; Potemkin, D.I.; Snytnikov, P.V.; Belyaev, V.D.; Ishchenko, A.V.; Svintsitskiy, D.A.; Sobyanin, V.A. Highly dispersed Rh-, Pt-, Ru/Ce0.75Zr0.25O2-δ catalysts prepared by sorption-hydrolytic deposition for diesel fuel reforming to syngas. Appl. Catal. B—Environ. 2018, 237, 237–244. [Google Scholar] [CrossRef]
  14. Kim, D.H.; Kang, J.S.; Lee, Y.J.; Park, N.K.; Kim, Y.C.; Hong, S.I.; Moon, D.J. Steam reforming of n-hexadecane over noble metal-modified Ni-based catalysts. Catal. Today 2008, 136, 228–234. [Google Scholar] [CrossRef]
  15. Ming, Q.; Healey, T.; Allen, L.; Irving, P. Steam reforming of hydrocarbon fuels. Catal. Today 2002, 77, 51–64. [Google Scholar] [CrossRef]
  16. Li, L.; Zuo, S.W.; An, P.F.; Wu, H.Y.; Hou, F.; Li, G.Z.; Liu, G.Z. Hydrogen production via steam reforming of n-dodecane over NiPt alloy catalysts. Fuel 2020, 262, 116469. [Google Scholar] [CrossRef]
  17. Liu, J.; Xiao, Z.; Zhang, X. Pt and ceria promoted Ni/SBA-15 for n-dodecane steam reforming for hydrogen production. Chem. Ind. Eng. 2019, 36, 1–10. [Google Scholar]
  18. Miyamoto, M.; Arakawa, M.; Oumi, Y.; Uemiya, S. Influence of metal cation doping on Ru/CeO2/Al2O3 catalyst for steam reforming of desulfurized kerosene. Int. J. Hydrogen Energy 2015, 40, 2657–2662. [Google Scholar] [CrossRef]
  19. Wang, T.; Liu, S.; Wang, L.; Liu, G. High-performance Rh/CeO2 catalysts prepared by L-lysine-assisted deposition precipitation method for steam reforming of toluene. Fuel 2023, 341, 127736. [Google Scholar] [CrossRef]
  20. Li, Z.W.; Li, M.; Ashok, J.; Sibudjing, K. NiCo@NiCo phyllosilicate@CeO2 hollow core shell catalysts for steam reforming of toluene as biomass tar model compound. Energy Convers. Manag. 2019, 180, 822–830. [Google Scholar] [CrossRef]
  21. Zhang, H.C.; Xiao, Z.R.; Yang, M.; Zou, J.J.; Liu, G.Z.; Zhang, X.W. Highly dispersible cerium-oxide modified Ni/SBA-15 for steam reforming of bio-mass based JP10. Chin. J. Chem. Eng. 2022, 43, 255–265. [Google Scholar] [CrossRef]
  22. Tribalis, A.; Panagiotou, G.D.; Bourikas, K.; Sygellou, L.; Kennou, S.; Ladas, S.; Lycourghiotis, A.; Kordulis, C. Ni catalysts supported on modified alumina for diesel steam reforming. Catalysts 2016, 6, 11–23. [Google Scholar] [CrossRef]
  23. Meloni, E.; Martino, M.; Palma, V. A short review on Ni based catalysts and related engineering issues for methane steam reforming. Catalysts 2020, 10, 352. [Google Scholar] [CrossRef]
  24. Li, P.; Zhang, S.L.; Xiao, Z.R.; Zhang, H.; Ye, F.; Gu, J.M.; Wang, J.D.; Li, G.Z.; Wang, D.S. Ni-TiO2 catalysts derived from metal-organic framework for efficient photo-thermal CO2 methanation. Fuel 2024, 357, 129817. [Google Scholar] [CrossRef]
  25. Xiao, Z.; Li, P.; Zhang, H.; Zhang, S.; Tan, X.; Ye, F.; Gu, J.; Zou, J.-j.; Wang, D. A comprehensive review on photo-thermal co-catalytic reduction of CO2 to value-added chemicals. Fuel 2024, 362, 130906. [Google Scholar] [CrossRef]
  26. Kim, H.; Lee, Y.H.; Lee, H.; Seo, J.C.; Lee, K. Effect of Mg contents on catalytic activity and coke formation of mesoporous Ni/Mg-aluminate spinel catalyst for steam methane reforming. Catalysts 2020, 10, 828. [Google Scholar] [CrossRef]
  27. Xiao, Z.R.; Li, Y.T.; Hou, F.; Wu, C.; Pan, L.; Zou, J.J.; Wang, L.; Zhang, X.W.; Liu, G.Z.; Li, G.Z. Engineering oxygen vacancies and nickel dispersion on CeO2 by Pr doping for highly stable ethanol steam reforming. Appl. Catal. B—Environ. 2019, 258, 117940. [Google Scholar] [CrossRef]
  28. Xiao, Z.; Wu, C.; Wang, L.; Xu, J.; Zheng, Q.; Pan, L.; Zou, J.; Zhang, X.; Li, G. Boosting hydrogen production from steam reforming of ethanol over nickel by lanthanum doped ceria. Appl. Catal. B—Environ. 2021, 286, 119884. [Google Scholar] [CrossRef]
  29. Sugisawa, M.; Takanabe, K.; Harada, M.; Kubota, J.; Domen, K. Effects of La addition to Ni/Al2O3 catalysts on rates and carbon deposition during steam reforming of n-dodecane. Fuel Process. Technol. 2011, 92, 21–25. [Google Scholar] [CrossRef]
  30. Xiao, Z.R.; Wu, C.; Li, L.; Li, G.Z.; Liu, G.Z.; Wang, L. Pursuing complete and stable steam reforming of n-dodecane over nickel catalysts at low temperature and high LHSV. Int. J. Hydrogen Energy 2017, 42, 5606–5618. [Google Scholar] [CrossRef]
  31. Ye, F.; Qian, J.; Xia, J.J.; Li, L.F.; Wang, S.J.; Zeng, Z.X.; Miao, J.; Ahamad, M.; Xiao, Z.R.; Zhang, Q.R. Effect photoelectrocatalytic degradation of pollutants over hydrophobic carbon felt loaded with Fe-doped porous carbon nitride via direct activation of molecular oxygen. Environ. Res. 2014, 249, 118497. [Google Scholar] [CrossRef]
  32. Xiao, Z.R.; Li, L.; Wu, C.; Li, G.Z.; Liu, G.Z.; Wang, L. Ceria-promoted Ni–Co/Al2O3 catalysts for n-dodecane steam reforming. Catal. Lett. 2016, 146, 1780–1791. [Google Scholar] [CrossRef]
  33. Xiao, Z.R.; Zheng, Q.C.; Zhang, X.W.; Li, L.; Wang, L.; Li, G.Z. Synthesis of Ni-Co catalysts supported on flower-like MgAl composite oxide for hydrogen production by n-dodecane steam reforming. Chin. J. Inorg. Chem. 2021, 37, 629–637. [Google Scholar]
  34. Xiao, Z.R.; Ji, S.; Hou, F.; Li, Y.T.; Zhang, H.C.; Wang, L.; Zhang, X.W.; Liu, G.Z.; Zou, J.J.; Li, G.Z. n-Dodecane steam reforming catalyzed by Ni-Ce-Pr catalysts. Part 1: Catalyst preparation and Pr doping. Catal. Today 2018, 316, 78–90. [Google Scholar] [CrossRef]
  35. Xiao, Z.R.; Zhang, X.W.; Hou, F.; Wu, C.; Wang, L.; Li, G.Z. Tuning metal-support interaction and oxygen vacancies of ceria supported nickel catalysts by Tb doping for n-dodecane steam reforming. Appl. Surf. Sci. 2020, 503, 144319. [Google Scholar] [CrossRef]
  36. Xiao, Z.; Zhang, X.; Wang, L.; Li, G. Optimizing the preparation of Ni-Ce-Pr catalysts for efficient hydrogen production by n-dodecane steam reforming. Int. J. Energy Res. 2019, 44, 1828–1842. [Google Scholar] [CrossRef]
  37. Oemar, U.; Ang, P.S.; Hidajat, K.; Kawi, S. Promotional effect of Fe on perovskite LaNixFe1−xO3 catalyst for hydrogen production via steam reforming of toluene. Int. J. Hydrogen Energy 2013, 38, 5525–5534. [Google Scholar] [CrossRef]
  38. Chen, G.K.; Shen, Q.W.; Zhang, X.; Cai, Z.W.; Shao, Z.C.; Li, S.; Yang, G.G. Methanol steam reforming over La1-xSrxCeO3-δ catalysts for hydrogen production: Optimization of operating parameters. Catalysts 2023, 13, 248. [Google Scholar] [CrossRef]
  39. Wang, Y.; Qian, S.; Chen, Y.; Yan, B.; Cheng, Y. Effect of the calcination temperature of LaNiO3 on the structural properties and reaction performance of catalysts in the steam reforming of methane. Catalysts 2023, 13, 356. [Google Scholar] [CrossRef]
  40. Das, S.; Bhattar, S.; Liu, L.N.; Wang, Z.G.; Xi, S.B.; Spivey, J.J.; Kawi, S. Effect of partial Fe substitution in La0.9Sr0.1NiO3 perovskite-derived catalysts on the reaction mechanism of methane dry reforming. ACS Catal. 2020, 10, 12466–12486. [Google Scholar] [CrossRef]
  41. Pereñiguez, R.; Gonzalez-delaCruz, V.M.; Caballero, A.; Holgado, J.P. LaNiO3 as a precursor of Ni/La2O3 for CO2 reforming of CH4: Effect of the presence of an amorphous NiO phase. Appl. Catal. B—Environ. 2012, 123–124, 324–332. [Google Scholar] [CrossRef]
  42. Yu, J.B.; Muhetaer, A.; Gao, X.W.; Zhang, Z.Z.; Yang, Y.Y.; Li, Q.; Chen, L.X.; Liu, H.C.; Xu, D.S. Highly active hydrogen-rich photothermal reverse water gas shift reaction on Ni/LaInO3 perovskite catalysts with near-unity selectivity. Angew. Chem. Int. Ed. 2023, 62, e202303135. [Google Scholar] [CrossRef]
  43. Wu, G.W.; Li, S.R.; Zhang, C.X.; Wang, T.; Gong, J.L. Glycerol steam reforming over perovskite-derived nickel-based catalysts. Appl. Catal. B—Environ. 2014, 144, 277–285. [Google Scholar] [CrossRef]
  44. De Lima, S.M.; da Silva, A.M.; da Costa, L.O.O.; Assaf, J.M.; Jacobs, G.; Davis, B.H.; Mattos, L.V.; Noronha, F.B. Evaluation of the performance of Ni/La2O3 catalyst prepared from LaNiO3 perovskite-type oxides for the production of hydrogen through steam reforming and oxidative steam reforming of ethanol. Appl. Catal. A 2010, 377, 181–190. [Google Scholar] [CrossRef]
  45. Liu, H.; Yu, J. Catalytic performance of Cu-Ni/La0.75Sr0.25Cr0.5Mn0.5O3-δ for dry methane reforming. Int. J. Energy Res. 2022, 46, 10522–10534. [Google Scholar] [CrossRef]
  46. Bussi, J.; Musso, M.; Quevedo, A.; Faccio, R.; Romero, M. Structural and catalytic stability assessment of Ni-La-Sn ternary mixed oxides for hydrogen production by steam reforming of ethanol. Catal. Today 2017, 296, 154–162. [Google Scholar] [CrossRef]
  47. Kim, T.; Song, K.H.; Yoon, H.; Chung, J.S. Steam reforming of n-dodecane over K2Ti2O5-added Ni-alumina and Ni-zirconia (YSZ) catalysts. Int. J. Hydrogen Energy 2016, 41, 17922–17932. [Google Scholar] [CrossRef]
  48. Li, D.L.; Koike, M.; Chen, J.H.; Nakagawa, Y.; Tomishige, K. Preparation of Ni-Cu/Mg/Al catalysts from hydrotalcite-like compounds for hydrogen production by steam reforming of biomass tar. Int. J. Hydrogen Energy 2014, 39, 10959–10970. [Google Scholar] [CrossRef]
  49. Jing, J.Y.; Wei, Z.H.; Zhang, Y.B.; Bai, H.C.; Li, W.Y. Carbon dioxide reforming of methane over MgO-promoted Ni/SiO2 catalysts with tunable Ni particle size. Catal. Today 2020, 356, 589–596. [Google Scholar] [CrossRef]
  50. Xiao, Z.R.; Li, P.; Zhang, S.L.; Gu, J.M.; Wang, D.S. Effects of metal-support interactions and interfaces on catalytic performance over M2O3- (M = La, Al-) supported Ni catalysts. Int. J. Energy Res. 2023, 2023, 6504914. [Google Scholar] [CrossRef]
  51. Li, D.; Zeng, L.; Li, X.Y.; Wang, X.; Ma, H.Y.; Assabumrungrat, S.; Gong, J.L. Ceria-promoted Ni/SBA-15 catalysts for ethanol steam reforming with enhanced activity and resistance to deactivation. Appl. Catal. B—Environ. 2015, 176, 532–541. [Google Scholar] [CrossRef]
  52. Chen, M.Q.; Wang, Y.S.; Yang, Z.L.; Liang, T.; Liu, S.M.; Zhou, Z.S.; Li, X.J. Effect of Mg-modified mesoporous Ni/Attapulgite catalysts on catalytic performance and resistance to carbon deposition for ethanol steam reforming. Fuel 2018, 220, 32–46. [Google Scholar] [CrossRef]
  53. Xue, Q.Q.; Li, Z.W.; Yi, H.L.; Jiang, Z.; Yan, B.H.; Wang, Y.J.; Luo, G.S. Controlling sintering and carbon deposition with post-coated MgO confined group VIII metal-based catalysts towards durable high-temperature steam reforming. Appl. Catal. B—Environ. 2022, 318, 121874. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the catalysts that after calcination (a) and (b) catalysts after hydrogen reduction. The symbol of Δ, ♥, ▲, ♠ and ♣ represented NiO, La2O3, LaNiO3 perovskite, CeO2 and Ni, respectively.
Figure 1. XRD patterns of the catalysts that after calcination (a) and (b) catalysts after hydrogen reduction. The symbol of Δ, ♥, ▲, ♠ and ♣ represented NiO, La2O3, LaNiO3 perovskite, CeO2 and Ni, respectively.
Catalysts 14 00186 g001
Figure 2. (a) TEM images and (b) HRTEM images of LaNiO3, (c) the element mapping of LaNiO3 after hydrogen reduction. (d) TEM images and (e) HRTEM images of LaCePrNiO3, (f) the element mapping of LaCePrNiO3 after hydrogen reduction.
Figure 2. (a) TEM images and (b) HRTEM images of LaNiO3, (c) the element mapping of LaNiO3 after hydrogen reduction. (d) TEM images and (e) HRTEM images of LaCePrNiO3, (f) the element mapping of LaCePrNiO3 after hydrogen reduction.
Catalysts 14 00186 g002
Figure 3. (a) N2 adsorption–desorption isotherms and (b) pore size distributions of the as-prepared sample.
Figure 3. (a) N2 adsorption–desorption isotherms and (b) pore size distributions of the as-prepared sample.
Catalysts 14 00186 g003
Figure 4. H2-TPR profiles of the as-prepared catalysts.
Figure 4. H2-TPR profiles of the as-prepared catalysts.
Catalysts 14 00186 g004
Figure 5. XPS spectra of the catalysts after reduction.
Figure 5. XPS spectra of the catalysts after reduction.
Catalysts 14 00186 g005
Figure 6. The conversion of n-dodecane and gas product rate for (a) LaNiO3, (b) LaCeNiO3, (c) LaCePrNiO3, (d) LaCeTbNiO3, (e) LaCeSmNiO3 and (f) comparison of the conversion of n-dodecane. Reaction condition: 600 °C, 0.1 MPa, LHSV of C12H26 15 mL/gcat∙h, with mole ratio of H2O:C12H26 at 24.
Figure 6. The conversion of n-dodecane and gas product rate for (a) LaNiO3, (b) LaCeNiO3, (c) LaCePrNiO3, (d) LaCeTbNiO3, (e) LaCeSmNiO3 and (f) comparison of the conversion of n-dodecane. Reaction condition: 600 °C, 0.1 MPa, LHSV of C12H26 15 mL/gcat∙h, with mole ratio of H2O:C12H26 at 24.
Catalysts 14 00186 g006
Figure 7. The relative contents of the gas products for steam reforming of n-dodecane. (a) LaNiO3, (b) LaCeNiO3, (c) LaCePrNiO3, (d) LaCeTbNiO3 and (e) LaCeSmNiO3. Reaction condition: 600 °C, 0.1 MPa, LHSV of C12H26 15 mL/gcat∙h, with mole ratio of H2O:C12H26 at 24.
Figure 7. The relative contents of the gas products for steam reforming of n-dodecane. (a) LaNiO3, (b) LaCeNiO3, (c) LaCePrNiO3, (d) LaCeTbNiO3 and (e) LaCeSmNiO3. Reaction condition: 600 °C, 0.1 MPa, LHSV of C12H26 15 mL/gcat∙h, with mole ratio of H2O:C12H26 at 24.
Catalysts 14 00186 g007
Figure 8. The conversion of n-dodecane and gas-production rate for (a) LaNiO3 and (b) LaCePrNiO3. Reaction condition: 600 °C, 0.1 MPa, LHSV of C12H26 15 mL/gcat∙h, with mole ratio of H2O:C12H26 at 48.
Figure 8. The conversion of n-dodecane and gas-production rate for (a) LaNiO3 and (b) LaCePrNiO3. Reaction condition: 600 °C, 0.1 MPa, LHSV of C12H26 15 mL/gcat∙h, with mole ratio of H2O:C12H26 at 48.
Catalysts 14 00186 g008
Figure 9. The TG and DTG analysis of the spent catalysts.
Figure 9. The TG and DTG analysis of the spent catalysts.
Catalysts 14 00186 g009
Figure 10. Raman spectra of the spent catalysts.
Figure 10. Raman spectra of the spent catalysts.
Catalysts 14 00186 g010
Table 1. The textural properties of as-prepared sample.
Table 1. The textural properties of as-prepared sample.
SampleBET Surface Area
(m2/g)
Pore Volume
(cm3/g)
Average Pore Size
(nm)
LaNiO32.840.6514.29
LaCeNiO31.150.2613.00
LaCePrNiO33.300.7617.59
LaCeSmNiO33.280.7523.84
LaCeTbNiO33.020.7520.69
Table 2. XPS of La 3d for LaNiO3, LaCeNiO3 and LaCePrNiO3.
Table 2. XPS of La 3d for LaNiO3, LaCeNiO3 and LaCePrNiO3.
SampleNamePeak Position (eV)AreaPercentages (%)
LaNiO3La 3d5/283589,853.5735.4
838.5556,589.0522.3
La 3d3/2852.0360,555.3723.8
855.3946,967.8818.5
LaCeNiO3La 3d5/283560,051.5735.8
838.5937,870.5222.6
La 3d3/2851.9741,413.9624.6
855.4728,454.6717.0
LaCePrNiO3La 3d5/2834.8372,638.2235.0
838.4947,519.3622.9
La 3d3/2851.9949,820.9124.0
855.4137,590.9118.1
Table 3. XPS of O 1s for LaNiO3, LaCeNiO3 and LaCePrNiO3.
Table 3. XPS of O 1s for LaNiO3, LaCeNiO3 and LaCePrNiO3.
SampleNamePeak Position (eV)AreaPercentages (%)
LaNiO3OL528.913,140.819.1
OV530.8226,106.8337.9
OSurf531.6917,864.2925.9
Oadv532.5111,747.8117.1
LaCeNiO3OL528.8110,728.1520.0
OV530.8518,226.533.9
OSurf531.6614,346.3826.7
Oadv532.410,391.3119.4
LaCePrNiO3OL528.9516,498.1924.8
OV530.9123,063.934.8
OSurf531.7416,776.4525.2
Oadv532.610,121.515.2
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

Guo, K.; Zhang, H.; Zhang, C.; Guo, X.; Li, H.; Xiao, Z. Perovskite-Derivative Ni-Based Catalysts for Hydrogen Production via Steam Reforming of Long-Chain Hydrocarbon Fuel. Catalysts 2024, 14, 186. https://doi.org/10.3390/catal14030186

AMA Style

Guo K, Zhang H, Zhang C, Guo X, Li H, Xiao Z. Perovskite-Derivative Ni-Based Catalysts for Hydrogen Production via Steam Reforming of Long-Chain Hydrocarbon Fuel. Catalysts. 2024; 14(3):186. https://doi.org/10.3390/catal14030186

Chicago/Turabian Style

Guo, Kai, Hui Zhang, Changxuan Zhang, Xining Guo, Huiying Li, and Zhourong Xiao. 2024. "Perovskite-Derivative Ni-Based Catalysts for Hydrogen Production via Steam Reforming of Long-Chain Hydrocarbon Fuel" Catalysts 14, no. 3: 186. https://doi.org/10.3390/catal14030186

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop