Aakash Rajpoota,
Afaq Ahmad Khan
a and
Ejaz Ahmad
*ab
aGreenCat Laboratory, Department of Chemical Engineering, Indian Institute of Technology (Indian School of Mines), Dhanbad, 826004, India. E-mail: ejaz@iitism.ac.in
bNaresh Vashisht Centre for Hydrogen and CCUS Technologies, Indian Institute of Technology (Indian School of Mines), Dhanbad, 826004, India
First published on 7th July 2025
The present study reports the synthesis of a series of nanocatalysts (NCs) comprising iron (Fe), cobalt (Co), and zinc (Zn) supported on γ-Al2O3 nanopowder (Fe5–Co–Zn/γ-Al2O3) by tuning the metals stoichiometric ratio using a facile co-precipitation approach. The powder X-ray diffraction patterns and Raman spectra revealed that the freshly synthesized Fe5–Co–Zn/γ-Al2O3 is composed of hexagonal Fe2O3, cubic Co3O4, and hexagonal ZnO phases. The NCs exhibited efficient performance in CH4 decomposition, yielding turquoise hydrogen (H2) and carbon nanotubes (CNTs) as solid carbon by-products. The active metals consisted of 50% Fe, 5% Co, and 15% Zn as Fe5Co0.5Zn1.5 given a maximum CH4 conversion of 90% and an H2 yield of 92.2%, which are considerably higher than those of many other previously reported NCs. The structural characteristics of the spent NCs (S-Fe5Co2Zn0 to S-Fe5Co0Zn2) were examined, showing variation in the growth and quality of CNT. Moreover, the electron microscopy micrographs suggested that the CNTs growth plausibly followed the base-growth model. The CNTs are also effectively isolated from spent NCs. This study paves the way for a streamlined approach to designing and unravelling the intricate distribution of active metals on suitable supports, offering deeper insights into optimizing CH4 decomposition into H2 and high-quality CNTs.
In contrast, natural gas is a widely used fossil fuel made primarily of 70% to 95% methane (CH4) and possesses high transportability. CH4 can be converted into H2 using several processes, such as partial oxidation,13 steam reforming,14 plasma reforming,15 and catalytic decomposition.16 Catalytic methane decomposition (CMD) is a highly promising process that can generate hydrogen without emitting CO or CO2 and requires less energy than steam and plasma reforming procedures.12 The CDM process is an endothermic reaction, as shown by the following equation:
![]() | (1) |
The process is widely recognized as pyrolysis of methane, sometimes known as thermal methane cracking.17 The CDM process can also be carried out at an operating temperature of around ∼700 to ∼900 °C. The choice of a catalyst plays a crucial role in reducing the necessary temperature and increasing the amount of desired product formation. Several factors influence CDM efficiency, including the catalyst metal type/composition, the support material type, and the interactions between the catalyst and support.18 The catalyst support must have thermal, chemical, and mechanical stability to ensure process safety and consistency. Additionally, it should offer a large surface area to facilitate the uniform distribution of active sites.19 Furthermore, the experimental reaction conditions influence the catalyst/support performance in CDM processes.20 The catalysts originated from group VIII, including nickel (Ni), cobalt (Co), and iron (Fe), are highly regarded for their exceptional catalytic capabilities and cost-effectiveness in the CDM process.21 The Ni-based catalysts have shown high activity in CH4 decomposition. However, it is highly prone to coking (carbon deposition), short lifespan, sintering, high-temperature stability, selectivity concerns, etc.16,22,23 As a consequence, the incorporation of Co and Fe into Ni-based catalysts has been studied to improve their performance and durability.24–26 However, the process of restoring deactivated Ni-based catalysts is still uncertain and not commercially viable at present. The researchers also developed beyond Ni-based catalysts such as 60%Fe/Al2O3,27 60%Fe/CeO2,28 Co/SiO2, Co/Nb2O5,18 Co/γ-Al2O3,29 and Zn–CoFe2O4,30 ZnFe2O4.31 Nevertheless, efforts are underway to develop an ideal catalyst to achieve a high turquoise hydrogen yield and produce high-quality solid carbon material.
Numerous investigations focused on CMD reactions indicate that the advancement of tri-metallic catalysts shows greater potential for enhancing reactivity and durability at ideal CH4 conversion temperatures.32 The non-noble metals-based trimetallic catalysts are promising and cost-effective due to their range of adjustable physicochemical properties. Fe-based catalyst is an attractive contender in CMD because of its low cost and partially full 3d orbitals, which allow it to assist the dissociation of hydrocarbons. Catalysts with added cobalt operate at lower temperatures and exhibit greater catalytic activity compared to those with only Fe-catalyst.33,34 The catalytic stability is also affected by the type of support utilized in catalyst preparation. Various supports are reported, such as Al2O3, La2O3, SiO2, MgO, H-ZSM-5, CeO2, TiO2, and CaO, to synthesize catalysts for CMD.35,36 Among these supports, Al2O3 has emerged as a promising choice for the CMD, probably due to its extensive specific surface area, effective interaction between metal and support, and excellent stability.37 In comparison, SiO2 demonstrates weak metal–support bonding, low acidity, and inadequate thermal stability, making it susceptible to metal sintering and carbon deposition.35 CeO2, on the other hand, contains abundant surface oxygen species that facilitate the oxidation of carbon deposits and the reoxidation of metal sites. Nevertheless, it suffers from moderate acidity, a relatively small surface area, poor thermal stability, and reduced reducibility due to solid solution formation.35,38 Meanwhile, TiO2 facilitates CH4 activation and dissociation, the formation of reduced TiOx species during the reaction can lead to the coverage of active metal particles. This encapsulation phenomenon potentially blocks active site accessibility, consequently impairing catalytic performance.35 So, the selection of Al2O3 as a support is a suitable choice to synthesize a catalyst for efficient methane decomposition.35,38 While there have been notable progressions in the synthesis of tri-metallic catalysts, the ability to Fe, Co, and Zn-based catalysts with controlled stoichiometry remains unachieved.
Herein, we report the synthesis of trimetallic Fe5–Co–Zn/γ-Al2O3 nanocatalysts (NCs) with tunable Co and Zn stoichiometry for CH4 decomposition in a temperature-programmed surface reaction (Scheme 1). The addition of Zn to Fe–Co catalysts is intended to study its role in improving dispersion, modifying metal-support interactions, and enhancing solid carbon structure, which can enhance catalytic efficiency and durability. The potential synergistic effects between these metals remain understudied, highlighting the novelty of this research work. The nanocatalysts were designed with a constant alumina support (30 wt%) and fixed Fe (50 wt%) content while varying the Co and Zn content from 0 to 20 wt% through a co-precipitation method. The synthesized nanocatalysts (NCs) exhibited effective performance in breaking down CH4, resulting in the production of turquoise hydrogen (H2) and carbon nanotubes (CNTs) as solid carbon by-products. Further, the reactivity for CH4 conversion, H2 yields, and the effect of reaction time of the various synthesized nanocatalysts are studied. Additionally, various analytical techniques were utilized to investigate the physico-chemical characteristics of fresh, reduced, and spent nanocatalysts.
![]() | ||
Scheme 1 A schematic illustration of the cobalt (Co) and zinc (Zn) stoichiometric ratio-controlled synthesis of trimetallic Fe5–Co–Zn/γ-Al2O3 nanocatalysts. |
Subsequently, the quartz tube underwent degassing with argon for 60 minutes to establish an inert environment. Following this, the methane and argon gases were introduced into the system at a volume ratio of 1:
19 and maintained at this ratio throughout the 150 minute reaction period. Additionally, the space velocity was held constant at 1.2 litre per hour per gram of NCs. The outlet of the reactor was connected to a mass spectroscopy apparatus equipped with a thermal conductivity detector (TCD) for the examination and determination of composition and molecular weight. The conversion of methane, the yield of turquoise hydrogen, and carbon production were determined utilizing eqn (2) and (3).
![]() | (2) |
![]() | (3) |
Raman spectroscopic analysis was also performed to confirm the formation of Co3O4, ZnO, and Fe2O3 in Fe5Co2Zn0 to Fe5Co0Zn2 (Fig. 1b). The two peaks with the frequency of 194.1 and 689.8 cm−1 in Fe5Co2Zn0 denoted as Raman active modes F 2g and A 1g, respectively is well aligned with the Co3O4.41 Furthermore, an increased in Zn wt% with decreasing Co wt% (Fe5Co1.5Zn0.5 to Fe5Co0.5Zn1.5), new peaks appeared along with the Co3O4 Raman peaks. Further, the increase in Zn wt% (Fe5Co0Zn2), F 2g (194.1 cm−1), and A 1g (689.8 cm−1) modes have completely disappeared. However, four peaks with the frequency of 403, 433, 591, and 615 cm−1 assigned as Raman active modes of E1, E2, A1, and E1, respectively, corresponded to the ZnO.42 Additionally, the two most prominent peaks detected in all NCs correspond to the primary Raman active bands, identified as A 1g and Eg, which are attributed to the α-Fe2O3 phase.43 So, the Raman results also correspond to our powder-XRD outcomes and indicate that the synthesized trimetallic nanocatalysts are composed of Co3O4, ZnO, and α-Fe2O3.
X-ray photoelectron spectroscopy (XPS) analysis of freshly synthesized trimetallic Fe5Co1.5Zn0.5, Fe5Co1Zn1, and Fe5Co0.5Zn1.5 was performed to investigate the chemical composition and electronic state on the surface of the NCs. The survey spectrum peaks clearly indicate that the NCs are composed of iron (Fe), cobalt (Co), zinc (Zn), aluminium (Al), and oxygen (O) elements (Fig. S1a†). The core-level spectra of Fe 2p, Co 2p, Zn 2p, O 1s, and Al 2p were deconvoluted to investigate the valence state present in the trimetallic NCs (Fig. 2 and S1†). The obtained binding energies of the valence states are presented in Table S3.† The core-level spectrum of Fe 2p is fitted into two pairs of peaks (2p3/2 and 2p1/2), with the two peaks of 2p3/2 corresponding to the Fe3+(Oh) and Fe3+(Td) valence states, respectively (Fig. 2a). Moreover, the presence of two binding energy peaks is a result of Fe3+ forming octahedral (Oh) and tetrahedral (Td) bonds with oxygen.44 The core-level Co 2p spectra contain two major peaks related to Co 2p3/2 and Co 2p1/2, which are further deconvoluted into four peaks (Fig. 2b). The first pair of peaks with lower binding energy signifies the presence of Co3+. In contrast, the second pair of peaks with higher binding energy indicate the Co2+ valence state.45 The Zn 2p core-level spectrum was fitted into two prominent peaks corresponding to Zn 2p3/2 and Zn 2p1/2, with the obtained binding energies indicating the Zn2+ valence state (Fig. 2c).46 Further, O 1s spectrum was deconvoluted into three peaks around 529.5 eV, 530.8 eV, and 531.8 eV, corresponding to metal-bonded oxygen (M–O), aluminium-bonded oxygen (Al–O), and adsorbed oxygen (Oads), respectively (Fig. S1b†).47 In addition, the Al 2p spectrum exhibits two prominent peaks around 73 eV and 75 eV that resemble the Al3+ valence state and Al–OH, respectively (Fig. S1c†).47,48 So, the obtained results indicated that the valence states present in freshly synthesized trimetallic nanocatalysts are Fe3+, Co3+, Co2+, Zn2+, and Al3+, which further suggests that O is chemically bonded to Fe, Co, Zn, and Al in the form of Fe–O, Zn–O, Co–O, and Al–O.
![]() | ||
Fig. 2 Core-level XPS spectra: (a) Fe2p, (b) Co 2p, (c) Zn 2p of trimetallic Fe5–Co–Zn/γ-Al2O3 nanocatalysts. |
The Brunauer–Emmett–Teller (BET) surface area (SBET), pore width, and pore volume of the Fe5–Co–Zn/γ-Al2O3 NCs with varying Co and Zn wt% ratios were determined using N2 adsorption–desorption isotherms (Fig. S2†). All the synthesized NCs displayed a type IV isotherm with an H3 hysteresis loop, signifying their predominantly mesoporous structure (inset Fig. S2†). The SBET, pore width, and pore volume of the NCs vary based on the stoichiometric ratio of Co and Zn, as shown in Table 1. As the Zn wt% increased, the SBET values decreased, indicating the formation of larger particles and a resulting reduction in the surface-to-volume ratio. The pore width of the NCs was determined using the Barrett–Joyner–Halenda (BJH) method, revealing that the pore width of the trimetallic Fe5Co1.5Zn0.5, Fe5Co1Zn1, and Fe5Co0.5Zn1.5 NCs is increased with a decrease in Co wt% and an increase in Zn wt%, which might be due to the inter-particle pores formation. Moreover, the pore width of bimetallic Fe5Co2Zn0 and Fe5Co0Zn2 NCs suggested that the Zn-containing NCs formed a bigger pore. Furthermore, the pore width of the synthesized trimetallic NCs is substantially greater than the molecular size of CH4 (0.436 nm), leading to better surface adsorption, which further increases the CH4 decomposition rate into H2 and CNTs.
Nanocatalysts | SBET (m2 g−1) | Pore volume (cm3 g−1) | Pore width (nm) |
---|---|---|---|
Fe5Co2Zn0 | 40.3 | 0.230 | 24.1 |
Fe5Co1.5Zn0.5 | 36.1 | 0.211 | 19.0 |
Fe5Co1Zn1 | 31.8 | 0.193 | 25.1 |
Fe5Co0.5Zn1.5 | 27.7 | 0.192 | 29.8 |
Fe5Co0Zn2 | 26.2 | 0.175 | 28.9 |
The surface morphology of Fe5–Co–Zn/γ-Al2O3 NCs was analyzed using field-emission scanning electron microscopy (FE-SEM) micrographs (Fig. S3†). The micrograph of Fe5Co2Zn0 displays a bead-like structure with some flakes stacked on the surface, indicating the formation of a hetero-structured morphology (Fig. S3a†). As the Zn content increases to 15% and Co decreases to 5% (Fe5Co1.5Zn0.5 to Fe5Co0.5Zn1.5), the density of the stacked flakes increases (Fig. S3(b)–(d)†). However, the flakes became more prominent when Zn content was further increased to 20% (Fe5Co0Zn2) (Fig. S3e†). So, the obtained micrographs suggested that the Zn promotes the formation of a more flake-like structure on the bead-like surface. Further, the TEM micrograph of NCs confirms the growth of irregular flake structures with variable sizes (Fig. S4†). The high-resolution TEM (HR-TEM) micrograph displays that the flakes are more uniformly distributed in trimetallic Fe5Co1.5Zn0.5, Fe5Co1Zn1, and Fe5Co0.5Zn1.5 NCs (Fig. S4e, h and k†). However, the bimetallic Fe5Co2Zn0 and Fe5Co0Zn2 NCs show the stacked small flakes growth with variable shapes and sizes (Fig. S4b and n†). The selected area electron diffraction (SAED) patterns of Fe5–Co–Zn/γ-Al2O3 NCs showed that the polycrystalline in nature (Fig. S4c, f, i, l and o†).
Further, the Fe, Co, and Zn content in the freshly synthesized Fe5Co2Zn0 to Fe5Co0Zn2 NCs was measured using inductively coupled plasma optical emission spectroscopy (ICP-OES). The procedure entailed digesting 2 mg of the sample in aqua regia for 6 hours, followed by dilution prior to measurement. The ICP-OES analysis revealed that the concentrations of cobalt (Co) and zinc (Zn) varied according to the stoichiometric ratios of Co and Zn employed during catalyst synthesis (Table S4†).
Similarly, the trimetallic Fe5Co1.5Zn0.5 to Fe5Co0.5Zn1.5 NCs exhibit an increasing trend in H2 yield with an increase in Zn content from 5 to 15 wt% and a decrease in Co content from 15 to 5 wt% (Fig. 3c). However, further increasing the Zn content to 20 wt% (Fe5Co0Zn2) results in a decrease in H2 yield. The highest H2 yield, 92.2%, was achieved for trimetallic Fe5Co0.5Zn1.5 after 25 minutes of reaction time. The H2 yields for Fe5Co2Zn0, Fe5Co1.5Zn0.5, Fe5Co1Zn1, and Fe5Co2Zn0 NCs were 79.4%, 77.2%, 91.7%, and 73.3%, respectively.
The H2 yield slightly decreased after the reaction time reached 150 minutes, which closely aligns with the trends observed in CH4 conversion. The reduction after 25 minutes of reaction time in CH4 conversion and H2 yield can be attributed to the formation of encapsulating carbon, which obstructs access to active sites by covering their surfaces (Fig. 3d). Specifically, the maximum H2 yield for Fe5Co0.5Zn1.5 NCs decreased from 92.2% to 67.2% (Table S6†).
The selectivity of H2 production is determined by comparing the actual amount of hydrogen generated to the theoretical amount expected from the catalytic decomposition of CH4. The calculation of H2 selectivity is based on eqn (4);
![]() | (4) |
H2 selectivity was determined at the point of maximum CH4 conversion (25 minutes). It is observed that the H2 selectivity of trimetallic NCs is increased as the Co wt% decreased from 15 to 5% (Table 2). The trimetallic Fe5Co0.5Zn1.5 exhibited the highest selectivity, achieving 92.1% for H2 production. Furthermore, the findings indicate that the CH4 conversion efficiency and H2 production performance of the synthesized nanocatalysts at elevated temperatures (750 °C) are significantly affected by the tuning of the active metal's ratio (Co and Zn). The CH4 decomposition reaction was also repeated up to two times using the synthesized NCs, and the CH4 conversion (%) and H2 yield (%) remained consistent across the second and third trials, demonstrating the reproducibility of the experiments (Table S7†). Additionally, the obtained CH4 conversion and H2 yields are significantly higher than the various previously reported catalysts, including Co/SiO2, Co/Nb2O5, 60% Fe/Al2O3, 60% Fe/CeO2, Co/γ-Al2O3, Zn–Co–Fe2O3, and ZnFe2O4.18,27,28,30–32 Moreover, a comparison table of conversion, yield, and reaction time of some other previously reported results is presented in Table S8.† The turnover frequency (TOF) at 150 min was also determined to elaborate on the effectiveness of the nanocatalyst for H2 production. The TOF was determined by calculating the active sites of the catalysts using H2 temperature-programmed desorption analysis, and the corresponding values are presented in Table S9.†49 The Fe5Co0.5Zn1.5 gave the maximum active sites of 0.068 mmole/gram; however, Fe5Co1Zn1 gave the maximum TOF as 165.4 h−1 at 150 minutes with active sites of 0.051 mmole gram−1. The TOF of the most efficient catalyst, Fe5Co0.5Zn1.5, was determined as 156.5 h−1 at 150 minutes, suggesting that not all active sites are involved in the reaction might be due to the lower flow rate of CH4 during reaction. However, Fe5Co0.5Zn1.5 provided maximum CH4 conversion and H2 yields (Fig. 3). Further, it was noted that the active sites value of the catalyst is raised to Fe5Co0.5Zn1.5 and decreased by further increase of Zn content to Fe5Co0Zn2. So, the obtained results indicate that suitable Co and Zn are beneficial to achieve better active sites and stability for CH4 conversion.
S. No. | Nanocatalyst | Actual H2 produced (millimoles) | Theoretical H2 produced (millimoles) | Selectivity of H2 (%) |
---|---|---|---|---|
1 | Fe5Co2Zn0 | 0.473 | 0.595 | 79.5 |
2 | Fe5Co1.5Zn0.5 | 0.460 | 0.595 | 77.3 |
3 | Fe5Co1Zn1 | 0.546 | 0.595 | 91.7 |
4 | Fe5Co0.5Zn1.5 | 0.548 | 0.595 | 92.1 |
5 | Fe5Co0Zn2 | 0.436 | 0.595 | 73.2 |
As we discussed earlier, along with H2, the solid carbon was also formed as a by-product in a catalytic CH4 decomposition reaction. So, the carbon content in spent catalysts was determined using CHNS analysis after 150 minutes of reaction time. The obtained results indicate that the carbon content increased in S-Fe5Co2Zn0 to S-Fe5Co0.5Zn1.5 and then decreased in Fe5Co0Zn2 with a maximum carbon content of 56.38% in S-Fe5Co0.5Zn1.5 (Table 3). The amounts of hydrogen (H), nitrogen (N), and sulfur (S) were minimal compared to carbon. The analysis also suggested that trimetallic Fe5Co0.5Zn1.5 promotes the maximum CNTs formation and highest yield of H2. Additionally, the amount of CNTs produced is influenced by the quantity of H2 generated and the duration of the reaction. So, the obtained results suggest that the Co and Zn ratios play a crucial role in determining the selectivity, quantity, and quality of the products.
S. No. | Sample weight (mg) | Nanocatalysts | N [%] | C [%] | H [%] | S [%] |
---|---|---|---|---|---|---|
1 | 2.38 | S-Fe5Co2Zn0 | 0.01 | 33.40 | 0.05 | 0.01 |
2 | 2.73 | S-Fe5Co1.5Zn0.5 | 0 | 44.74 | 0.05 | 0.03 |
3 | 2.10 | S-Fe5Co1Zn1 | 0.01 | 50.04 | 0.05 | 0.04 |
4 | 2.24 | S-Fe5Co0.5Zn1.5 | 0 | 56.38 | 0.06 | 0.04 |
5 | 2.62 | S-Fe5Co0Zn2 | 0.01 | 49.26 | 0.05 | 0.05 |
Powder X-ray diffraction analysis of reduced and spent Fe5Co2Zn0 to Fe5Co0Zn2 was performed to identify the changes in phase and crystallographic structure after 150 minutes of reaction time (Fig. 4). The obtained diffraction pattern of the reduced as well as spent NCs reveals the most intense peaks around 2θ = 44°, which is corresponds to the planes (111) of the cubic phase of metallic γ-Fe with the space group of Fmm (JCPDF no. 005-3803, a = b = c = 3.61 Å) and cubic phase of metallic α-Co with the space group of Fm
m (JCPDF no. 004-4989, a = b = c = 3.54 Å). When the Zn content increased up to 20 wt%, a new diffraction peak appeared at 2θ = 36.7°, which is aligned with the (002) plane of the hexagonal phase of metallic Zn with space group of P63/mmc (JCPDF no. 042-1013, a = b = c = 2.66 Å). In Fig. 4b, the spent NCs showed another diffraction peak at 2θ = 26° of the (002) plane, which is well matched with the hexagonal phase of carbon (C1) with the space group P63/mmc (JCPDS no. 061-7290, a = b = 2.47 Å, c = 6.93 Å). So, the obtained diffraction patterns suggested that the freshly synthesized bimetallic and trimetallic metal oxide NCs are reduced to their metallic phase during and after the catalytic CH4 decomposition reaction.
![]() | ||
Fig. 4 Powder-XRD diffraction patterns: (a) reduced Fe5–Co–Zn/γ-Al2O3, and (b) spent Fe5–Co–Zn/γ-Al2O3 NCs. |
Raman spectroscopy was employed to investigate the degree of graphitization and crystallinity of the deposited CNTs. In the Raman spectra, two clear bands can be observed, one near ∼1570 cm−1 (referred to as the G-band), which is linked to the graphitization process and results from in-plane lattice vibrations of the CC (sp2) bond (Fig. 5a). The additional band, noted at ∼1340 cm−1 (D-band), resulted from the structural disorder present in the graphene structure. The intensity ratio of the D and G-band can reveal the quality of CNTs, as it is linked to the degree of structural disorder in the material. The spectral data, which included D- and G-band intensities, were converted into ratios (ID/IG). The minimum value of ID/IG is 0.37 for S-Fe5Co0.5Zn1.5, whereas the highest value is 0.49 for both S-Fe5Co2Zn0 and S-Fe5Co1.5Zn0.5. The ID/IG values varied from 0.37 to 0.49 for S-Fe5Co2Zn0 to S-Fe5Co0Zn2, as presented in Table S10.† The Raman findings indicated that the formation of CNTs was successfully achieved, and the quality of CNTs can be modified by adjusting the stoichiometric ratio of Co and Zn in the nanocatalyst.
![]() | ||
Fig. 5 (a) Raman spectra of the in the spectral range 100 cm−1–4000 cm−1 (λ = 532 nm), and (b) FTIR patterns of the spent nanocatalyst of S-Fe5Co2Zn0 to S-Fe5Co0Zn2. |
To further investigate the functional characteristics of CNTs, FTIR spectroscopy was performed in the range of 550–4000 cm−1 of S-Fe5Co2Zn0 to S-Fe5Co0Zn2 (Fig. 5b). A broad band observed at 3442 cm−1 was attributed to hydrogen-bonded O–H groups from moisture. The two separate peaks at 2980–2820 cm−1 were associated with the asymmetric and symmetric stretching of the C–H bond. The peaks between 1490–1350 cm−1 can be ascribed to asymmetric bending vibration of the C–H bond. The band appeared around 1740 cm−1 corresponding to the stretching vibration of the CO bond. Additionally, the band at ∼1627 cm−1 is stretching vibration of aromatic C
C bonds in the sp2 carbon domain.41 The three peaks observed at 1110, 809, and 657 cm−1 were assigned to the stretching vibration of the –O–R group, the aromatic out-of-plane C–H deformations, and the M–O bond, respectively.50
The morphology of spent Fe5–Co–Zn/γ-Al2O3 NCs was examined using FE-SEM micrographs (Fig. 6 and S6†). The micrographs of bimetallic S-Fe5Co2Zn0 show that the nanotubes are spread on the surface of catalysts (Fig. S6a†). In the trimetallic NCs, as the Zn content increases to 15% and Co decreases to 5% (Fe5Co1.5Zn0.5 to Fe5Co0.5Zn1.5), the density of the nanotubes significantly increases (Fig. 6d). Further increase in Zn content to 20% (Fe5Co0Zn2), the nanotubes density on the surface of the catalyst is slightly decreased (Fig. S6b†). It is displayed that the trimetallic S-Fe5Co0.5Zn1.5 formed more prominent nanotubes. To further analyse the morphology of CNTs, TEM micrographs of the S-Fe5Co2Zn0 to S-Fe5Co0Zn2 are assessed to elaborate the morphological characteristic at higher resolution. The obtained results revealed that the length and quality of CNTs are significantly influenced by the stoichiometric ratio of Co and Zn (Fig. 6d–f, S7 and S8†). The micrographs showed that as the Zn content increased to 15% and Co decreased to 5%, the nanotube length and uniformity improved. However, when the Zn content reached 20%, the growth of CNTs became distorted. Notably, the S-Fe5Co0.5Zn1.5 produced more uniform CNTs compared to the other synthesized NCs (Fig. 6d and S8†). The inset micrographs of Fig. 6 and S7† display that the CNTs are wrapped around the tops of the catalyst particles, indicating multi-walled carbon nanotubes (MWCNTs) formation, which seems to be grown by a base-growth mechanism.51 The HR-TEM micrographs of S-Fe5Co2Zn0 to S-Fe5Co0Zn2 revealed characteristic fringes of CNTs, with the interplanar spacing in the region denoted by a black dotted box measured at 0.346 to 0.348 nm. These values correspond to the (002) plane of hexagonal carbon (C1) and are consistent with our powder-XRD results (inset Fig. 6d–f and inset Fig. S7†). The SAED patterns of S-Fe5Co2Zn0 to S-Fe5Co0Zn2 confirm the polycrystalline nature of the NCs (Fig. S9†). The STEM-EDS elemental mappings show a uniform distribution of carbon (C) and metals across the catalyst surface, indicating the stability and formation of high-quality CNTs (Fig. S10a–e†). The oxygen observed in the spent NCs is attributed to the presence of γ-Al2O3 and the adsorption of atmospheric moisture on the surface.
![]() | ||
Fig. 6 FE-SEM micrographs of spent-trimetallic nanocatalysts (a) S-Fe5Co1.5Zn0.5, (b) S-Fe5Co1Zn1, (c) S-Fe5Co0.5Zn1.5. |
The XPS analysis of our best catalyst (S-Fe5Co0.50Zn1.5) was performed to investigate the chemical composition and electronic state on the surface after the CH4 decomposition reaction for 150 minutes (Fig. 7). The survey spectrum peaks clearly revealed that the S-Fe5Co0.5Zn1.5 consists of iron (Fe), cobalt (Co), zinc (Zn), carbon (C), aluminium (Al), and oxygen (O), elements (Fig. S11†). To examine the valence states within the S-Fe5Co0.5Zn1.5, the core-level spectra of Fe 2p, Co 2p, Zn 2p, C 1s, O 1s, and Al 2p were deconvoluted. The core-level spectrum of Fe 2p is fitted with three peaks of 2p3/2 at 707.2 eV, 711.3 eV, and 714.1 eV corresponding to the Fe(0), Fe3+(Oh) and Fe3+(Td) valence states, respectively (Fig. 7a).
![]() | ||
Fig. 7 XPS core-level spectra: (a) Fe 2p, (b) Co 2p, (c) Zn 2p, (d) C 1s, (e) O 1s, and (f) 2p of S-Fe5Co0.5Ni1.5. |
The core-level Co 2p spectra contain major peaks related to Co 2p3/2 with a binding energy of 783.4 eV, which signifies the presence of Co2+ valence state (Fig. 7b).45 The Zn 2p core-level spectrum was fitted into a peak 2p3/2 at 1022.9 eV indicating the Zn2+ valence state (Fig. 7c).46 The presence of Fe3+, Co2+, and Zn2+ in the S-Fe5Co0.5Zn1.5 due to the aerials oxidation of catalysts after reaction. The core-level C 1s spectrum was fitted into three peaks, 284.3 eV, 285.1 eV, and 289.6 eV, indicating the presence of sp2/sp3 carbon (CC/C–C), –C–O, and –COO group of CNTs, which coincides with the origin of D-band in our Raman spectra (Fig. 7d and 5a).52 Additionally, the O 1s spectrum was deconvoluted into two peaks at approximately 532.2 eV and 533.8 eV, corresponding to adsorbed oxygen (Oads) and water (OH2O), respectively (Fig. 7e). Furthermore, the Al 2p spectrum shows two distinct peaks near 75.1 eV and 76.2 eV, indicating the presence of Al2O3. The obtained XPS binding energies reveal downshifts in the binding energies of Fe 2p and Co 2p in S-Fe5Co0.5Zn1.5 compared to fresh Fe5Co0.5Zn1.5 (Table S2†), indicating elevated electron density at the Fe and Co centres. This is attributed to electron donation from Zn2+ ions (3d10 electronic configuration), which facilitate charge redistribution within the multi-metallic system. In the Fe5Co0.5Zn1.5 nanocatalyst, a decrease in Co2+ concentration was observed, while in the spent catalyst, an increase in Co3+ concentration was observed, as shown in Table S11.† This indicates electron transfer between Co and Fe during the reaction. Similarly, for Fe species, the concentration of Fe3+(Oh) sites decreased, whereas the Fe3+(Td) sites increased, suggesting a redistribution of electrons and coordination environment, further supporting the occurrence of metal–metal electronic interactions within the catalyst. So, the synergistic effect between Fe, Co, and Zn plays significant role in the CH4 decomposition to turquoise H2 and CNTs productions.
![]() | ||
Fig. 8 (a) TEM micrograph of CNTs, (b) dark-field micrograph of CNTs, and (c) STEM-EDS image mapping of separated CNTs from S-Fe5Co0.5Zn1.5. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5na00237k |
This journal is © The Royal Society of Chemistry 2025 |