Next Article in Journal
Genetical and Biochemical Basis of Methane Monooxygenases of Methylosinus trichosporium OB3b in Response to Copper
Previous Article in Journal
Towards a Mechanistic Understanding of the Slagging Propensities of Petroleum Coke: Lessons Learned from Its Co-Combustion with Natural Gas in Oxygen-Enriched Atmospheres
Previous Article in Special Issue
Dry Reforming of Methane over Li-Doped Ni/TiO2 Catalysts: Effect of Support Basicity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress on Stability Control on Ni-Based Catalysts for Methane Dry Reforming

by
Minghui Wei
1 and
Xuerong Shi
1,2,*
1
School of Materials Engineering, Shanghai University of Engineering Science, Shanghai 201620, China
2
National Key Laboratory of High Efficiency and Low Carbon Utilization of Coal, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, China
*
Author to whom correspondence should be addressed.
Methane 2024, 3(1), 86-102; https://doi.org/10.3390/methane3010006
Submission received: 17 November 2023 / Revised: 13 December 2023 / Accepted: 27 December 2023 / Published: 6 February 2024
(This article belongs to the Special Issue Methane Dry Reforming)

Abstract

:
CO2 reforming of CH4 (DRM) utilizes the greenhouse gases of CH4 and CO2 to obtain the synthesis gas, benefiting the achievement of carbon neutrality. However, the deactivation of Ni-based catalysts caused by sintering and carbon deposition limits the industrial application. Focusing on stability improvement, this review first summarizes the reaction mechanism and deactivation mechanism in DRM and then discusses the impact of catalyst active components, supports, and interfacial structure. Finally, we propose the design direction of stable Ni-based catalysts towards DRM, providing guidance for the future development of catalysts suitable for industrial production.

1. Introduction

Methane (CH4) and carbon dioxide (CO2) are two major greenhouse gases but also important carbon resources [1,2,3,4]. Catalytic conversion and utilization of these two greenhouse gases through chemical pathways is of positive significance for mitigating the greenhouse effect and closing the carbon cycle [5]. Additionally, CH4 and CO2 are important components of biogas, which come from a wide range of sources [6]. Methane dry reforming (DRM) can convert CH4 and CO2 into carbon-rich synthesis gas, which is particularly suitable for hydroformylation synthesis to fulfill the high-value utilization of biogas. DRM, which is strongly endothermic thermodynamically, needs to be performed under high temperature (700–1000 °C) conditions. On the one hand, this leads to high energy consumption, and on the other hand, there is catalyst deactivation caused by high-temperature sintering and carbon deposition. Therefore, how to improve the stability and low-temperature activity of the DRM catalyst has become the key to industrial application [7]. Among various metal-based DRM catalysts, the inexpensive earth-abundant Ni-based catalysts stand out because they show high catalytic activity, comparable to noble metals such as Ru and Rh [8]. However, Ni-based catalysts also deliver excellent C-H bond cleavage ability. The formed carbon species deposit on the Ni surface and form stable carbon structures, leading to the block of metal-active sites and inducing deactivation [9,10,11,12].
To understand and seek possible solutions to catalyst deactivation issues, this review first reveals the reaction mechanism and deactivation mechanism of DRM and then discusses the influence of catalyst active components, supports, and microenvironment. Finally, the development direction of stable DRM catalysts is proposed, which provides guidance for the design of catalysts suitable for industrial applications (Figure 1).

2. DRM Reaction Mechanism and Deactivation Mechanism

The reaction network of DRM is complex. The reaction (Entry 1), Table 1 is accompanied by side reactions such as the reverse water gas shift (RWGS) reaction (Entry 2) and carbon deposition reaction, including CH4 dissociation (Entry 3), CO disproportionation, namely, the Boudouard reaction (Entry 4) and CO redox (Entry 5).
As can be seen from the above reaction equation, DRM is strongly endothermic, and a high temperature is required to promote the reaction. Due to the presence of the side reaction of (Entry 2), the CO:H2 mole ratio is less than 1. Since the DRM reaction is carried out in the thermodynamic carbon deposition zone (Entry 3), it is impossible to improve the carbon deposition of the DRM reaction by changing the thermodynamic conditions of the reaction. Therefore, this review focuses on possible kinetic ways to increase the reaction rate of DRM while decreasing the reaction rates of side reactions by the adjustment of influencing factors.

2.1. Reaction Mechanism

The DRM mechanism has been extensively studied [13,14]. It is generally believed that DRM reactions on Ni-based catalysts follow these steps:
(1) Activation of CH4. It generally occurs on the active site of the metal and at the interface between the metal and the support. CH4 adsorbs on these active metal surfaces and gradually dissociates to C* or CH* [15,16,17].
(2) Activation of CO2. The activation of CO2 usually occurs on the support or at the interface between the support or the support rather than on metal. The obtained H from the CH4 dissociation on Ni may assist CO2 activation via COOH* or HCOO* species [18]. The rate-limiting step varies depending on the system. The dissociation of CH4 on the catalyst is generally considered to be a rate control step for DRM [18]. M-O* species formed on metal surfaces after preferential activation of CO2 can promote methane activation [19]. However, some studies have shown that CO2 dissociation at high temperatures may be a rate-limiting step [20]. For bifunctional catalysts, such as Pd/ZrO2, it is a multi-step-controlled reaction [21].
(3) Oxidation and decomposition of intermediates. C* or CH*, which is from the CH4 dissociation, is oxidized by O* to produce CO* or CHO*, and then CHO* is decomposed into CO* and H* [21]. The origin of O* is diverse, which could come from the decomposition of CO2 or from the oxide support [22,23,24]. For those supports with high oxygen release capacity and lattice oxygen mobility, such as La2O3 [23] and CeO2 [23], direct oxidation of CHx species by the lattice oxygen (Olattice) via a Mars−van Krevelen-type redox mechanism is proposed. In this mechanism, the active carbon intermediates from methane dissociation are oxidized directly by the lattice oxygen of the support to form CO and oxygen vacancies, avoiding the formation of carbon species on the Ni surface [23]. The oxygen vacancies then serve as the activation sites for CO2 dissociation into CO and O [23].
(4) Formation of hydroxyl groups. H* generated by CH4 decomposition could diffuse from the surface of active metal particles to the support under the DRM reaction condition and react with the O* species generated by CO2 decomposition to form OH species [23].
(5) Syngas formation. In the final step, syngas is formed as a result of a radical recombination.

2.2. Catalyst Deactivation Mechanism

In the DRM reaction, the deactivation of the catalyst is mainly caused by carbon deposition and sintering of the active components.
Carbon deposition is the process of carbon species covering the active sites, reducing the contact between reactants and active metals. Carbon species can be classified into different types based on morphology, properties, and relative deactivation rate. It includes the adsorbed atomic carbon Cα, polymeric carbon filaments Cβ, graphitic carbon Cc, and carbon whiskers Cv (Figure 2). Cα is formed from the dissociation of reactants and products of DRM, CO2, CH4, and CO on the catalyst. When the temperature is above 250 °C, the generation rate of Cα will exceed its gasification rate and accumulate on the surface. As it accumulates, they react together to convert into Cβ. As the temperature continues to rise, Cα and Cβ will convert into carbon whiskers Cv and graphite carbon Cc. It is noteworthy that not every kind of carbon species reduces the activity of the catalyst. For example, Cα and Cβ can be eliminated by oxidants formed during the reaction with little impact on the catalytic activity. However, at 500–900 °C, the carbon whiskers Cv exhibit thermodynamic stability with a thickness of over 80 atomic diameters, making it difficult to oxidize and remove. In addition, when generating a large amount of Cv, it will cause the active metal components to move and aggregate. Graphite carbon Cc is also difficult to eliminate through oxidation reactions [25]. It will cover the surface of the catalyst, even completely envelop the active sites, block pores, isolate internal active components, and prevent them from contacting reactants, thus deactivating the catalysts. In addition, Cc can also block the reactor, increase the load pressure of the catalytic bed, and damage the structure of the catalyst. In summary, at the typical DRM temperature, 500–900 °C, the main carbon species are carbon whiskers Cv and graphite carbon Cc [25]. Therefore, for Ni-based catalysts in the DRM process, it is crucial to solve the problem of carbon deposition to maintain the activity and stability simultaneously.
Sintering is the migration and agglomeration of the active metal at high temperatures, which reduces the number of active sites, resulting in a decrease in the activity of the catalyst [26]. The oxidation of active metals during the reaction and its reaction with the support or promoters, as well as the collapse of the pore structure of the catalyst support at high temperatures, will also cause catalyst deactivation. Although it is difficult to avoid carbon deposition and sintering of the active components, the above problems can be effectively mitigated by the design and modulation of the catalyst.
Additionally, the possible presence of temperature hysteresis phenomenon in the typical endothermic reaction of DRM could be one of the reasons for deactivation. However, the mechanism of its occurrence is still unclear. Wang et al. [27] studied the hysteresis phenomenon of the Ni/SBA-15 catalyst in the DRM reaction and deeply analyzed the reasons and mechanisms behind this phenomenon (Figure 3). The ignition temperature (Tign) is 425 °C, and the text temperature is 295 °C. At temperatures below Tign, the surface Ni0 is oxidized to NiO by CO2 or O in the support, making the reaction difficult to proceed. At Tign, NiO reduction and methane cracking occur simultaneously, generating carbon nanotubes that inhibit the oxidation of Ni0. When the reaction temperature is higher than Tign, carbon nanotubes generated by methane cracking form a unique catalyst structure with surface Ni, which promotes the catalyst to still have catalytic activity at temperatures lower than Tign. When it is below Text (the activity decreases to 0% at Text), CH4 cannot continue to reduce NiO, the cycle between the NiO and Ni is broken, and the catalyst is deactivated [27] (Figure 3).

3. Regulation of Carbon Deposition on Ni-Based DRM Catalysts

Many strategies, such as structural and electronic engineering, are employed to adjust the properties of the catalysts. For the control of carbon deposition on Ni-based DRM catalysts, the design includes the following aspects.

3.1. Regulation of Active Metals

3.1.1. Ni Configuration

The configuration of Ni matters during the catalytic reaction [28,29]. The zero-valenced Ni centers, namely metallic Ni, are required for DRM, while the limited amount of metallic Ni minimizes carbon deposition [30,31,32]. Wang et al. [15] found that for the Ni-Zr/SiO2 catalyst, the Ni0 species was oxidized to a NiO species under the reaction conditions and could not be restored, leading to its deactivation.
The carbon deposition causes the deformation and recombination of the active metal particles. Due to the friction exerted by carbon deposits, the Ni particles in the Ni-based catalyst may become pear-shaped and transfer to elongated conical at high temperatures. In addition, since high Ni dispersion reduces metal sintering and hinders Cv formation, the stability of the catalyst can be enhanced by increasing the dispersion of Ni and adopting the appropriate operating temperature for the Ni-based catalyst. Note that agglomeration and sintering can again increase above an optimal limit of metal dispersion.

3.1.2. Ni Particle Size

The size of Ni particles is another important factor controlling the stability of DRM catalysts. Recently, in order to fully utilize the metal center [33], a series of single-atom catalysts have been developed for DRM applications. Mohcin Akri, et al. [34] designed and prepared a stable Ni single-atom catalyst through interaction with Ce-doped hydroxyapatite (HAP), showing a high activity of DRM and excellent carbon resistance ability. Experimental and computational studies have shown that isolated nickel atoms are inherently carbon-resistant because they have the unique property of activating only the first carbon-hydrogen bond in CH4, thus avoiding the deep decomposition of methane into deposited carbon species. It was found that the cerium doping of HAP support stabilized the atomically dispersed Ni, which significantly improved the anti-sintering performance of the 0.5Ni1/HAP-Ce catalyst. After 65 h reactions, the conversion of CO2 and CH4 retains 90% of the initial with a small decrease. The HAP-Ce support has almost no activity towards DRM, indicating that its activity mainly comes from atomically dispersed Ni. The main role of Ce is to stabilize atomically dispersed Ni rather than to inhibit carbon deposition. This observation is consistent with results from recently reported Fe single-atom catalysts catalyzing non-oxidized coupling reactions of CH4 with ethylene and aromatics [35], in which no carbon deposition was observed. Theoretical calculations also show that the oxide-supported Pt single-atom catalyst can only efficiently activate the first C-H bond in CH4, matching well with the experimental results [36].

3.1.3. Second Active Metal

As a non-precious metal element, Ni shows high catalytic activity but low anti-deposition ability towards the DRM reaction process. While precious metals (Ru, Rh, Ir, Pd, Pt, etc.) own high activity and good carbon-resistant properties, they are expensive. Therefore, adding a small number of precious metals as co-active components [37] could achieve excellent activity and stability simultaneously with high cost-effectiveness [38]. In general, a large number of active Ni components in Ni-based catalysts migrate to the outside of the pore, causing Ni to sinter and forming large-sized particles. The addition of a second component, such as Pt, can greatly avoid the diffusion of active metals to the outside, decrease the size of metal particles, and reduce the sintering of metal particles in the pore. Recently, Li et al. [39] designed and constructed a highly dispersed Ni-Ir/MgAl2O4 alloy catalyst for the DRM reaction. Based on the synergistic effect between Ni, Ir, and MgAl2O4, high catalytic activity and stability were achieved concurrently. Ni is the main activation center of CH4, while the Ir can effectively utilize the surface carbonates formed by the adsorption of CO2 on the MgAl2O4 support to generate active oxygen species, which can further be used for the elimination of the carbon species generated after the activation of CH4 (Figure 4). Based on the equilibrium mechanism between the active site and carbon deposition, the dynamic balance between carbon deposition and elimination can be achieved by regulating the active site so that the carbon deposition in this system will not continue to increase with the increased reaction time. The optimal Ni3Ir6/MgAl2O4 catalyst achieved near-zero carbon deposition during a 600 h high-temperature DRM reaction [39].
In sum, based on the investigation of the influence of the Ni configuration, the size of the Ni particles, and doped metal components on catalyst stability, it is found that the effective dispersion of Ni and the avoidance of active components sintering can prolong the service life of catalysts.

3.2. Regulation of Supports

The support is generally inert in the DRM reaction. However, it can affect the stability and mechanical strength of the catalyst via the interaction between the active component and the support. Moreover, the surface area, pore structure, thermal stability, oxidation-reducibility, and surface alkalinity of the support can affect the carbon deposition behavior of the Ni-based catalyst [40].
Ni-based DRM catalysts generally use porous oxides as supports, such as metal oxides γ-Al2O3 [12,41], SiO2 [40,42], MgO [43], ZrO2 [11,44], CeO2 [45] with rich acidic/alkali sites or good reducibility [46]. The structure, acidity, and alkalinity of the support, defects, and the use of the new catalyst support will affect the carbon deposition [40,42,47,48].

3.2.1. Support Structure

As described above, the support can affect the particle size and dispersion of the active metal in the catalyst and directly affect the activity and stability of the DRM catalysts. For the Ni-based catalysts, the use of a support with a high specific surface area and a finely controlled pore structure can obtain small Ni nanoparticles, improve the dispersion of the active component, prevent the sintering of metal particles, and reduce carbon deposition [40]. Al2O3-supported Ni-based catalysts, one of the widely employed DRM catalysts, exhibit a high catalytic activity because of their good structural characteristics, strong metal–support interaction (SMSI) effect, and high dispersion of Ni particles. The porous 3D structure enables the Ni nanoparticles to remain uniformly dispersed in the catalyst. In addition, the formation of the spinel structure NiAl2O4 during DRM makes the catalyst highly stable. However, the generation of the NiAl2O4 phase also limits the metal utilization of the active Ni site, thereby reducing the DRM reactivity. The formed NiAl2O4 spinel is unstable and can be reduced to Ni during the DRM reaction by a reductive gas mixture of CO and H2. After adding an appropriate amount of MgO to the Ni@Al2O3 catalyst, the spinel structure MgAl2O4 generated during the reaction also contributes to the resistance to carbon deposition [49].
ZrO2 is another commonly used support due to its high oxygen storage capacity and high thermodynamic stability [50]. The crystal structure of ZrO2 is closely related to the catalytic activity and stability of the catalysts [21]. Naeem et al. [51] prepared Ni@Al2O3, Ni@CeO2, and Ni@ZrO2 using polyol and surfactant-assisted methods. Among these catalysts, Ni@ZrO2 prepared using the polyol method delivers the highest reactivity, while the Ni@ZrO2 obtained using the surfactant-assisted method shows the lowest reactivity. This difference in catalytic performance is due to the phase state of ZrO2. The ZrO2 of the monoclinic crystal system was mainly prepared using the polyol method, while the ZrO2 of the cubic crystal system was mainly prepared using the surfactant-assisted method [51]. The former shows better redox properties than the latter and favors the adsorption and activation of CO2 better, leading to higher reactivity [51].
The distribution of active metal components at different locations of the support also affects carbon deposition. For example, Yan et al. [52] prepared and designed two types of Ni/CeO2-SiO2 catalysts, including CeO2 (Ni/CeO2-SiO2-P) in close contact with Ni nanoparticles or CeO2 (Ni/CeO2-SiO2-C) away from the Ni nanoparticles for DRM. Compared with Ni/CeO2-SiO2-C, Ni/CeO2-SiO2-P exhibits superior low-temperature activity and H2/CO ratio [52]. At 700 °C, the CO2 and CH4 conversion rates of the former (87.3% and 78.5%) were higher than those of the latter (80.5% and 67.8%). Ni/CeO2-SiO2-P was stable in long-term tests, while Ni/CeO2-SiO2-C showed poor stability, and the activity was significantly reduced within 10 h. The improved performance and stability of the Ni/CeO2-SiO2-P resulted from rich reactive oxygen species and more accessible formate species located at the metal-support interface [52] (Figure 5).

3.2.2. Acidity and Alkalinity of the Support Surface

In the high-temperature DRM process, the support affects the acidity and alkalinity of the catalyst, and the change in acidity and alkalinity further affects the reaction mechanism and carbon deposition. For Ni-based catalysts, the support properties affect the adsorption and dissociation of CO2 and CH4 on the catalyst and determine the reaction pathway of DRM [53]. If the inert support (SiO2) is used as the catalyst support, the reaction follows a single-function pathway, that is, CH4 and CO2 are adsorbed and activated on the surface of the active metal, and the inert support is only used as the skeleton of the catalyst, providing large pore volume, good thermal stability, and good dispersibility for the active metal [42,53]. CO2 is an acid gas, and using an alkaline support may lead to an increase in the adsorption capacity of CO2. Introducing the Lewis alkali support reduces the formation of deposited carbon and improves stability [54]. As illustrated in Reference [51], the Ni@ZrO2 prepared using the surfactant-assisted method with high basicity shows better coke-resistant ability but lower reactivity compared with the one prepared by the surfactant-assisted method with low basicity. However, the basic support decreases the adsorption of CH4 and may reduce the DRM reaction activity. The acidic support reduces the adsorption of CO2 but promotes the CH4 adsorption, which may favor the generation of side products to result in increased carbon deposits. When the weakly acidic or weakly basic (CeO2, La2O3) support is employed as the support of the Ni-based catalyst, DRM takes place via a bifunctional reaction pathway, that is, CH4 is activated on the active metal, and CO2 is activated on the support or interface [55,56,57]. Alkaline additives, such as alkali metals, are used to increase the surface alkalinity of the catalyst, thereby accelerating the activation of CO2 and the gasification of carbon [58].

3.2.3. Support Defects

Surface defects of metal oxides provide a way to stabilize thermodynamically unstable metal atoms and as active sites to promote adsorption and chemical transformation of reactants [21].
For non-redox support Al2O3, the surface defects are also used in regulating the catalytic reactivity of reactions. In general, the so-called defect sites on Al2O3 are mainly related to the coordination of the unsaturated Al site and the adjacent lattice O atoms. The under-coordinately tricoordinated Al3+ sites are closely related to surface dehydroxylation. Zhang et al. [59] proposed a new comprehension of the special role of defect sites on γ-Al2O3. Ni/CeO2-SiO2 catalyst was prepared using the plasma decomposition method, which is characterized by close contact between the CeO2 and Ni nanoparticles. In previous work, plasma-decomposed nickel catalysts showed smaller nickel particle sizes, fewer defect sites, and stronger nickel-support interactions compared to thermal decomposed nickel catalysts. The development of the acid etching strategy has been applied to the directional removal of OH groups on the surface of the γ-Al2O3, thereby generating the necessary vacancy defects to anchor the active Ni site, and the moderate interaction of Ni with the modified γ-Al2O3 support has improved the utilization of Ni. These designed defects promote the adsorption and activation of CO2 and the activation of CH4, thus improving the stability and activity of traditional Ni/Al2O3 catalysts [59].
Redox oxide supports, such as CeO2 and TiO2, have a high concentration of highly reactive oxygen species and unique redox properties (M4+↔M3+) [60,61,62,63,64,65,66]. The surface defects anchor Ni, forming a strong interaction between the metal and the support, preventing the metal from sintering. Reactive oxygen species can act as local sources or reservoirs of oxygen species, which are involved in reactions that occur on oxide surfaces [23]. The effective barrier for methane activation decreases to 0.15 eV on Ni/CeO2−x(111) compared with 0.9 eV on Ni(111) [67]. Recent studies show that Co-substituted CeO2 is more stable than Ni-substituted CeO2 under the DRM reaction conditions at 800 °C. Surface lattice oxygen reactivity plays a vital role in catalytic stability. The vacancy formation is favored in Co-substituted CeO2 as compared to Ni-substituted CeO2, leading to a higher availability of surface oxygen in Co-substituted catalyst than Ni-substituted CeO2 for the reaction [68]. Interfacial metal-Olattice favors the O-assisted CH4 dissociation, forms CHxO intermediate, and finally yields CO. The reacted Olattice can be re-supplied by CO2 dissociation at the metal-CeO2 interface [69].
It is worth pointing out that recent works mostly use combined strategies to regulate catalyst activity and stability jointly. For example, Li et al. [70] claim that the high Ni dispersion, intensified Ni–support interaction, the enlarged oxygen vacancies, the increased t-ZrO2 content, and enhanced reducibility of NiO led by oxygen vacancies co-boost the superior catalytic activity and stability of the developed LA-Ni/ZrO2 catalyst. Although a larger amount of coke is deposited on the spent LA-Ni/ZrO2 catalyst in comparison with that on the spent Ni/ZrO2, the developed LA-Ni/ZrO2 illustrates much higher catalytic stability than Ni/ZrO2, ascribed to the superior thermal sintering resistance of Ni nanoparticles and the different coke morphologies induced by enforced interaction of Ni and the ZrO2 support [70]. Similarly, among a series of Ni-CexZr1−xO2 and Ni-CexTi1−xO2 catalysts with varying Ce/Zr and Ce/Ti molar ratios (x = 0, 0.25, 0.5, 0.75, and 1), Ni-Ce0.5Zr0.5O2 and Ni-Ce0.5Ti0.5O2 delivered the highest apparent activity and stability for combined steam and CO2 reforming of CH4 in each series due to small Ni particles, strong metal–support interaction, and high oxygen vacancy concentration [71]. Ding et al. [72] also find that doping the optimized amount of cerium (Ce/Mg = 0.12) into Ni/MgO catalyst can improve activity and stability, which are attributed to the easy reducibility of NiOx species, small Ni particle size, and abundant oxygen vacancies resulting from the doping of cerium.

3.2.4. New Catalyst Support

Employing novel materials, such as MXene, as the support of Ni-based catalysts has also received much attention. For example, MXene is a class of layered two-dimensional (2D) transition metal carbides [73]. Due to its large specific surface area, high conductivity and strong redox properties, and the presence of hydroxyl groups or terminal oxygen on the surface, the multilayer structure of V2CTxMXene maintains a certain degree of extension in the environment of inert gas (N2) and reducing gas (H2). And in the presence of CO2 and air, V2CTxMXene can be completely oxidized (V2CTx is converted to V2O5) [74]. After 100 h of reaction, its stability is still very high; no carbon deposition has been found, showing excellent resistance towards carbon deposition. It has the potential to be used as a precursor, support, or catalyst for the DRM reaction [74]. Another 2D support is nitride, such as BN. Because BN has the advantages of high thermal stability, strong chemical inertness, good thermal conductivity, and corrosion resistance, it shows greater application potential than traditional oxides under harsh reaction environments. As expected, BN-supported Ni-based catalysts show good resistance to carbon deposition during DRM [19]. However, due to the inert surface of BN, Ni particles are prone to sintering on the surface of the BN, which leads to an unstable Ni-BN interface. Based on this, a research group first reduced the particle size of Ni particles by reducing the Ni loading (0.3 wt%) to increase the proportion of Ni-BN interfaces and then confined the Ni particles to the edge of BN by functionalizing amorphous SiO2 generated by functional group decomposition on boron nitride (BNf), thereby constructing a stable Ni-BN interface. The results show that Ni/BNf shows excellent resistance to carbon deposition during DRM, and Ni remains in a metallic state after reaction for 10 h. For comparison, under the same DRM conditions, no carbon deposition is observed on the surface of the Ni/BNf catalyst, while a few deposited carbons are produced on the surface of the Ni/α-Al2O3 catalyst. Additionally, under the condition of the CH4 pyrolysis experiment alone, no obvious carbon deposition is observed on the Ni/BNf catalyst (Figure 6), while carbon deposition on the Ni/α-Al2O3 catalyst surface continues to occur, and deposited carbon covered all surfaces of the catalyst. In situ tests and theoretical calculations show that the catalyst promotes the first C-H cleavage and inhibits the last C-H bond cleavage of CH4. When CH4 and CO2 are injected at the same time, CO2 preferentially activates on the Ni surfaces. Altogether, it makes the catalyst present excellent resistance to carbon deposition [19].
In sum, the carbon deposition behavior of the Ni-based catalyst can be adjusted by the pore structure, oxidation-reducibility, and surface alkalinity of the support.

3.3. Interface Structure Control

The interfacial structure is widely confirmed to influence the properties of the composites [75], such as the resistance towards carbon deposition during Ni-based catalysts catalyzing DRM [76]. Recently, studies have shown that adjusting the interface and strengthening the interaction between Ni and the support play a vital role in the synthesis of efficient and stable Ni-based DRM catalysts. Agglomeration of Ni nanoparticles can be avoided by regulating the SMSI between the metal (Ni) and the support. Compared to bare Ni NPs, the greater the contribution of the Ni–support interface, the stronger the adsorption or activation of the reactants [77]. Stronger Ni–support interactions yield more uniform distribution and smaller Ni nanoparticles, which is conducive to high activity and stability. Jeon et al. [76] propose the strategy of suppressing Ni nanoparticle aggregation during the reduction process via in situ reduction using dry methane (CH4/CO2) as the reducing gas. It increases the exposed interface between the Ni and CeO2 and results in large amounts of the Ni3+ phase at the catalyst surface, of which SMSI is dramatically enhanced (Figure 7). SMSI originates from Ni-doped sites on the CeO2 (denoted as the Ni3+ phase). The more Ni3+ on the surface, the more SMSI is formed near the exposed Ni nanoparticles, the smaller the size of Ni nanoparticles prepared, and the better the durability and activity. The developed catalyst achieved a methane conversion of 60% for 550 h at a low DRM operating temperature of 600 °C [76].

3.4. Special Structure Design

The configuration of the Ni centers, size of Ni particles, and exposed crystal surface are the key factors affecting its catalytic performance. Because Ni particles are on the surface of the catalyst and are exposed to the reaction stream directly, the supported Ni nanocatalysts prepared by conventional impregnation or precipitation migrate and aggregate during the DRM. Therefore, the design of special structures [78,79,80,81,82], e.g., core-shell [24], to protect metal active centers is the key to improving the stability of the catalysts. Employing this strategy, researchers improve the resistance ability of Ni-based DRM catalysts towards sintering and carbon deposition.

3.4.1. Reverse Metal-on-Oxide

Recently, some research teams innovatively constructed nanostructured oxides on the surface of metals to form reverse oxide-on-metal catalysts [44,64]. Due to the interaction between the oxide and the metal, the oxide formed on the metal surface is a single-layer dispersed, metastable structure, and metal active centers with coordination unsaturated are formed at the oxide-metal boundary, which can efficiently activate O2, H2O, and other molecules, and can be successfully used in low-temperature catalytic oxidation, etc. Another group has successfully grown nano-ZrO2 films on Pt, Cu, Pd, and Ni, denoted as ZrO2@M (M = Ni, Pt, Cu, Pd), using chemical vapor deposition and applying them to the DRM reaction. This new type of reverse catalyst shows excellent resistance to carbon deposition. It was found that the activity and stability of DRM are closely related to Zr0. Graphitic deposits are anti-segregated into Ni0 nanoparticles to provide restored CH4 adsorption sites and near-surface/dissolved C atoms, which migrate toward the Ni0/ZrO2 interface and induce local ZrxCy formation. The resulting oxygen-deficient carbidic phase boundary sites assist in kinetically enhanced CO2 activation toward CO [44,64].

3.4.2. Encapsulated Structure

Encapsulating active metal nanoparticles with a protecting shell, such as a carbon layer or zeolite [83], is an effective way to increase their stability. For Ni-based catalysts, using the zeolite with a special pore structure as the protector is widely employed so that the migration of Ni is inhibited effectively, thereby improving the resistance to carbon deposition. For example, based on this strategy, researchers encapsulate active Ni nanoparticles by the silicon-aluminum MFI zeolite (Ni@HZSM-5) [83]. They further regulate the pore environment of zeolite to enhance the hydrogen spillover so as to keep enough active hydrogen species available around Ni nanoparticles [83]. This facilitates the reduction of CO2 to CO and suppresses the side reaction of RWGS. In a CO2-rich atmosphere, the methane on the catalytic material can reduce CO2 by 2.9 (1 mol of methane reduces 2.9 mol CO2), which is better than the most advanced super-DRM process reported so far. The CO yield of 3.9 molCO molCH4−1 on this catalytic material can be carried out for a long time. It is expected to realize the direct conversion and utilization of CO2-rich natural gas [83].
Similarly, Wang et al. [84] prepared small-sized bimetallic NiCo nanoparticles encapsulated in silicalite-2 (S-2) using the in situ growth method. Elemental segregation is observed during the activation process, in which Ni and Co atoms migrate to the outer surface and center of the NiCo alloy, respectively [84]. The optimal Ni0.2Co0.3@S-2 catalyst shows stable and high CH4 and CO2 conversion within 100 h without carbon deposition [84]. This localized effect, combined with the precise dynamic balance of carbon and oxygen content on the catalyst surface, makes it have high catalytic performance [84].
In another example, a catalyst of attapulgite-derived MFI (ADM) zeolite-encapsulated Ni-Co alloys was fabricated using a one-pot method. Co in Ni–Co alloys transfers electrons to Ni, forming electron-rich Ni metal sites, which significantly improves the ability of C-H bond cleavage [47]. Meanwhile, the ADM zeolite not only stabilized metallic sites by pore or layer confinement but also offered massive CO2 adsorption/activation centers. These improve the ability of the anti-sintering and anti-coking formation as well as the superior DRM activity of the optimal 10Ni1Co@ADM-0.1 catalyst [47].
Active metal nanoparticles can also be encapsulated using surface coating oxides (e.g., SiO2, >10 nm). He et al. [85] grew Ni nanoparticles in situ from the MgO–Al2O3 matrix and surrounded by a thin layer of 1–2 nm multielement oxide (Ni@NiSiAlMgOx), and synchronously controlled the size, exposed crystal planes, surface acidity and alkalinity, and redox state of the Ni nanoparticles, thereby improving the stability of Ni catalysts in DRM reactions (Figure 8). The oxide thin layer efficiently isolated and confined the Ni nanoparticles, avoiding fast C–C bond formation on an open facet and protecting the Ni nanoparticles from sintering [85]. The appropriately sized Ni particles of around 15 nm balanced the carbon decomposition and elimination rates. The nearby basic sites assist CO2 adsorption and accelerate carbon elimination. The cooperative metallic Ni and the surficial MEO thin layer containing Ni2+ enhanced the anti-sintering of Ni@NiSiAlMgOx in a redox atmosphere of MDR [85].
As summarized in Table 2, the CO:H2 mole ratio is less than 1 due to the presence of side reaction, the conversion rate of CO2 is usually higher than that of CH4, and the stability can reach 550 h. In addition to the above regulation strategies, the stability of the catalyst can also be improved using other strategies, which are not discussed in this review in detail.

4. Conclusions and Outlook

The rapid carbon deposition and catalyst sintering induced by high reaction temperature limit the industrial applications of Ni-based catalysts toward DRM reaction. In this review, the mechanisms of reaction and deactivation during DRM reaction are analyzed, and the influencing factors and possible solutions regarding the stability of Ni-based catalysts are summarized. The catalytic activity and stability are co-affected by the active metal, support, interface, etc. The stability of the catalyst can be improved by multi-scale regulation. The metal dispersion is related to stability, which has an optimal limit. Increasing the surface basicity of the support could favor the activation of CO2 and consequently inhibit carbon deposition on the catalyst. The protection of active centers by the construction of special structures, e.g., encapsulating Ni by oxides, can also improve the stability of catalysts.
Future research can focus on the following aspects: (1) Develop new preparation methods of catalysts to increase the dispersion of Ni particles on the support and adjust the interaction between the metal and the support; (2) increase the number of adsorbed oxygen species on the catalyst surface by the release of partial lattice oxygen, and promote the combination of adsorbed oxygen with adsorbed carbon to form CO, eliminating carbon deposition on the catalyst surface; (3) couple to the reaction of CO hydrogenation, modifying the reaction thermodynamically.
This review provides guidance for the rational design of stable Ni-based catalysts in the industrial application of DRM.

Author Contributions

Conceptualization, M.W. and X.S.; formal analysis, M.W.; investigation, M.W. and X.S.; resources, M.W. and X.S.; writing—original draft preparation, M.W.; writing—review and editing, X.S.; supervision, X.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Shanghai Natural Science Foundation, grant number 23ZR1425500, the foundation of the National Key Laboratory of High Efficiency and Low Carbon Utilization of Coal, grant number Grant No. J23-24-603, Shanghai Local Universities Capacity Building Project of Science and Technology Innovation Action Program, grant number 21010501700, Class III Peak Discipline of Shanghai—Materials Science and Engineering (High-Energy Beam Intelligent Processing and Green Manufacturing).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Meng, C.; Zhao, G.; Shi, X.-R.; Chen, P.; Liu, Y.; Lu, Y. Oxygen-deficient metal oxides supported nano-intermetallic InNi3C0.5 toward efficient CO2 hydrogenation to methanol. Sci. Adv. 2021, 7, eabi6012. [Google Scholar] [CrossRef] [PubMed]
  2. Meng, C.; Zhao, G.; Shi, X.-R.; Nie, Q.; Liu, Y.; Lu, Y. Electronic modulation of InNi3C0.5/Fe3O4 by support precursor toward efficient CO2 hydrogenation to methanol. Appl. Catal. B 2022, 316, 121699. [Google Scholar] [CrossRef]
  3. Zhu, J.; Zhao, G.; Meng, C.; Chen, P.; Shi, X.-R.; Lu, Y. Superb Ni-foam-structured nano-intermetallic InNi3C0.5 catalyst for hydrogenation of dimethyl oxalate to ethylene glycol. Chem. Eng. J. 2021, 426, 130857. [Google Scholar] [CrossRef]
  4. Chen, P.; Zhao, G.; Shi, X.-R.; Zhu, J.; Ding, J.; Lu, Y. Nano-Intermetallic InNi3C0.5 Compound Discovered as a Superior Catalyst for CO2 Reutilization. iScience 2019, 17, 315–324. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, M.; Wu, Z.; Kong, X.; Zhang, X.; Tan, L.; Guo, H.; Li, C.-J. One-pot synthesis of toluene from methane and methanol catalyzed by GaN nanowire. Nano Res. 2023, 16, 6512–6516. [Google Scholar] [CrossRef]
  6. Zhang, X.; Witte, J.; Schildhauer, T.; Bauer, C. Life cycle assessment of power-to-gas with biogas as the carbon source. Sustain. Energy Fuels 2020, 4, 1427–1436. [Google Scholar] [CrossRef]
  7. Palmer, C.; Upham, D.C.; Smart, S.; Gordon, M.J.; Metiu, H.; McFarland, E.W. Dry reforming of methane catalysed by molten metal alloys. Nat. Catal. 2020, 3, 83–89. [Google Scholar] [CrossRef]
  8. Kawi, S.; Kathiraser, Y.; Ni, J.; Oemar, U.; Li, Z.; Saw, E.T. Progress in Synthesis of Highly Active and Stable Nickel-Based Catalysts for Carbon Dioxide Reforming of Methane. ChemSusChem 2015, 8, 3556–3575. [Google Scholar] [CrossRef]
  9. Sheng, K.; Luan, D.; Jiang, H.; Zeng, F.; Wei, B.; Pang, F.; Ge, J. NixCoy Nanocatalyst Supported by ZrO2 Hollow Sphere for Dry Reforming of Methane: Synergetic Catalysis by Ni and Co in Alloy. ACS Appl. Mater. Interfaces 2019, 11, 24078–24087. [Google Scholar] [CrossRef]
  10. Yu, Y.-X.; Yang, J.; Zhu, K.-K.; Sui, Z.-J.; Chen, D.; Zhu, Y.-A.; Zhou, X.-G. High-Throughput Screening of Alloy Catalysts for Dry Methane Reforming. ACS Catal. 2021, 11, 8881–8894. [Google Scholar] [CrossRef]
  11. Dou, J.; Zhang, R.; Hao, X.; Bao, Z.; Wu, T.; Wang, B.; Yu, F. Sandwiched SiO2@Ni@ZrO2 as a coke resistant nanocatalyst for dry reforming of methane. Appl. Catal. B 2019, 254, 612–623. [Google Scholar] [CrossRef]
  12. Li, Y.; Wang, Q.; Cao, M.; Li, S.; Song, Z.; Qiu, L.; Yu, F.; Li, R.; Yan, X. Structural evolution of robust Ni3Fe1 alloy on Al2O3 in dry reforming of methane: Effect of iron-surplus strategy from Ni1Fe1 to Ni3Fe1. Appl. Catal. B 2023, 331, 122669. [Google Scholar] [CrossRef]
  13. Li, J.; Croiset, E.; Ricardez-Sandoval, L. Effect of Metal–Support Interface During CH4 and H2 Dissociation on Ni/γ-Al2O3: A Density Functional Theory Study. J. Phys. Chem. C 2013, 117, 16907–16920. [Google Scholar] [CrossRef]
  14. Xu, Y.; Li, J.; Jiang, F.; Xu, Y.; Liu, B.; Liu, X. Insight into the anti-coking ability of NiM/SiO2 (M=ZrO2, Ru) catalyst for dry reforming of CH4 to syngas. Int. J. Hydrogen Energy 2022, 47, 2268–2278. [Google Scholar] [CrossRef]
  15. Wang, Y.; Yao, L.; Wang, Y.; Wang, S.; Zhao, Q.; Mao, D.; Hu, C. Low-Temperature Catalytic CO2 Dry Reforming of Methane on Ni-Si/ZrO2 Catalyst. ACS Catal. 2018, 8, 6495–6506. [Google Scholar] [CrossRef]
  16. Köpfle, N.; Götsch, T.; Grünbacher, M.; Carbonio, E.A.; Hävecker, M.; Knop-Gericke, A.; Schlicker, L.; Doran, A.; Kober, D.; Gurlo, A.; et al. Zirconium-Assisted Activation of Palladium To Boost Syngas Production by Methane Dry Reforming. Angew. Chem. Int. Ed. 2018, 57, 14613–14618. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, M.; Xiang, Y.; Chen, W.; Wu, S.; Zhu, Z.-Z.; Cao, X. SiFeN6-graphene: A promising dual-atom catalyst for enhanced CO2-to-CH4 conversion. Appl. Surf. Sci. 2024, 643, 158724. [Google Scholar] [CrossRef]
  18. Wang, P.; Shi, X.-R.; Zhang, Y.; Wei, M. The influence of Co nucleation in Co-Ni single atom alloy on low-temperature methane dry reforming with DFT simulations and microkinetic modeling. Mol. Catal. 2023, 549, 113515. [Google Scholar] [CrossRef]
  19. Deng, J.; Gao, M.; Hasegawa, J.; Zhang, X.; Wang, A.; Chen, A.; Zhang, D. Unravelling the Anomalous Coking Resistance over Boron Nitride-Supported Ni Catalysts for Dry Reforming of Methane. CCS Chem. 2023, 5, 2111–2124. [Google Scholar] [CrossRef]
  20. Chen, S.; Zaffran, J.; Yang, B. Descriptor Design in the Computational Screening of Ni-Based Catalysts with Balanced Activity and Stability for Dry Reforming of Methane Reaction. ACS Catal. 2020, 10, 3074–3083. [Google Scholar] [CrossRef]
  21. Shi, X.R.; Wang, P.; Jing, C.; Wu, K.; Xu, S.; Klötzer, B. The formation of O vacancy on ZrO2/Pd and its effect on methane dry reforming: Insights from DFT and microkinetic modeling. Appl. Surf. Sci. 2023, 619, 156679. [Google Scholar] [CrossRef]
  22. Das, S.; Bhattar, S.; Liu, L.; Wang, Z.; Xi, S.; Spivey, J.J.; Kawi, S. Effect of Partial Fe Substitution in La0.9Sr0.1 NiO3 Perovskite-Derived Catalysts on the Reaction Mechanism of Methane Dry Reforming. ACS Catal. 2020, 10, 12466–12486. [Google Scholar] [CrossRef]
  23. Laosiripojana, N.; Sutthisripok, W.; Assabumrungrat, S. Synthesis gas production from dry reforming of methane over CeO2 doped Ni/Al2O3: Influence of the doping ceria on the resistance toward carbon formation. Chem. Eng. J. 2005, 112, 13–22. [Google Scholar] [CrossRef]
  24. Das, S.; Ashok, J.; Bian, Z.; Dewangan, N.; Wai, M.H.; Du, Y.; Borgna, A.; Hidajat, K.; Kawi, S. Silica–Ceria sandwiched Ni core–shell catalyst for low temperature dry reforming of biogas: Coke resistance and mechanistic insights. Appl. Catal. B 2018, 230, 220–236. [Google Scholar] [CrossRef]
  25. Abdulrasheed, A.; Jalil, A.A.; Gambo, Y.; Ibrahim, M.; Hambali, H.U.; Hamid, M.Y.S. A review on catalyst development for dry reforming of methane to syngas: Recent advances. Renew. Sustain. Energy Rev. 2019, 108, 175–193. [Google Scholar] [CrossRef]
  26. Liang, Y.; Douthwaite, M.; Huang, X.; Zhao, B.; Tang, Q.; Liu, L.; Dong, J. Zero-oxidation state precursor assisted fabrication of highly dispersed and stable Pt catalyst for chemoselective hydrogenation of α,β-unsaturated aldehydes. Nano Res. 2023, 16, 6085–6093. [Google Scholar] [CrossRef]
  27. Wang, S.; Wang, Y.; Yao, L.; Hu, C. Temperature hysteresis in dry reforming of methane on Ni/SBA-15 catalyst: Low temperature activity originated from the synergy of Ni0 and carbon nanotubes. Appl. Catal. B 2023, 333, 122756. [Google Scholar] [CrossRef]
  28. Zhang, M.; Ding, L.; Zheng, J.; Liu, L.; Alsulami, H.; Kutbi, M.A.; Xu, J. Surface modification of carbon fibers with hydrophilic Fe3O4 nanoparticles for nickel-based multifunctional composites. Appl. Surf. Sci. 2020, 509, 145348. [Google Scholar] [CrossRef]
  29. Liu, C.; Shi, X.; Yue, K.; Wang, P.; Zhan, K.; Wang, X.; Xia, B.Y.; Yan, Y. S-Species-Evoked High-Valence Ni2+δ of the Evolved β-Ni(OH)2 Electrode for Selective Oxidation of 5-Hydroxymethylfurfural. Adv. Mater. 2023, 35, 2211177. [Google Scholar] [CrossRef]
  30. Rogers, J.L.; Mangarella, M.C.; D’Amico, A.D.; Gallagher, J.R.; Dutzer, M.R.; Stavitski, E.; Miller, J.T.; Sievers, C. Differences in the Nature of Active Sites for Methane Dry Reforming and Methane Steam Reforming over Nickel Aluminate Catalysts. ACS Catal. 2016, 6, 5873–5886. [Google Scholar] [CrossRef]
  31. Sumi, T.; Murata, D.; Kitamura, H.; Kubota, S.; Miyake, K.; Uchida, Y.; Miyamoto, M.; Nishiyama, N. Formation of Ni Species Anchored on a Silicalite-1 Zeolite Framework as a Catalyst with High Coke Deposition Resistance on Dry Reforming of Methane. Cryst. Growth Des. 2023, 23, 3308–3313. [Google Scholar] [CrossRef]
  32. Manabayeva, A.M.; Mäki-Arvela, P.; Vajglová, Z.; Martinéz-Klimov, M.; Tirri, T.; Baizhumanova, T.S.; Grigor’eva, V.P.; Zhumabek, M.; Aubakirov, Y.A.; Simakova, I.L.; et al. Dry Reforming of Methane over Ni–Fe–Al Catalysts Prepared by Solution Combustion Synthesis. Ind. Eng. Chem. Res. 2023, 62, 11439–11455. [Google Scholar] [CrossRef]
  33. Gu, Y.; Cao, Z.; Zhao, M.; Xu, Y.; Lu, N. Single-Atom Fe Nanozyme with Enhanced Oxidase-like Activity for the Colorimetric Detection of Ascorbic Acid and Glutathione. Biosensors 2023, 13, 487. [Google Scholar] [CrossRef] [PubMed]
  34. Akri, M.; El Kasmi, A.; Batiot-Dupeyrat, C.; Qiao, B. Highly Active and Carbon-Resistant Nickel Single-Atom Catalysts for Methane Dry Reforming. Catalysts 2020, 10, 630. [Google Scholar] [CrossRef]
  35. Guo, X.; Fang, G.; Li, G.; Ma, H.; Fan, H.; Yu, L.; Ma, C.; Wu, X.; Deng, D.; Wei, M.; et al. Direct, nonoxidative conversion of methane to ethylene, aromatics, and hydrogen. Science 2014, 344, 616–619. [Google Scholar] [CrossRef]
  36. Akri, M.; Zhao, S.; Li, X.; Zang, K.; Lee, A.F.; Isaacs, M.A.; Xi, W.; Gangarajula, Y.; Luo, J.; Ren, Y.; et al. Atomically dispersed nickel as coke-resistant active sites for methane dry reforming. Nat. Commun. 2019, 10, 5181. [Google Scholar] [CrossRef]
  37. Gould, T.D.; Montemore, M.M.; Lubers, A.M.; Ellis, L.D.; Weimer, A.W.; Falconer, J.L.; Medlin, J.W. Enhanced dry reforming of methane on Ni and Ni-Pt catalysts synthesized by atomic layer deposition. Appl. Catal. A Gen. 2015, 492, 107–116. [Google Scholar] [CrossRef]
  38. Wang, G.; Liang, Y.; Song, J.; Li, H.; Zhao, Y. Study on High Activity and Outstanding Stability of Hollow-NiPt@SiO2 Core–Shell Structure Catalyst for DRM Reaction. Front. Chem. 2020, 8, 220. [Google Scholar] [CrossRef]
  39. Li, H.; Hao, C.; Tian, J.; Wang, S.; Zhao, C. Ultra-durable Ni-Ir/MgAl2O4 catalysts for dry reforming of methane enabled by dynamic balance between carbon deposition and elimination. Chem. Catal. 2022, 2, 1748–1763. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Zhang, G.; Liu, J.; Li, T.; Wang, Y.; Zhao, Y.; Li, G.; Zhang, Y. Dry reforming of methane over Ni/SiO2 catalysts: Role of support structure properties. Fuel 2023, 340, 127490. [Google Scholar] [CrossRef]
  41. Fakeeha, A.H.; Ibrahim, A.A.; Arafat, Y.; Atia, H.; Abasaeed, A.E.; Al-Fatesh, A.S. Iridium promoted Ni-Co/Al2O3-ZrO2 catalyst for dry reforming of methane. Can. J. Chem. Eng. 2018, 96, 955–960. [Google Scholar] [CrossRef]
  42. Song, Q.; Ran, R.; Wu, X.; Si, Z.; Weng, D. Dry reforming of methane over Ni catalysts supported on micro- and mesoporous silica. J. CO2 Util. 2023, 68, 102387. [Google Scholar] [CrossRef]
  43. Zeng, F.; Zhang, J.; Xu, R.; Zhang, R.; Ge, J. Highly dispersed Ni/MgO-mSiO2 catalysts with excellent activity and stability for dry reforming of methane. Nano Res. 2022, 15, 5004–5013. [Google Scholar] [CrossRef]
  44. Haug, L.; Thurner, C.; Bekheet, M.F.; Ploner, K.; Bischoff, B.; Gurlo, A.; Kunz, M.; Sartory, B.; Penner, S.; Klötzer, B. Pivotal Role of Ni/ZrO2 Phase Boundaries for Coke-Resistant Methane Dry Reforming Catalysts. Catalysts 2023, 13, 804. [Google Scholar] [CrossRef]
  45. Odedairo, T.; Chen, J.; Zhu, Z. Metal–support interface of a novel Ni–CeO2 catalyst for dry reforming of methane. Catal. Commun. 2013, 31, 25–31. [Google Scholar] [CrossRef]
  46. Gao, X.; Lin, W.; Ge, Z.; Ge, H.; Kawi, S. Modification Strategies of Ni-Based Catalysts with Metal Oxides for Dry Reforming of Methane. Methane 2022, 1, 139–157. [Google Scholar] [CrossRef]
  47. Liang, D.; Wang, Y.; Chen, M.; Xie, X.; Li, C.; Wang, J.; Yuan, L. Dry reforming of methane for syngas production over attapulgite-derived MFI zeolite encapsulated bimetallic Ni-Co catalysts. Appl. Catal. B 2023, 322, 122088. [Google Scholar] [CrossRef]
  48. Chatla, A.; Almanassra, I.W.; Kallem, P.; Atieh, M.A.; Alawadhi, H.; Akula, V.; Banat, F. Dry (CO2) reforming of methane over zirconium promoted Ni-MgO mixed oxide catalyst: Effect of Zr addition. J. CO2 Util. 2022, 62, 102082. [Google Scholar] [CrossRef]
  49. Shang, Z.; Li, S.; Li, L.; Liu, G.; Liang, X. Highly active and stable alumina supported nickel nanoparticle catalysts for dry reforming of methane. Appl Catal. B 2017, 201, 302–309. [Google Scholar] [CrossRef]
  50. Zhang, J.; Lv, Z.; Wang, L.; Guo, Y. Preparation of C–N co-doped zirconia electrospun nanofibers and their humidity sensing properties. Mater. Adv. 2022, 3, 7265–7271. [Google Scholar] [CrossRef]
  51. Naeem, M.A.; Al-Fatesh, A.S.; Fakeeha, A.H.; Abasaeed, A.E. Hydrogen production from methane dry reforming over nickel-based nanocatalysts using surfactant-assisted or polyol method. Int. J. Hydrogen Energy 2014, 39, 17009–17023. [Google Scholar] [CrossRef]
  52. Yan, X.; Hu, T.; Liu, P.; Li, S.; Zhao, B.; Zhang, Q.; Jiao, W.; Chen, S.; Wang, P.; Lu, J.; et al. Highly efficient and stable Ni/CeO2-SiO2 catalyst for dry reforming of methane: Effect of interfacial structure of Ni/CeO2 on SiO2. Appl. Catal. B 2019, 246, 221–231. [Google Scholar] [CrossRef]
  53. Velty, A.; Corma, A. Advanced zeolite and ordered mesoporous silica-based catalysts for the conversion of CO2 to chemicals and fuels. Chem. Soc. Rev. 2023, 52, 1773–1946. [Google Scholar] [CrossRef]
  54. Kim, D.H.; Seo, J.-C.; Kim, Y.J.; Kim, J.; Yoon, S.; Ra, H.; Kim, M.-J.; Lee, K. Ni-Co alloy catalyst derived from NixCoy/MgAl2O4 via exsolution method for high coke resistance toward dry reforming of methane. Catal. Today 2024, 425, 114337. [Google Scholar] [CrossRef]
  55. Singh, S.; Zubenko, D.; Rosen, B.A. Influence of LaNiO3 Shape on Its Solid-Phase Crystallization into Coke-Free Reforming Catalysts. ACS Catal. 2016, 6, 4199–4205. [Google Scholar] [CrossRef]
  56. Wang, M.; Zhao, T.; Dong, X.; Li, M.; Wang, H. Effects of Ce substitution at the A-site of LaNi0.5Fe0.5O3 perovskite on the enhanced catalytic activity for dry reforming of methane. Appl. Catal. B 2018, 224, 214–221. [Google Scholar] [CrossRef]
  57. Bonmassar, N.; Bekheet, M.F.; Schlicker, L.; Gili, A.; Gurlo, A.; Doran, A.; Gao, Y.; Heggen, M.; Bernardi, J.; Klötzer, B.; et al. In Situ-Determined Catalytically Active State of LaNiO3 in Methane Dry Reforming. ACS Catal. 2020, 10, 1102–1112. [Google Scholar] [CrossRef]
  58. Zhang, Z.; Zhang, X.; Zhang, L.; Gao, J.; Shao, Y.; Dong, D.; Zhang, S.; Liu, Q.; Xu, L.; Hu, X. Impacts of alkali or alkaline earth metals addition on reaction intermediates formed in methanation of CO2 over cobalt catalysts. J. Energy Inst. 2020, 93, 1581–1596. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Zu, Y.; He, D.; Liang, J.; Zhu, L.; Mei, Y.; Luo, Y. The tailored role of “defect” sites on γ-alumina: A key to yield an efficient methane dry reforming catalyst with superior nickel utilization. Appl. Catal. B 2022, 315, 121539. [Google Scholar] [CrossRef]
  60. Köpfle, N.; Ploner, K.; Lackner, P.; Götsch, T.; Thurner, C.; Carbonio, E.; Hävecker, M.; Knop-Gericke, A.; Schlicker, L.; Doran, A.; et al. Carbide-Modified Pd on ZrO2 as Active Phase for CO2-Reforming of Methane—A Model Phase Boundary Approach. Catalysts 2020, 10, 1000. [Google Scholar] [CrossRef]
  61. Mayr, L.; Shi, X.; Köpfle, N.; Klötzer, B.; Zemlyanov, D.Y.; Penner, S. Tuning of the copper–zirconia phase boundary for selectivity control of methanol conversion. J. Catal. 2016, 339, 111–122. [Google Scholar] [CrossRef]
  62. Köpfle, N.; Mayr, L.; Lackner, P.; Schmid, M.; Schmidmair, D.; Götsch, T.; Penner, S.; Kloetzer, B. Zirconium-Palladium Interactions during Dry Reforming of Methane. ECS Trans. 2017, 78, 2419–2430. [Google Scholar] [CrossRef]
  63. Bikaljevic, D.; Rameshan, R.; Köpfle, N.; Götsch, T.; Mühlegger, E.; Schlögl, R.; Penner, S.; Memmel, N.; Klötzer, B. Structural and kinetic aspects of CO oxidation on ZnOx-modified Cu surfaces. Appl. Catal. A Gen. 2019, 572, 151–157. [Google Scholar] [CrossRef]
  64. Haug, L.; Thurner, C.; Bekheet, M.F.; Bischoff, B.; Gurlo, A.; Kunz, M.; Sartory, B.; Penner, S.; Klötzer, B. Zirconium Carbide Mediates Coke-Resistant Methane Dry Reforming on Nickel-Zirconium Catalysts. Angew. Chem. Int. Ed. 2022, 61, e202213249. [Google Scholar] [CrossRef]
  65. Thalinger, R.; Stöger-Pollach, M.; Hetaba, W.; Feuerbacher, M.; Klötzer, B.; Penner, S. Electron microscopy investigations of metal-support interaction effects in M/Y2O3 and M/ZrO2 thin films (M = Cu, Ni). Mater. Chem. Phys. 2013, 143, 167–177. [Google Scholar] [CrossRef]
  66. Wolfbeisser, A.; Klötzer, B.; Mayr, L.; Rameshan, R.; Zemlyanov, D.; Bernardi, J.; Föttinger, K.; Rupprechter, G. Surface modification processes during methane decomposition on Cu-promoted Ni–ZrO2 catalysts. Catal. Sci. Technol. 2015, 5, 967–978. [Google Scholar] [CrossRef]
  67. Liu, Z.; Grinter, D.C.; Lustemberg, P.G.; Nguyen-Phan, T.; Zhou, Y.; Luo, S.; Waluyo, I.; Crumlin, E.J.; Stacchiola, D.J.; Zhou, J.; et al. Dry Reforming of Methane on a Highly-Active Ni-CeO2 Catalyst: Effects of Metal-Support Interactions on C−H Bond Breaking. Angew. Chem. 2016, 128, 7581–7585. [Google Scholar] [CrossRef]
  68. Yadav, P.K.; Patrikar, K.; Mondal, A.; Sharma, S. Ni/Co in and on CeO2: A comparative study on the dry reforming reaction. Sustain. Energy Fuels 2023, 7, 3853–3870. [Google Scholar] [CrossRef]
  69. Shen, D.; Li, Z.; Shan, J.; Yu, G.; Wang, X.; Zhang, Y.; Liu, C.; Lyu, S.; Li, J.; Li, L. Synergistic Pt-CeO2 interface boosting low temperature dry reforming of methane. Appl. Catal. B 2022, 318, 121809. [Google Scholar] [CrossRef]
  70. Li, W.; Zhao, Z.; Ding, F.; Guo, X.; Wang, G. Syngas Production via Steam–CO 2 Dual Reforming of Methane over LA-Ni/ZrO2 Catalyst Prepared by L-Arginine Ligand-Assisted Strategy: Enhanced Activity and Stability. ACS Sustain. Chem. Eng. 2015, 3, 3461–3476. [Google Scholar] [CrossRef]
  71. Zhao, Y.; Qi, L.; Cheng, Z.; Zhou, Z. Syngas Production via Combined Steam and Carbon Dioxide Reforming of Methane over Ni-CexM1−xO2(M = Ti or Zr) Catalysts. Ind. Eng. Chem. Res. 2022, 61, 12978–12988. [Google Scholar] [CrossRef]
  72. Ding, X.; Yang, Y.; Li, Z.; Huang, P.; Liu, X.; Guo, Y.; Wang, Y. Engineering a Nickel–Oxygen Vacancy Interface for Enhanced Dry Reforming of Methane: A Promoted Effect of CeO2 Introduction into Ni/MgO. ACS Catal. 2023, 13, 15535–15545. [Google Scholar] [CrossRef]
  73. Liu, J.; Wang, F.; Guo, R.; Liu, Y.; Zhang, M.; Sunarso, J.; Liu, D. Size-regulated co-doped hetero-interfaced 3D honeycomb MXene as high performance electromagnetic absorber with anti-corrosion performance. Appl. Surf. Sci. 2023, 645, 158846. [Google Scholar] [CrossRef]
  74. Wysocka, I.; Karczewski, J.; Gołąbiewska, A.; Łapiński, M.; Cieślik, B.M.; Maciejewski, M.; Kościelska, B.; Rogala, A. Nickel phase deposition on V2CTx/V2AlC as catalyst precursors for a dry methane reforming: The effect of the deposition method on the morphology and catalytic activity. Int. J. Hydrogen Energy 2023, 48, 10922–10940. [Google Scholar] [CrossRef]
  75. Shi, X.-R.; Zhang, Y.; Zong, S.; Gu, W.; Ma, P.; Lu, N. Interaction of Ethylene with Irn (n = 1–10): From Bare Clusters to γ-Al2O3-Supported Nanoparticles. Nanomaterials 2019, 9, 331. [Google Scholar] [CrossRef]
  76. Jeon, O.S.; Lee, H.; Lee, K.S.; Paidi, V.K.; Ji, Y.; Kwon, O.C.; Kim, J.P.; Myung, J.H.; Park, S.Y.; Yoo, Y.J.; et al. Harnessing Strong Metal-Support Interaction to Proliferate the Dry Reforming of Methane Performance by in Situ Reduction. ACS Appl. Mater. Interfaces 2022, 14, 12140–12148. [Google Scholar] [CrossRef]
  77. Sun, Y.; Liu, F.; Wang, X.; Lu, K.; Liu, X.; Huang, Y.; Yu, F.; Chen, Y. Highly selective CO2 electroreduction to CO by the synergy between Ni–N–C and encapsulated Ni nanoparticles. Dalton Trans. 2023, 52, 928–935. [Google Scholar] [CrossRef]
  78. Zhang, R.; Lu, N.; Zhang, J.; Yan, R.; Li, J.; Wang, L.; Wang, N.; Lv, M.; Zhang, M. Ultrasensitive aptamer-based protein assays based on one-dimensional core-shell nanozymes. Biosens. Bioelectron. 2020, 150, 111881. [Google Scholar] [CrossRef]
  79. He, X.; Xie, S.; Xu, J.; Yin, X.-B.; Zhang, M. Reactive Template-Engaged Synthesis of NiSx/MoS2 Nanosheets Decorated on Hollow and Porous Carbon Microtubes with Optimal Electronic Modulation toward High-Performance Enzyme-like Performance. Inorg. Chem. 2023, 62, 8033–8042. [Google Scholar] [CrossRef]
  80. Yang, L.; Jin, Z.; Zheng, J.; Zhang, B.; Xu, J.; Yin, X.-B.; Zhang, M. In Situ Construction of Co-MoS2/Pd Nanosheets on Polypyrrole-Derived Nitrogen-Doped Carbon Microtubes as Multifunctional Catalysts with Enhanced Catalytic Performance. Inorg. Chem. 2022, 61, 542–553. [Google Scholar] [CrossRef]
  81. Yang, L.; He, X.; Xu, J.; Yin, X.-B.; Zhang, M. Construction of hierarchical Co, Fe Co-doped MoS2/polypyrrole heterostructure as an efficient bifunctional catalyst for enzyme-like and nitrophenol catalysis. Colloids Surf. A Physicochem. Eng. Asp. 2022, 655, 130225. [Google Scholar] [CrossRef]
  82. Xiao, Y.; Gong, W.; Zhao, M.; Zhang, M.; Lu, N. Surface-engineered prussian blue nanozymes as artificial receptors for universal pattern recognition of metal ions and proteins. Sens. Actuators B Chem. 2023, 390, 134006. [Google Scholar] [CrossRef]
  83. Zhu, Q.; Zhou, H.; Wang, L.; Wang, L.; Wang, C.; Wang, H.; Fang, W.; He, M.; Wu, Q.; Xiao, F.-S. Enhanced CO2 utilization in dry reforming of methane achieved through nickel-mediated hydrogen spillover in zeolite crystals. Nat. Catal. 2022, 5, 1030–1037. [Google Scholar] [CrossRef]
  84. Wang, J.; Fu, Y.; Kong, W.; Li, S.; Yuan, C.; Bai, J.; Chen, X.; Zhang, J.; Sun, Y. Investigation of Atom-Level Reaction Kinetics of Carbon-Resistant Bimetallic NiCo-Reforming Catalysts: Combining Microkinetic Modeling and Density Functional Theory. ACS Catal. 2022, 12, 4382–4393. [Google Scholar] [CrossRef]
  85. He, L.; Li, M.; Li, W.-C.; Xu, W.; Wang, Y.; Wang, Y.-B.; Shen, W.; Lu, A.-H. Robust and Coke-free Ni Catalyst Stabilized by 1–2 nm-Thick Multielement Oxide for Methane Dry Reforming. ACS Catal. 2021, 11, 12409–12416. [Google Scholar] [CrossRef]
Figure 1. Illustration of DRM.
Figure 1. Illustration of DRM.
Methane 03 00006 g001
Figure 2. Carbon formation, deposition, and transformation pathways on the Ni-based catalyst [25]. Reprinted with permission from Elsevier, copyright 2019.
Figure 2. Carbon formation, deposition, and transformation pathways on the Ni-based catalyst [25]. Reprinted with permission from Elsevier, copyright 2019.
Methane 03 00006 g002
Figure 3. Schematic diagram of the variation in catalysts in the DRM hysteresis reaction under different conditions [27]. Reprinted with permission from Elsevier, copyright 2023.
Figure 3. Schematic diagram of the variation in catalysts in the DRM hysteresis reaction under different conditions [27]. Reprinted with permission from Elsevier, copyright 2023.
Methane 03 00006 g003
Figure 4. Catalytic performances of Ni-Ir catalysts in DRM and the characterization of spent catalysts [39]. (A) CH4 conversion on Ni/MgAl2O4, Ni3Ir1/MgAl2O4, and Ir/MgAl2O4 (reaction conditions for the catalytic test: 650 °C, CH4/CO2 = 1/1, GHSV = 40,000 mL g−1 h−1, and 1 bar). SEM images of (B) Ni/MgAl2O4, (C) Ni3Ir1/MgAl2O4, (D) Ir/MgAl2O4, (E) XRD patterns of the spent catalysts. (F) The schematic diagram of the mechanism to activate CO2 with the assistance of carbonate [39]. Reprinted with permission from Elsevier, copyright 2022.
Figure 4. Catalytic performances of Ni-Ir catalysts in DRM and the characterization of spent catalysts [39]. (A) CH4 conversion on Ni/MgAl2O4, Ni3Ir1/MgAl2O4, and Ir/MgAl2O4 (reaction conditions for the catalytic test: 650 °C, CH4/CO2 = 1/1, GHSV = 40,000 mL g−1 h−1, and 1 bar). SEM images of (B) Ni/MgAl2O4, (C) Ni3Ir1/MgAl2O4, (D) Ir/MgAl2O4, (E) XRD patterns of the spent catalysts. (F) The schematic diagram of the mechanism to activate CO2 with the assistance of carbonate [39]. Reprinted with permission from Elsevier, copyright 2022.
Methane 03 00006 g004
Figure 5. (A) TEM images of Ni/CeO2-SiO2-C and Ni/CeO2-SiO2-P. (B) Stability test at 700 °C [52]. (C,D) Illustration of the reaction pathways of DRM. Reprinted with permission from Elsevier, copyright 2019.
Figure 5. (A) TEM images of Ni/CeO2-SiO2-C and Ni/CeO2-SiO2-P. (B) Stability test at 700 °C [52]. (C,D) Illustration of the reaction pathways of DRM. Reprinted with permission from Elsevier, copyright 2019.
Methane 03 00006 g005
Figure 6. (A,B) TEM images of the Ni/BNf and XAS images. (C) Activity and (D) stability of catalyst [19]. The figures are published on CCS chemistry and are available online [19].
Figure 6. (A,B) TEM images of the Ni/BNf and XAS images. (C) Activity and (D) stability of catalyst [19]. The figures are published on CCS chemistry and are available online [19].
Methane 03 00006 g006
Figure 7. (A) Illustration of the formation and role SMSI in anti-coking formation during DRM. (B) CH4 and CO2 conversion at 600 °C with dry methane (CH4/CO2/N2 = 1:1:2) and GHSV of 10,000 h−1 of the fabricated catalyst after CH4/CO2 reduction. Insert: Scheme of the concentration gradient of the Ni3+ and Ce3+ fraction at the bulk and surface of the nanoparticles. Reprinted with permission from [76]. Copyright {2022} American Chemical Society.
Figure 7. (A) Illustration of the formation and role SMSI in anti-coking formation during DRM. (B) CH4 and CO2 conversion at 600 °C with dry methane (CH4/CO2/N2 = 1:1:2) and GHSV of 10,000 h−1 of the fabricated catalyst after CH4/CO2 reduction. Insert: Scheme of the concentration gradient of the Ni3+ and Ce3+ fraction at the bulk and surface of the nanoparticles. Reprinted with permission from [76]. Copyright {2022} American Chemical Society.
Methane 03 00006 g007
Figure 8. (A) High–resolution STEM image; (B) long–term stability tests of Ni@NiSiAlMgOx at 673 and 873 K; (C) schematic representation of the MEO thin layer–confined and –stabilized Ni catalyst and the reference sample. Reprinted with permission from [85]. Copyright {2021} American Chemical Society.
Figure 8. (A) High–resolution STEM image; (B) long–term stability tests of Ni@NiSiAlMgOx at 673 and 873 K; (C) schematic representation of the MEO thin layer–confined and –stabilized Ni catalyst and the reference sample. Reprinted with permission from [85]. Copyright {2021} American Chemical Society.
Methane 03 00006 g008
Table 1. Standard reaction enthalpy of reactions involved in DRM.
Table 1. Standard reaction enthalpy of reactions involved in DRM.
EntryElementary StepΔHθ (kJ/mol)
1CO2 + CH4 → 2H2 + 2CO+247
2CO2 + H2 → CO + H2O+41
3CH4 → C + 2H2+75
42CO → C + CO2−172
5CO + H2 → C + H2O−175
Table 2. Performances of the Ni-based catalysts toward DRM.
Table 2. Performances of the Ni-based catalysts toward DRM.
CatalystReaction ConditionH2/COCO2 Conversion RateCH4 Conversion RateStability (h)Ref.
Ni@NiSiAlMgOxT = 873 K,
pCH4:CO2 = 1:1,
GHSV = 30,000 mL gcat−1 h−1
0.7840%31%~320[85]
Ni-cerium oxide nanocatalystsT = 873.15 K,
pCH4:CO2:N2 = 1:1:2,
GHSV = 10,000 h−1
N.A.60%40%550[76]
Ni/BNfT = 1023.15 K, pCH4:CO2 = 1:1, WHSV = 25,000 mL g−1 h−10.90N.A.60%100[19]
Ni/CeO2-SiO2-CT = 973.15 K, pCH4:CO2:Ar = 1:1:2, GHSV = 48,000 mL g−1 h−10.8580.5%67.8%50[52]
Ni/CeO2-SiO2-PT = 973.15 K, pCH4:CO2:Ar = 1:1:2, GHSV = 48,000 mL g−1 h−10.8987.3%78.5%50[52]
Ni/MgAl2O4T = 923.15 K, pCH4:CO2 = 1:1, GHSV = 40,000 mL g−1 h−10.9671.5%59.4%100[39]
Ni3Ir1/MgAl2O4T = 923.15 K, pCH4:CO2 = 1:1, GHSV = 40,000 mL g−1 h−10.9773.0%61.0%100[39]
Ir/MgAl2O4T = 923.15 K, pCH4:CO2 = 1:1, GHSV = 40,000 mL g−1 h−10.9463.6%50.2%100[39]
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

Wei, M.; Shi, X. Research Progress on Stability Control on Ni-Based Catalysts for Methane Dry Reforming. Methane 2024, 3, 86-102. https://doi.org/10.3390/methane3010006

AMA Style

Wei M, Shi X. Research Progress on Stability Control on Ni-Based Catalysts for Methane Dry Reforming. Methane. 2024; 3(1):86-102. https://doi.org/10.3390/methane3010006

Chicago/Turabian Style

Wei, Minghui, and Xuerong Shi. 2024. "Research Progress on Stability Control on Ni-Based Catalysts for Methane Dry Reforming" Methane 3, no. 1: 86-102. https://doi.org/10.3390/methane3010006

Article Metrics

Back to TopTop