Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Effects of alloying for steam or dry reforming of methane: a review of recent studies

Maki Torimoto and Yasushi Sekine *
Applied Chemistry, Waseda University, 3-4-1, Okubo, Shinjuku, Tokyo 169-8555, Japan. E-mail: ysekine@waseda.jp

Received 12th January 2022 , Accepted 18th February 2022

First published on 15th March 2022


Abstract

Methane is an important chemical resource, not only in natural gas but also in biogas, which can be regarded as a renewable energy resource. Reforming of methane with steam or carbon dioxide, which is important for producing hydrogen and syngas, is conducted at high temperatures using heterogeneous catalysts. To achieve high activity, stability, and low carbon deposition, many studies have been conducted in recent years on the use of alloys as active sites in these catalysts. This review presents a summary of recent studies of alloy catalysts, for which various secondary metals have been added to active metals, such as Ni. Then we summarize the current status of these alloys in terms of their structure, electronic state, and adsorption. The reported effects of alloying include improvement of dispersion and reducibility of the supported metal, change in catalytic performance such as activity and selectivity, and improvement of durability against carbon deposition, sulphur poisoning, and sintering. The directions of future research and development are summarized in terms of sulphur resistance, sintering inhibition, and high activity at low temperatures.


1. Introduction

1.1 Roles of alloys for catalysis

Supported metal catalysts have been used widely in various chemical processes up to the present day. The supported metal catalyst activity varies depending on the metal particle size, shape, structure, composition, electronic state, and so on.1 Therefore, increasing the catalyst activity and selectivity by controlling the structure, composition, and electronic states of supported metals is important for developing high-performance catalysts. In this sense, alloy catalysts, in which the original metal is intermixed with second, third, or more metals, have been studied widely as one method of modifying the properties of supported metals.2–4 Bi-metallic catalysts have various mixing states: a random solid-solution alloy, an intermetallic compound (IMC), a surface alloy, alloys with a core–shell structure, and others.5 Solid-solution alloys are classifiable into two types: substitutional and interstitial alloys. The former is an alloy in which substitution of metals occurs when the second metal has similar characteristics, such as the size or electronic properties, as the original metal. The latter is an alloy for which the second metal is sufficiently small to incorporate with the lattice void of the original metals. Unlike these random alloys, IMC has an ordered structure comprising two metal elements, yielding a crystal structure that differs from that of the mother metal.2,5 As for the bi-metal system of bulk, the change in the Gibbs free energy of mixing is often used to explain the mixing state of the two metals. The change in Gibbs free energy of mixing, ΔGmix, is calculable using the change in enthalpy, ΔHmix, and in entropy, ΔSmix during alloy formation (eqn (1)).
 
ΔGmix = ΔHmixTΔSmix(1)
If ΔGmix is negative, then a random alloy or intermetallic compound is formed spontaneously in the mixing condition. The sign and the magnitude of the ΔHmix value are important because ΔSmix is assumed to be positive.2 If the bond formation between the original metal and the second metal is exothermic (ΔHmix < 0), then ΔGmix is always negative and alloy formation occurs spontaneously. However, if the bond formation between the original metal and the second metal is endothermic (ΔHmix > 0), formation of the alloy depends on temperature. At high temperatures, the contribution of TΔSmix is sufficiently large to make ΔGmix negative. However, at a low temperature, ΔGmix is expected to be positive, which causes separation of the formed alloy. Furthermore, the transition from a random solid-solution alloy to an intermetallic compound can be estimated using eqn (1).2 Considering the transition from disorder to an ordered structure, ΔSmix is expected to be negative. Therefore, the transition is entropically unfavourable. By contrast, ΔHmix would be a large, negative value because of the stronger intermetallic bond energy. Therefore, whereas the low-temperature condition is favourable for the formation of intermetallic compounds, the high-temperature condition produces a disordered structure.

In recent years, alloys of some new types, such as single-atom alloys (SAAs)6,7 and high entropy alloys (HEAs),8,9 have received much attention. The former, SAAs, are alloys in which a very small amount of the second metal, such as 1[thin space (1/6-em)]:[thin space (1/6-em)]10 or 1[thin space (1/6-em)]:[thin space (1/6-em)]100, is added to the original metal. The second metal atoms are atomically dispersed on the original metal surface. In many SAA systems, a small amount of 8, 9, or 10 group metal (Pd, Pt, Rh, Ni, Ru, etc.) is added to the 11 group metals (Cu, Ag, Au).7 Earlier studies of SAAs have reported numerous benefits, including higher activity or selectivity, increased stability for coke formation, and increased stability for CO poisoning. In research in this field, STM investigations using a single crystal substrate have been conducted widely, in addition to XAFS and HAADF-STEM measurements.7 The latter type mentioned above, HEAs, are generally categorized as alloys comprising at least five nearly equimolar metal components (5–35 atom % for each component), although the definition remains controversial.8,9 Because configurational entropy increases as the number of components increases, HEAs have high stability (the high entropy effect). Reportedly, HEAs show interesting and unexpected catalytic performance for various reactions.

Effects of alloys on modifying the metal properties are often explained with two terms: ligand effect and ensemble effect.3,10,11 The ligand effect represents changes of the electronic structure caused by electron transfer between the original metal and the second metal. The ensemble effect represents a geometric effect by which the ensembles of active metals are diluted by the addition of second metals. The ensemble effect can suppress undesired side reactions such as coke formation. Nevertheless, it sometimes engenders decreasing activity of the main reaction when ensembles of active metals are necessary for activating reactants.3 Therefore, the metal components should be chosen carefully depending on the purpose.

As described above, alloys have numerous benefits, including increased catalytic activity, selectivity, and stability for various catalytic reactions. However, the role of the second metal is expected to be different for each reaction. It cannot be understood simply by a general explanation. For this review, we specifically examine the steam and dry reforming reaction of methane to gain a deep and comprehensive understanding of alloy effects.

1.2 Increasing demand for chemical usage of methane

The use of CH4 as a new alternative energy source and as a chemical material source instead of crude oil has become popular.12–15 Actually, CH4 is the main component of natural gas, shale gas, and bio-gas by fermentation.16 The chemical utilization of CH4 includes steam reforming of CH4 (SRM), dry reforming of CH4 (DRM), partial oxidation of CH4 (POM), autothermal reforming (ATR), and oxidative coupling of CH4 (OCM). Among these, SRM is the most commonly used process to produce H2; DRM is important for the effective utilization of greenhouse gases by converting CH4 and CO2 into industrially important syngas, a mixture of CO and H2. In fact, according to the Sixth Assessment Report (AR6) of the Intergovernmental Panel on Climate Change (IPCC) published in 2021, the gases exerting the strongest greenhouse effect are CO2, CH4, and N2O, in that order. Therefore, improving the methane reforming process efficiency to reduce the amount of CO2 emitted and to use greenhouse gases for chemical production is expected to become increasingly important for building a sustainable society. According to the International Energy Agency's (IEA) Energy Technology Perspectives 2020, 75 Mt of hydrogen are produced each year from natural gas (76%) and coal (23%). They are used for industrial applications such as oil refining (33%) and ammonia synthesis (27%).

SRM is a process to produce hydrogen from CH4 (eqn (2)). Water gas shift (WGS) occurs sequentially, thereby increasing the H2 yield (eqn (3)).

 
CH4 + H2O → CO + 3H2 ΔH0298 = +206 kJ mol−1(2)
 
H2O + CO → CO2 + H2 ΔH0298 = −41 kJ mol−1(3)
Although SRM has a long history and although it is a mature process, many issues remain to be resolved.13–15,17–23 First, industrially, SRM is conducted under high-temperature and high-pressure conditions such as 1000–1300 K and 3–25 bar, because of the stability of CH4 and a large endothermic reaction. Such harsh conditions lead to consumption of huge amounts of energy, requiring high costs because of the need for heat exchangers and heat-resistant materials.22 Next, Ni catalysts, which are mainly used industrially, deactivate easily because of coke formation, sintering, and poisoning by impurities such as S species in the feedstock. Noble metal catalysts (Pd, Rh, Ir, Ru, Pt, etc.) exhibit not only high activity but also high durability.22 However, because of their high cost, noble metal catalysts have not yet been industrialized. Therefore, development of a catalyst that can operate efficiently with high durability at low temperatures is desired in SRM.

The DRM process produces syngas, a mixture of CO and H2, from CH4 and CO2, as described in eqn (4).24–28 Sequentially, a reverse water gas shift reaction (RWGS) occurs (eqn (5)).

 
CH4 + CO2 → 2CO + 2H2 ΔH0298 = + 247 kJ mol−1(4)
 
CO2 + H2 → CO + H2O ΔH0298 = + 41 kJ mol−1(5)
Syngas is a raw material for various chemical reactions such as FT synthesis, methanol synthesis, and DME synthesis.28 Although DRM is an attractive reaction, its commercialization has been hindered by difficult challenges, which are almost identical to those for SRM: the requirement for high temperatures (over 1000 K) and rapid catalyst deactivation. Furthermore, DRM has a higher ratio of C to H than SRM, which makes it more prone to carbon formation.

1.3 Contents of this review

As described earlier, SRM and DRM have similar directions and issues to be developed. In both cases, the needs for high temperature and the catalyst durability are important issues. Many methods to improve activity, selectivity, and stability, have been reported. However, this review specifically examines methods to control the properties of the supported metals for thermal catalytic SRM and DRM and provides a detailed and comprehensive summary. In recent years, comprehensive review papers on SRM have been reported by Zhang et al.23 in 2021 and by Chen et al.22 in 2020. However, no report of the relevant literature describes a study of alloys. Many review papers of recent years have described DRM.24–35 Among these papers, Aziz et al.27 and Yentekakis et al.25 respectively produced reviews for bimetallic catalysts in 2020 and 2021. Bian et al.26 presented a review for nickel-based alloys in 2017. The present review assesses effects of alloys in SRM and DRM, with depth and breadth of comparison and with verification of the similarities and differences between the two reactions.

2. Properties of metals that exhibit high activity for SRM and DRM

The order of catalytic activities on active metals for SRM has been reported as shown below.36,37
Rh > Ru > Ni > Ir = Pd = Pt > Co > Fe
In addition, Ferreira-Aparicio et al.38 calculated the TOF in DRM for various metals and showed that the order differs for Al2O3 support and SiO2 support. The order of TOF of the methane reaction rate on each support is presented below.
Ni > Ru > Rh, Ir (SiO2 support)

Rh > Ni > Ir > Pt, Ru > Co (Al2O3 support)
For SRM and DRM, it is common to use Ni catalysts, which are active and inexpensive, supported on metal oxides such as Al2O3, which have high heat resistance.21,39 Other popular supports include durable MgAlOx support, mesoporous materials with high specific surface area, CeO2 base support with high oxygen storage capacity, and perovskite oxide support.40 However, Ni catalysts are prone to carbon deposition, sintering, and sulphur poisoning. Catalyst development is therefore still ongoing. Noble metal catalysts are more active. They have the salient advantage of being less prone to carbon deposition than Ni catalysts, but they are expensive and difficult to use industrially.41Table 1 shows prices for various metals as of December 2021, showing that the prices of noble metals are much higher than that of Ni. The most popular research is therefore to use the Ni catalyst as a basis for improving catalytic performance by improving the performance of supports and metals.
Table 1 The recent average prices of precious and base metals as of 2021 by Tanaka Precious Metal
Metal Pt Au Ag Pd Rh Ir Ru
$ per g 33.7 58.7 0.8 83.4 608.3 168.3 23.5


The rate-limiting step in SRM is reportedly dissociative adsorption of CH4, i.e., cleavage of the C–H bond of CH4 that occurs on a metal surface.42–45 In DRM, the mainstream theory is that the CH4 activation occurs on the metal, but it is generally accepted that activation of CO2 might occur on the metal or the support depending on the catalyst and operating conditions.46 Therefore, the key steps in the DRM process mechanism are regarded as being the adsorption and dissociation of CH4 on the metal site, adsorption and dissociation of CO2 on the support or the metal site, or both. Details of the DRM reaction mechanism have been summarized in reports by Yentekakis et al.,25 Fan et al.,40 and Ranjekar and Yadav.46

Details of the structure dependence of these reactions were investigated by Vogt et al.47 in 2020. Fig. 1 presents the possible reaction pathways on metals in SRM and DRM. Activation of H2O and CH4 requires σ-bond cleavage, whereas the activation of CO2 requires π-bond cleavage (Fig. 1(a)). An important point here is that the sites with higher activity for C–H activation and C[double bond, length as m-dash]O activation might be different. Reportedly, C–H cleavage occurs preferentially at co-ordination-unsaturated sites, whereas π-bond activation and cleavage occur preferentially at defect sites. Therefore, these reactions might have different active sites. Fig. 1(c) shows the general relation between supported metal particle size and TOF, where class 1 is for structure insensitivity, class 2 is for π-bond structure sensitivity, and class 3 is for σ-bond sensitivity. In class 2, the TOF is classified further as increasing and then decreasing or remaining constant. When the rate-limiting step involves the activation of a σ-bond, such as a C–H bond, it is usually shown as class 3 in Fig. 1(c). In this class, the TOF generally increases as the size of the metal nanoparticles decreases because the reaction proceeds predominantly on the co-ordination-unsaturated side. However, when the C[double bond, length as m-dash]O cleavage is rate-limiting, a certain amount of site co-ordination is required, which is classified as class 2. Consequently, pathway 1 (Fig. 1(b)) might be much more predominant for catalysts where C–H activation is difficult, whereas pathway 2 might be predominant for catalysts where the recombination of C and O adatoms is very slow compared to the activation of C–H bonds.


image file: d2cy00066k-f1.tif
Fig. 1 (a) The σ-bond cleavage is necessary for activation of H2O or CH4, and the π-bond cleavage is necessary for CO2 activation. (b) Possible two pathway for SRM and DRM; pathway 1 is a formyl-intermediate route and pathway 2 is a direct carbide route. (c) General classification of TOF dependence of metal particle diameter. Class 1 is for structure insensitivity, class 2 is for π-bond structure sensitivity, and class 3 is for σ-bond structure sensitivity. Reprinted with permission from ref. 47. Copyright 2019 American Chemical Society. https://pubs.acs.org/doi/10.1021/acscatal.9b04193.

However, the Ni particle size dependence of TOF in SRM and DRM investigated using 1–6 nm Ni/SiO2 catalysts differs slightly from the general classification as shown in Fig. 1(c). Catalysts supporting very small Ni particles (<1.5 nm) were less active: maximum activity was achieved at 2–3 nm of metal particle. This might be true because the balance of active sites for σ-bond and π-bond is important. A combination of different sites is necessary. In addition, the isotopically labelled experiments showed that the CH4 activation is not the only rate-determining step. Results suggest that the formation of CO by recombination of C and O and that the desorption of CO are likely to be important kinetically limiting factors for both SRM and DRM.

Furthermore, earlier reports show that CH4 activation requires an ensemble of active metals rather than a single active metal atom.17,48 Rostrup-Nielsen17 reported in 1984 that three Ni atoms are involved in the reaction. Rocha et al.48 reported in 2019 that two adjacent Ni atoms are necessary for CH4 activation. It is often discussed that when the active site is an ensemble of active metals, the ensemble effect, in which the second metal dilutes the ensemble of active metals, engenders a decrease in the activity.3 However, ensemble dilution is highly effective at inhibiting coke formation (details are presented in chapter 3). Numerous reports have described important improvements in stability achieved by adding small amounts of the second metal.

Catalytic activity is often expected to correlate with the electronic state of the metal surface or d-band centre.49–51 Such explanations using DFT calculations have been reported also for CH4.45 Abild-Pedersen et al.45 reported the correlation between the surface d-band centre and the transition state energy of CH3–H bond dissociation. Consequently, it is expected that if the d-band centre is increased by alloying, then the activity would be higher. If the d-band centre were decreased by alloying, then the activity would be lower. Recently, Joo et al.52 associated the high activity of CoNiFe ternary catalyst in DRM to the d-band centre of the metal. However, they reported the existence of an optimal d-band centre value. Roy et al.42 also compared surface properties such as surface energy, work function, and density of states of Ni–Pt alloy with the activation barrier for methane dissociative adsorption, concluding that, among the surface parameters, the d-band centre can be relied upon as a descriptor for predicting the trend of activation energy for CH4 dissociation in some alloy systems. Ray et al.53 also presented correlation between the d-band centre and the TOF in the DRM reaction in Ni–M alloys (M = Fe, Co, Cu). Conversely, the TOF became lower when the d-band centre was closer to the Fermi level. Therefore, the d-band centre of the metal surface is generally reliable as a descriptor to predict the activation barrier of methane dissociative adsorption, but one must consider the possibility of an optimum value of the d-band centre or the possibility that the d-band centre alone cannot explain the behaviour.

3. Coke formation mechanism

As described above, carbon deposition on Ni catalysts is a major issue in SRM and DRM. To overcome this difficulty, understanding the coke formation mechanism is important. In this chapter, the mechanism of carbon deposition is summarized.17,54–57 Also, methods to suppress carbon deposition by alloying and other methods are described.

3.1 Formation of adsorbed C species

Formation of the C species, which engenders coke formation, occurs as side reactions such as the methane decomposition reaction (eqn (6)) and the Boudouard reaction (eqn (7)).
 
CH4 → C + 2H2 ΔH0298 = +75 kJ mol−1(6)
 
2CO → C + CO2 ΔH0298 = −171 kJ mol−1(7)
Whether the formation of a carbon species occurs or not is determined by the thermodynamic balance between the main reaction and the carbon deposition side reaction.58,59Fig. 2 presents the limit of carbon formation from CH4 at 25.5 atm. The left side of the central curve shows the H/C and O/C ratios at which there is potential for carbon deposition. As shown in Fig. 2, the larger the ratio of H or O to C is, the more difficult it is for carbon to form. For this reason, SRM processes are operated industrially under conditions with a great amount of steam, such as steam to CH4 ratios of 2.5–3.0,21 requiring much energy to supply the excess steam. In addition, Yoon et al.60 recently reported the necessity of considering not only the CH4 decomposition steps (steps 1–4) but also the CO gasification steps (steps 5–7) when considering the C species formation (Fig. 3). In these steps, step 4 or step 5 are alternatives. If steps 5–7 occur after step 3, then carbon deposition could be suppressed by CO gas production.
image file: d2cy00066k-t1.tif

image file: d2cy00066k-t2.tif

image file: d2cy00066k-t3.tif

Step 4 : CH* → C* + H*

Step 5 : CH* + O* → CHO*

Step 6 : CHO* → CO* + H*

Step 7 : CO* → CO(g)
They investigated the pathway of SRM on the Ni(111) surface and Ni-Ru and Ni-Rh alloy surfaces using DFT calculations and MD (molecular dynamics). First, the adsorption energy of CH4 on Ni is very weak (approx. −0.02 eV), indicating that CH4 molecules are physisorbed. Next, the C–H bond of CH4 is broken and CH4 decomposition proceeds to form CH* species (steps 1–3). Furthermore, two paths exist: one where CH* decomposes into C* and H* (step 4), and another where CH* reacts with O* to form CHO* (steps 5–7). Because the activation barrier of the former path (CH* → C* + H*) is high, the latter path (CH* + O* → CHO*) occurs more preferentially irrespective of Ni surfaces or Ni alloy surfaces. The activation barrier of Ru-doped Ni surface for step 5 was the smallest, which suggests that Ni-Ru has the lowest activation energy and that it is the most effective at suppressing carbon deposition.

image file: d2cy00066k-f2.tif
Fig. 2 The diagram of carbon limit. The left side of the bold curve has the potential for carbon deposition. Reprinted with permission from ref. 58. Copyright 2006 Elsevier.

image file: d2cy00066k-f3.tif
Fig. 3 The schematic images of reaction paths for SRM on Ni. Reprinted with permission from ref. 60. Copyright 2017 Elsevier.

In addition, the stability of the C atom on the metal surface is often discussed to explain whether the catalysts are prone to coke formation. Besenbacher et al.61 investigated the stability of the C atom on Au/Ni(111) by DFT calculation, which revealed that the C atoms were less stable on a Au-loaded Ni surface. Lower stability of the adsorbed C atoms on the metal leads to a greater likelihood that they will react with the adsorbed O species to produce CO, thereby suppressing the nucleation of graphite. Consequently, the stability of C atoms on alloy surfaces is expected to be an indicator of the ease of carbon deposition. Relevant details are presented in section 4.1.

3.2 Growth of C species making coke

Carbon atoms (Cα) bind other C species on the Ni metal to form larger carbon species.54,62,63Fig. 4 and Table 2 summarize coke formation of various types. The adsorbed atomic carbon (Cα) forms C–C bonds on the surface and rearranges to form polymeric amorphous carbon (Cβ), which is less reactive than Cα. Cβ species accumulate on the Ni surface, thereby forming graphite platelet films (Cc), or dissolve into Ni bulk. For reactions of heavy hydrocarbon, encapsulating carbons that cover Ni are sometimes produced. Then Ni carbide, Ni3C, is formed in the bulk (Cγ). Vermicular carbon or whiskers (Cv) are formed by the solid solution of C species in Ni metal. Whisker carbon is formed when C species formed on the Ni surface dissolve into the Ni metal, diffusing, and pushing up Ni particles from the opposite side (Fig. 4). Formation of numerous whisker carbons is expected to engender the destruction of the catalyst itself. The encapsulating carbons covering the Ni surface lead to decreased activity.
image file: d2cy00066k-f4.tif
Fig. 4 Formation, gasification, and transformation of carbon on nickel form carbon monoxide (a, g, s refer to adsorbed, gaseous, and solid states, respectively). Reprinted with permission from ref. 63, redrawn. Copyright 1982 Marcel Dekker Inc. Taylor & Francis ltd, https://http://www.tandfonline.com.
Table 2 The various structures of carbon formation. Reprinted with permission from ref. 63. Copyright 1982 Marcel Dekker Inc. Taylor & Francis ltd, https://www.tandfonline.com
Structural type Designation Temperature/K
Adsorbed, atomic carbon (surface carbide) Cα 473–673
Polymers, amorphous films, or filaments Cβ 523–773
Graphite (crystalline) platelet, films Cc 773–823
Vermicular (polymeric, amorphous) filaments or whiskers Cv 573–1273
Ni carbide (bulk), Ni3C Cγ 423–523


Factors such as the Ni ensemble size and particle size also play important roles in the growth of C species. Rostrup-Nielsen17 reported the numbers of Ni metal ensembles required for each process (eqn (8)). In eqn (8), subscript letters n, m, m′, respectively represent the numbers of Ni atoms required for CH4 dissociation, C species formation, and dissolution of C species into Ni metal. Although Ni ensembles are also required for CH4 dissociation, the necessary numbers for CH4 activation and formation of C species are smaller than that required for C species to dissolve into Ni (m′ > n, m). In addition, Rostrup-Nielsen17 described in that report that three Ni atoms are involved in the CH4 reforming reaction.

 
image file: d2cy00066k-t4.tif(8)
Furthermore, some papers have reported the Ni particle size effect on carbon. Kim et al.64 reported that carbon deposition is likely to occur when the Ni particles are larger than 7 nm. Vogt et al.47 reported that the maximum amount of surface carbon deposition was found at about 4 nm for SRM, and reported the amount of carbon increased with nanoparticle size for DRM. The low-coordinated Ni sites can be sites of carbon nucleation.65,66 Therefore, considering the coke formation mechanism described above, coke formation can be suppressed by loading small metal particles, by alloying the supported metal, by diluting the ensemble of active metals necessary for the dissolution-precipitation nucleation of carbon whiskers, or by blocking the low-coordinated sites which can be carbon nucleation sites.

4. Effects of alloying on steam reforming of methane (SRM)

This chapter summarizes effects of alloys on activity, selectivity, and durability in SRM. Industrially, Ni catalysts supported on Al2O3 are used widely.21,39 Noble metal catalysts (Rh, Ru, Pd, Pt, etc.) are highly active and stable, but they are too expensive for industrial use. Ni is cheap and common, but it presents difficulties such as catalyst deactivation because of carbon deposition, sintering, and poisoning, and lower activity than that of noble metals. Therefore, the development of catalysts with high activity and high durability is required. Many studies have examined catalysts using Ni alloyed not only with base metals, but also with noble metals.

Table 3 briefly presents the catalytic performance of the Ni-based catalysts supported on oxide supports introduced in this chapter. The respective sections provide relevant details. This table, which roughly categorizes the properties of metal, catalytic performance, and stability, presents the superiority or inferiority of each item compared to a single Ni catalyst. Blank spaces indicate that the corresponding survey has not been conducted. It is noteworthy that the alloy effects often vary depending on the catalyst, composition, oxide support, reaction conditions, etc. This table shows only representative features.

Table 3 The effect of Ni-based alloy for SRM
Properties of metal Catalytic performance Resistance Remarks
Metal reductivity Metal dispersion Activity Selectivity Long-term stability Coke Sulphur Sintering
Fe–Ni High High High High High Fe could enhance the resistance for sulphur
Co–Ni Low Low High Co block the low-coordinated Ni sites
Cu–Ni High High High WGS sel. High Cu could promote WGS
Sn–Ni Low High High Sn increases the apparent activation energy
Ru–Ni High High High High Not need the H2-treatment before activity tests
Rh–Ni High High High Low WGS sel. High High Rh–Ni has self-activation and self-regenerative activity
Ag–Ni Low High High Ag blocks the more active sites such as steps and edges
The apparent activation energy increases by Ag addition
Re–Ni High High Low WGS sel. High Re addition does not enhance WGS
Re–Ni has higher TOF than Ni
Ir–Ni High High High High High Ir addition could enhance the sintering resistance
Ir increases the barrier for CH4 dissociation
Pt–Ni High High High High High A small amount of Pt increases the metal dispersion, but excessive amounts of Pt lead to agglomeration
Pt increases the activation barrier for CH4 dissociation
Au–Ni High High/low High High High A barrier for CH4 dissociation increases by Au addition. C atom is less stable on Au/Ni surface


Some alloy catalysts reportedly have not only resistance to carbon deposition but also to sulphur and sintering, thereby providing improved long-term stability compared to Ni catalysts. In many alloys, negative effects such as an increase in the CH4 dissociation barrier have also been reported. In some cases, total activity was improved because of the effects of improved metal reducibility or dispersibility by addition of the second metal. In still other cases, WGS selectivity (CO selectivity) was reported as changed by the alloying.

4.1 Addition of base metal to Ni catalysts

The addition of base metals to Ni presents many benefits, including improved coke resistance, increased sulphur resistance, and improved reducibility and dispersibility of Ni, although the catalyst component is still inexpensive. However, some secondary metals reduce the initial activity slightly. Therefore, it will be necessary to control selection of the secondary metal and the composition ratio carefully.
Fe–Ni. Various structures such as core–shell, supported, and reductive precipitation of perovskite oxides have been reported for Ni catalysts doped with Fe. In addition, high activity, high resistance to carbon deposition, resistance to sulphur poisoning, and improved Ni reducibility have been reported. Tsodikov et al.67 reported that the core–shell catalysts with a Ni–Fe alloy nanoparticle core and a γ-Fe2O3 shell showed high activity for SRM and stability to coke formation and resistance to H2S (Fig. 5). The catalyst was produced by H2 reduction of Ni-loaded spinel Mg(FeAl)2O4 support. Its structure was characterized using XRD, XAFS, HRTEM, and Fe Mössbauer data. The reaction tests confirmed the core–shell structure: the core was a metallic NiFe alloy cluster; the shell was superparamagnetic γ-Fe2O3. They described that NiFe alloy provides high activity comparable to the commercial catalyst, and that the γ-Fe2O3 shell was the active site for the decomposition of H2S to S. Konstantinov et al.68 reported that Ni-Fe alloy catalysts prepared by epitaxial coating on the commercial spherical γ-Al2O3 showed high activity for SRM and showed resistance to H2S. Activity tests over Ni–Fe alloy catalysts and commercial SRM catalyst (NIAP 03-01) were conducted for 30 h with 30 ppm H2S. Although the activity of commercial NIAP 03–01 decreased after 10 h because of the H2S poisoning, that of Ni–Fe alloy catalyst remained high for 30 h. They assumed that the conversion of H2S to S occurred during the reaction. Subsequent XRD, TEM, and Mössbauer spectroscopy of Fe revealed that the catalyst after the activity tests contained MgFe0.1Al1.9O4 spinel phase, metallic Ni, γ-FeNi, and FeAl2O4 iron–aluminium spinel, but not sulphur-containing compounds. They inferred that both high SRM activity and decomposition of H2S to S occurred because Ni–Fe alloy particles bonded tightly to the support.
image file: d2cy00066k-f5.tif
Fig. 5 TEM images of Ni-loaded spinel Mg(FeAl)2O4 after reaction tests. The core–shell structure was observed; core is Ni–Fe alloy, shell is γ-Al2O3. Reprinted with permission from ref. 67. Copyright 2015 Elsevier.

Djaidja et al.69 also reported the increased coke resistance of Ni–M/MgO or (Ni–M–Mg)2Al (M = Fe, Cu). Their results showed that Ni–Fe or Ni–Cu alloying improves Ni reducibility, high activity, and high carbon tolerance. Furthermore, Provendier et al.70 and Thalinger et al.71 reported effects of Fe on Ni–perovskite supports. Provendier et al.70 investigated the series of LaNixFe1−xO3 (0 ≤ x ≤ 1) and found that although the no NiFe alloy was formed, the addition of Fe to perovskite supports can decrease the Ni particle size, thereby limiting coke formation. Thalinger et al.71 investigated two Ni–perovskite catalysts; Ni–La0.6Sr0.4FeO3−δ (lanthanum strontium ferrite, LSF) and Ni–SrTi0.7Fe0.3O3-δ (strontium titanium ferrite, STF). Ni–Fe alloy particles were observed after H2 treatment. These manifestations depend on the reducibility of supports. They described that, because LSF has more reducibility than STF, the degree of Ni–Fe alloying is high, resulting in suppressed catalytic performance. Consequently, optimization of the complex oxide support and the reduction conditions can lead to control of the supported Ni–Fe catalysts.

Co–Ni. You et al.65 reported the role of Co addition to Ni/Al2O3 catalysts. After Ni–Co/Al2O3 catalysts with 1–15%Co and 12%Ni were prepared using the co-impregnation method, their activity and stability toward coke were investigated. In fact, Co addition leads to decreased metal dispersion, and thereby activity at low temperature. However, 180 h stability tests conducted at 1073 K revealed that 7%Co–12%Ni/Al2O3 catalysts retained high activity of 95% for 180 h, although 12%Ni showed deactivation. Their results show that Co can block the active low coordinated Ni sites and that it can increase coke resistance.
Cu–Ni. Reportedly, Ni–Cu alloying provides benefits such as improved Ni reducibility, high activity, high carbon tolerance, and high WGS activity.

As described in the Fe–Ni section, Djaidja et al.69 investigated the increased coke resistance of Ni–M/MgO or (Ni–M–Mg)2Al (M = Fe, Cu). Results showed that Ni–Fe or Ni–Cu alloying offers improved Ni reducibility, high activity, and high carbon tolerance. Huang et al.72 reported that the effect of Cu addition to Ni supported on Sm-doped ceria, Gd-doped ceria, and α-Al2O3. Activity tests over 2wt%Ni catalysts showed that Ni catalyst supported by Sm-doped ceria had much higher activity at a lower temperature, the highest H2 production rate, and the lowest carbon formation rate. This markedly increased activity was driven by the high surface area and oxygen-ion conductivity of the doped ceria support. Next, a series of Ni–Cu catalysts supported on Sm-doped ceria support was prepared by the co-impregnation method and was examined its catalytic properties. Although Cu is inactive for SRM, 0.1wt%Cu–0.5wtNi catalyst showed a much higher CH4 conversion rate and higher H2 production rate than the 0.5wt%Ni catalyst. Because Cu catalyst is well known to show high WGS activity, the addition of Cu can increase the WGS activity and therefore SRM activity by increasing the CO removal rate.

Sn–Ni. Nikolla et al.66,73 reported that the addition of Sn to Ni catalyst can improve long-term stability, although the initial activity is slightly lower than that of Ni catalyst. 1wt%Sn or 5 wt%Sn was loaded on Ni/YSZ by the incipient wetness method. Fig. 6 presents results of long-term stability tests over these catalysts. As shown in Fig. 6, the addition of 1wt%Sn can drastically increase the stability, although 5 wt% of Sn led to decreased activity. In addition, the apparent activation barriers they calculated were 101 ± 4 kJ mol−1 for Ni, and 132 ± 4 for 1wt%Sn/Ni supported on YSZ.66 Moreover, DFT calculation revealed that Sn atoms preferentially displace low-coordinate Ni sites. Consequently, the main active sites were moved from under-coordinated Ni sites, where carbon nucleation occurs, to more abundant well-coordinated Ni sites. Amounts to Ni catalysts. Their effects on carbon deposition are mainly reported.
image file: d2cy00066k-f6.tif
Fig. 6 The long-term activity of Ni/YSZ, 1wt%Sn/Ni/YSZ, 5 wt%Sn/Ni/YSZ measured at the steam/carbon ratio of 0.5 and at 1073 K. the 1 wt% addition of Sn could improve the stability with a little of decrease in initial activity. Reprinted with permission from ref. 73. Copyright 2007 Elsevier.

4.2 Addition of noble metals to Ni catalysts

The addition of noble metals to Ni catalysts in small quantities has attracted much attention in recent years. In many cases, small amount of noble metals (Ru, Rh, Ag, Pt, Au, etc.) are added to Ni catalysts. Their effects on carbon deposition are mainly reported.
Ru–Ni. Nawfal et al.74 investigated SRM activity using catalysts impregnated with 0.5wt%Ru supported on NixMg6−xAl2 oxide prepared by a hydrotalcite route. Among these catalysts, 0.5wt%Ru/Ni6Al2Ox showed the highest activity. They required no reduction treatment before reactions. In fact, 0.5wt%Ru/Ni6Al2Ox remained highly active after 10 cycles of tests. The surface carbon did not decrease the activity. The Ru–Ni interaction was sufficient to show that Mg was not necessary for the oxide. In addition, as described in chapter 3, Yoon et al.60 calculated the activation energies of each step involving the CO gas evolution and predicted that the addition of Ru to alloy is effective for suppressing carbon deposition. Baek et al.75 reported that Ru/Ni/MgAl2O4 has self-activation property and that H2 treatment was not necessary.
Rh–Ni. The introduction of a small amount of Rh on Ni catalyst has been studied. Some benefits have been reported such as improved activity, high H2 yield, resistance to sintering, self-activation, and self-regenerative activity.

Katheria et al.76 investigated the effects of small amounts of Rh doped to Ni catalysts. First, xwt%Rh–15 wt%/MgAl2O4 (x = 0.1–1.0) were prepared using the sequential impregnation method. Activity tests over these catalysts were conducted under ambient and high-pressure conditions (Fig. 7). Activity tests at ambient pressure showed that a small amount of Rh (0.1wt%Rh) was able to increase the SRM activity and H2 yield drastically, but a further increase of Rh amount did not engender a further increase of activity or H2 yield. However, activity tests conducted at high pressures over these catalysts differed. The degrees of stability over 0.5wt%Rh and 1.0 wt%Rh loaded catalysts were similar but those over 0.1wt%Rh or 0wt%Rh loaded catalysts were lower. In addition, the CO selectivity of Rh–Ni catalysts was higher than that of Ni catalysts at both pressures. The presence of Rh on the surface and the smaller crystallite size of active metals by the addition of noble metal might be attributed to the higher CO selectivity.


image file: d2cy00066k-f7.tif
Fig. 7 The effect of Rh loading amount on methane conversion at pressure (i) 1 bar and (ii) 10 bar at 873 K; GHSV, 2.0 × 106 h−1; W/F is 0.34 gcat h mol−1. Reprinted with permission from ref. 76. Copyright 2019 Elsevier.

Morales-Cano et al.77 reported that Ir or Rh addition to Ni catalysts can suppress sintering. After they prepared 2.89wt%Ni/Al2O3, 0.97wt%Ir–2.91wt%Ni/Al2O3, and 1.12wt%Rh–2.92wt%Ni/Al2O3 by incipient wetness impregnation, aging treatments were conducted at 1073 K, under 30 bar, and H2O/H2 atmospheres for 240 h to simulate long-term stability. After the aging process, Ni particles of Ni/Al2O3 were larger than 100 nm, whereas Ni–Ir/Al2O3 and Ni–Rh/Al2O3 particles were smaller than 50 nm.

In addition, Li et al.78 reported that the Rh-doped MgAlO oxide has self-activation and self-regenerative activity, although Ru-doped MgAlO does not. These activities were achieved by the oxidative incorporation of surface NiO into the lattice Ni2+ of Mg(Al, Ni)O and the reductive transfer of lattice Ni2+ to surface Ni0 by hydrogen spill over, leading to continuous regeneration of active Ni metal particles.

Ag–Ni. Effects of the addition of Ag to Ni catalyst have been reported by Dam et al.79 and Wang et al.80 According to DFT calculations,80 Ag selectively substitutes for the more active sites of Ni, such as edges and steps. In the Ni–Ag alloy system, Ag is inactive to SRM. Moreover, the addition of Ag reduced the activity of adjacent Ni sites significantly because of charge donation from Ag to Ni. Fig. 8 portrays Arrhenius plots of Ni, Ni0.95Ag0.05, and Ni0.75Ag0.25 supported on MgAl2O4. The apparent activation energies of these catalysts were calculated respectively as 103.6 kJ mol−1 for Ni, 119.5 kJ mol−1 for Ni0.95Ag0.05, and 127.3 kJ mol−1 for Ni0.75Ag0.25. Consequently, Ag addition to Ni catalyst engenders decreased SRM activity, decreased TOF, and increased apparent activation energy.80 However, Ag can increase the carbon resistance by blocking the active sites on the steps for the nucleation and growth of the filamentous carbon.79 Dam et al.79 reported that only 3wt%Ag replaced with Ni led to a carbon amount that was one order of magnitude lower, thereby suppressing the formation of filamentous carbon.
image file: d2cy00066k-f8.tif
Fig. 8 The Arrhenius plots for 12wt%Ni (◆, blue), 12wt%Ni0.95Ag0.05 (▲, green) 12wt%Ni0.75Ag0.25 (■, orange). Reaction conditions: S/C = 3.5, CH4/H2 = 10, W/F0 between 0.018 and 0.037 g s mol−1. Reprinted with permission from ref. 80. Copyright 2017 RSC publishing.
Re–Ni. Xu et al.81 and Wang82 reported the superiority of Ni–Re alloy catalysts. Xu et al.81 reported the enhancement of catalytic activity and stability by adding Re to Ni monolithic catalysts. They loaded 5 wt%Ni and 0.5, 5, and 10 wt%Re directly on Ni monoliths without oxide support and designated them respectively as Ni5, Ni5Re0.5, Ni5Re5, and Ni5Re10. Fig. 9 shows the turnover frequency (TOF), defined as the number of H2 molecules formed per second per Ni active site, were calculated and shown. These results demonstrated that the increment of Re amount can enhance both the catalytic activity and stability. Furthermore, CO selectivity increased for Ni–Re alloy catalysts compared to Ni, suggesting that the addition of Re promoted the SRM reaction, but not the WGS reaction. XRD, in situ X-ray absorption fine structure spectroscopy (XAFS) during H2 reduction, and H2-TPR revealed that Re can drastically promote the reduction of NiO to Ni. Moreover, DFT calculations revealed that the adsorption energy of a hydrogen atom on Re(001) was smaller than that on Ni(111), Ni(011), and Ni(001).
image file: d2cy00066k-f9.tif
Fig. 9 TOFs as a function of TOS at (a) 873 K (b) 973 K. CH4/H2O/N2 = 10/13.6/30 mL min−1, GHSV = 6400 h−1. Reprinted with permission from ref. 81. Copyright 2020 RSC publishing.

Therefore, the H atom adsorbed more readily onto Re than on Ni during the reaction, thereby preventing oxidation of the surrounding Ni and enhancing the stability. Wang also reported Ni–Re alloy nanowire type catalysts, which showed low-temperature activity at 623 K.82

Ir–Ni. As described in the Rh–Ni section, Morales-Cano et al.77 reported that Ir or Rh addition to Ni catalysts can suppress sintering. Furthermore, Liu et al.83 reported increased coke resistance and dispersibility of metal particles over Ir doping to Ni/MgAl2O4 on bi-reforming of CH4 with steam and CO2. The formation of Ir–Ni alloy was confirmed after H2 treatment by HAADF-STM. The metal particles of Ir10Ni90/MgAl2O4 were 2–3 nm, although those of Ni/MgAl2O4 were 5–40 nm. Because nucleation of carbon occurs on large ensembles, coke formation occurs only with difficulty on such small metal particles. Furthermore, XPS measurements revealed that more C–O species were formed on Ir10Ni90/MgAl2O4 than in Ni/MgAl2O4, indicating that the addition of Ir promoted the oxophilicity of the metal surface. The increased amount of surface oxygenates was inferred as the reason for the increase in coke gasification rate during the catalytic process. Although Ir doping increased the apparent activation energy, it increased the coke resistance because of the small ensemble sizes, increased oxophilicity, and increased the barrier for CH4 dissociation.
Pt–Ni. Jaiswar et al.84 reported the effects of a small amount of Pt doping to Ni/MgAl2O4. They prepared the Pt doped 15wt%Ni catalysts supported on MgAl2O4 by the sequential impregnation method, varying Pt loading amount from 0.01–1.0 wt%. The prepared catalysts, xw%Pt–15w%Ni/MgAl2O4 were denoted as xPt15Ni. Three effects of Pt doping can occur: increased reduction of the supported metal, increased dispersion, and increased SRM activity. First, it was confirmed using H2-TPR measurements that, as Pt amount increased, the degree of reduction increased. The degree of reduction of 0Pt15Ni was 44.9%, but that of 1.0Pt15Ni was 67.7%. In addition, the average particle sizes of supported metals estimated from TEM images decreased by the addition of Pt. The average sizes of 0Pt15Ni, 0.1Pt15Ni, and 1.0Pt15Ni were, respectively, 14.9 nm, 7.6 nm, and 9.3 nm. Those sizes indicate that a small amount of Pt can increase the dispersibility of metal, but an excessive amount of Pt addition engenders agglomeration. Furthermore, activity tests over these prepared catalysts at 873 K at 1 or 10 bar pressures revealed that Pt doping amount of up to 0.1 wt% can increase the SRM activity, but a further increase of Pt led to the decrease of SRM activity. The activity was stable for 8 h over all catalysts during tests at 1 bar, but they decreased gradually during tests at 10 bar. Among the catalysts, 0.1Pt15Ni showed the highest activity and the latest rate of deactivation. The authors concluded that the high SRM activity of 0.1Pt15Ni corresponds to the high surface area of metal, and that stability corresponds to the small metal size because earlier reports show that the low particle size of the metal decreases the carbon formation rate.85

Related to this research, the effect of Pt doping to Ni catalysts or Pt–Ni alloy for CH4 dissociative chemisorption, the rate-determining step of SRM, has been studied using DFT calculation by Roy et al.42 As described herein, activation energies for CH4 dissociative chemisorption on Pt-doped Ni-based alloy and Ni-doped Pt-based alloy surface were calculated from transition-state calculations. In the case of Ni-based alloy, the activation barrier increased linearly with the amount of Pt. However, no linear increase or decrease trend was observed for Ni-doped Pt-based alloy. DFT calculation also demonstrates the relation between the activation energy and the d-band centre; the activation energy decreased as the values of the d-band centre increased. However, they also described that a single surface-based descriptor cannot provide a complete understanding of the reactivity changes.

Au–Ni. Details of Au–Ni alloys have been investigated by Besenbacher et al.61 using techniques such as STM and DFT calculation to study their effects on CH4 activation and carbon deposition. Some alloys do not mix in bulk but instead form stable alloys on the outermost surface: Au–Ni is of this type. In Fig. 10 (left), the black circles represent Au. The Ni around the Au has a different colour from that of normal Ni, signifying different electronic density of states. The DFT calculation shows that one neighbouring Au increases the barrier to CH4 dissociation on Ni atoms by 16 kJ mol−1, whereas two neighbouring Au atoms increase it by more than a factor of two. The dissociation on Au atoms is expected to have a higher barrier. The Au–Ni alloy effects on carbon deposition are also described in terms of the stability and coverage of C atoms adsorbed onto the Ni surface. Lower stability of the adsorbed C presents a stronger tendency to react with adsorbed O to form CO, and to give lower coverage of the C atoms. The energies of C atoms on the Ni(111) and Au/Ni(111) surfaces were calculated using DFT calculation (Fig. 10(right)). The stability of the C species at the Ni site next to Au is reduced considerably.
image file: d2cy00066k-f10.tif
Fig. 10 (left) the STM images of Ni(111) with (A) 2% and (B) 7% of a Au monolayer. The Ni atoms adjacent to Au looks brighter than normal Ni, indicating the different electronic states. (Right) The calculated chemisorption energy of C atom on the different positions A–D of Ni(111) or Au/Ni(111) by DFT calculation. The presence of Au significantly increases the chemisorption energy of C atom. Reprinted with permission from ref. 61. Copyright 1998 AAAS.

Rocha et al.48 also reported effects of Au addition to Ni/Al2O3 catalysts. The Au–Ni catalysts were prepared by mixing the Au solution with the reduced Ni/Al2O3. Diffuse reflectance infrared Fourier transform spectroscopy of adsorbed CO (DRIFTS-CO) revealed that the Au atoms blocked the low-coordinated Ni sites, such as steps and corners, and modified the Ni surface structure. In addition, by adding Au, the CO adsorption form changed from a bridge to a linear form. The electronic density was increased by the electron transfer from Au to Ni. Activity tests revealed that the addition of Au to Ni catalysts engenders a decrease the SRM activity. The authors conclude that blocking of the low coordinated Ni sites by Au, the decreased stability of intermediaries engender hindrance of the growth of graphitic carbon.

Palma et al.86 investigated effects of Au addition to NiLaO3 perovskite support and described that Au substitutes the step sites of Ni and suppresses carbon deposition. Although Au–Ni has a negative effect on CH4 activation because the Ni step sites are also active sites for C–H activation and because Au decreases the activity of surrounding Ni atoms, the total activity is enhanced because of increased Ni dispersion by Au addition. Chin et al.87 also reported decreased initial activity and deactivation rate. Sapountzi et al.88 reported that Au-modified Ni catalyst showed sulphur resistance and increased reducibility of NiO. In addition, Wang et al.89 used STM and DFT calculations to investigate single-atom alloys of Ni on the Au(111) surface. They described that Ni–Au can improve the surface activity of Au(111) considerably because of the high CO adsorption of Ni atom on Au(111). Therefore, Au–Ni catalysts might be catalysts with high stability, although they reduce the initial activity slightly.

4.3 Others

Co-based alloy catalyst. Shen et al.90 investigated single and bimetallic Co-based catalysts (Co, Co–Ni, Co–Cu, Co–Al) supported on CeO2. Among Co/CeO2 catalysts with different Co loading amounts, 12%Co/CeO2 showed the highest CH4 conversion. The Co–Ni and Co–Al catalysts showed superior activity to that of Co single catalysts. However, the H2 yield of Co–Al was much lower. The highest level of performance was exhibited by Co80Ni20 (total weight 16 wt%) catalysts.
Pd–Zn catalysts in an electric field. The author's group investigated SRM in an electric field.91–95 In this system, constant current is applied after two electrodes are inserted on and under the catalyst bed (Fig. 11(Left)). Manabe et al.91 and Okada et al.92 revealed that the reaction mechanism in an electric field differs from that of a conventional thermal catalytic reaction. In an electric field, the proton conduction on oxide supports occurs; then CH4 dissociative adsorption occurs because of collision of the proton. Results demonstrated further that the active sites for SRM in an electric field differ. The active sites for thermal catalysis are metals on the surface, but those in an electric field are metals at the perimeter.
image file: d2cy00066k-f11.tif
Fig. 11 (Left) The schematic images of the reactor. Two electrodes are inserted on and under the catalyst bed and the electrical current was applied. (Right) The calculated TOFs of PdZn/CZO with Zn/Pd ratios for thermal catalysis, SR, (○) and catalysis with an electric field, ER (●). Reaction conditions: CH4/H2O/Ar = 1/2/7, total flow rate = 100 sccm; catalysts weight, 100 mg; furnace temperature, 423 K for catalysis with an electric field, 673 K for thermal catalysis; input power 1.5 W. reprinted with permission from ref. 96. Copyright 2020 RSC publishing.

The author of this review investigated effects of Pd–Zn alloy on catalytic activity for SRM in an electric field.96 A series of Pd–Zn alloys with different composition ratios (Zn/Pd = 1/9, 1/4, 1/1, molar ratios) and fixed Pd loading weight (5 wt%) loaded on Ce0.5Zr0.5O2 were prepared and tested. Calculation of TOF over these catalysts revealed the different effects of Zn addition on TOFs with or without an electric field (Fig. 11(right)). In Fig. 11(right), TOF normalized by the number of metal atoms on the surface (TOF-s) was used for the thermal SRM and TOF normalized by the number of metal atoms at the perimeter was used for the SRM in an electric field based on the different reaction mechanisms. In the case of the normal thermal catalytic reaction, the TOF decreased concomitantly with increasing Zn content; Pd1Zn1 was almost inactive. This finding might be attributable to dilution of the Pd–Pd ensemble by Zn addition. It is noteworthy that Pd1Zn1, which was almost inactive in the thermal catalytic reaction, showed the highest TOF value in the electric field. This might indicate that ensembles of metals are no longer necessary for the reaction in the electric field. Furthermore, from the CO adsorption IR, the ligand effect was observed that the electron density of Pd increased concomitantly with increasing Zn addition. The intermediate of SRM in the electric field is regarded as cationic. The cationic intermediate might be more stable on electron-enriched Pd. Consequently, the effects of alloys for catalysis in the electric field differ greatly from those of thermal catalysis.

5. Effects of alloying on dry reforming of methane (DRM)

Because DRM is an attractive process that can convert the greenhouse gases of methane and CO2 into useful chemicals (CO and H2), numerous studies of DRM have been conducted to date. Widely various alloy catalysts have been investigated. Yentekakis et al.,25 Bian et al.,26 and Aziz et al.27 have provided comprehensive summaries of alloys in DRM. Therefore, this chapter introduces recent studies of alloy DRMs published over the last five years.

Alloy catalyst structures of several types have been introduced in this chapter. Some are metal-supported on Al2O3 or other supports, others are core–shell catalysts with alloy nanoparticles encased in oxides, whereas others investigated widely in recent years are alloy metal particles formed by hydrogen reduction of oxide precursors that include Ni and secondary metals in their composition. Table 4 briefly summarizes the catalytic performance of Ni-based catalysts supported on oxide supports introduced in this chapter. The dedicated sections provide relevant details for them. This table roughly categorizes the properties of metal, catalytic performance, and stability, and presents the superiority or inferiority of each item compared to a single Ni catalyst. Blank spaces show that the corresponding survey has not been conducted. It is noteworthy that the alloy effects often vary depending on the catalyst, composition, oxide supports, reaction conditions, etc. This table presents only representative features.

Table 4 The effect of Ni-based alloy catalysts for DRM
Properties of metal Catalytic performance Resistance Remarks
Metal reductivity Metal dispersion Activity Selectivity Long-term stability Coke Sulphur Sintering
Mn–Ni High Low/high High High Mn could increase the coke resistance
Fe–Ni High High High The oxidation and reduction of Fe occur during DRM, which results in high coke resistance
Co–Ni High High High H2/CO High High High High O species adsorbed on Co due to the strong affinity to oxygen affinity, promoting the gasification of C species
Cu–Ni High High/low High High A slight low The optimal Cu/Ni ratio exists
Zn–Ni High High The dilution of Ni ensembles by Zn could improve the coke resistance
Mo–Ni High/ low High Mo addition to Ni/Al2O3 could lead to a decrease in activity, whereas NiMo nanocatalysts on MgO single crystal support showed excellent performance
In–Ni High Low High The excess amount of In lowers the activity drastically
Sn–Ni Low High Sn atoms occupy the C nucleation site, but the excess amount of Sn lowers the activity drastically
Ru–Ni High High/low Low H2/CO High High Ru atoms occupy the more active sites, resulting in the lower activity and high coke resistance
Rh–Ni High Rh–Ni catalyst shows enhanced methane cracking and coke gasification
Pt–Ni Low High High Pt reduces the activity of CH4 decomposition


Many catalysts show higher coke resistance than that of Ni, with enhanced long-term stability by blocking the more active sites of Ni, such as edge sites or step sites. This feature also reduces the activation of methane, but in some alloy catalysts, it has also been noted that the total activity can be increased because of the effects of improved metal reducibility and dispersibility. In addition, some ternary catalysts reportedly show better catalytic performance than bimetallic catalysts show. It is expected that more good catalysts will be reported as research progresses. For details, please refer to 5.3, “others”.

5.1 Addition of the base metal to Ni catalysts

The addition of transition metals to Ni catalysts often changes activity and stability depending on the amount added. For example, Ni–Cu shows high activity and durability at the optimum Cu/Ni ratio, but carbon deposition tends to occur at higher or lower ratios. It is necessary to add an optimal amount of a second metal because it can not only enhance stability. It can also decrease the CH4 activity drastically.
Mn–Ni. Recently, Najfach et al.,97 and Ramezani et al.98 reported effects of Mn to Ni catalysts. They described effects of increased Ni dispersibility, which led to less coke formation. Ramezani et al.98 described that the Mn could increase the activity because of the higher dispersion of Ni and concluded that 10 wt%Ni–3wt%Mn/Al2O3 is the optimal catalyst, showing long-term stability for 20 h. Najfach et al.97 investigated Mn–Ni catalysts on various zeolite supports. They reported that NH4-ZSM5 and NH4-Y zeolites have benefits of Mn–Ni such as decreasing the coke formation, decreasing the Ni particle size, and increasing the Ni dispersibility, although the addition of Mn generally engenders decreasing activity.
Fe–Ni. Many studies have investigated effects of Fe–Ni on DRM. Tomishige et al.99 published a minireview on Ni–Fe alloy catalysts for reforming hydrocarbon including DRM in 2017. In recent years, Song et al.100 investigated a Fe–Ni catalyst supported on Al2O3. Also, Margossian et al.,101 Kim et al.102 and Theofanidis et al.103 investigated Ni–Fe catalysts on Mg(Al)O. Gunduz-Meric et al.104 investigated the core–shell NiFe@SiO2 catalysts. In addition, Theofanidis et al.105 reported improved stability and activity by addition of Pd to Ni–Fe/MgAl2O4. Furthermore, DFT calculation on the Ni2Fe overlayer of Ni(111) surface was conducted by Xu et al.106 and by Ray et al.107

Generally, the reaction mechanism on Ni–Fe catalysts has been demonstrated as presented below.100

 
Fe0 + xCO2xCO + FeOx(9)
 
CH4 → 2H2 + Cα(10)
 
FeOx + Cα → Fe0 + CO(11)
Fe0 metal is oxidized to FeOx by CO2. Then carbon species Cα is formed by CH4 decomposition. Finally, FeOx and Cα react to form Fe0 and CO. Furthermore, Kim et al.102 reported structural change during this oxidation–reduction mechanism (Fig. 12). Fe is partially oxidized to FeO and partially de-alloyed to form Ni-rich NiFe alloy; Fe migrates and preferentially forms FeO on the surface. The FeO reacts with Cα and produces CO. The reduced Fe restores the original Ni–Fe alloy. This dynamic mechanism occurring during the DRM reaction provides superior resistance for coke formation.


image file: d2cy00066k-f12.tif
Fig. 12 The schematic image of reaction mechanism of Ni–Fe alloy. Reprinted with permission from ref. 102. Copyright 2017 American Chemical Society. https://pubs.acs.org/doi/abs/10.1021/jacs.6b11487.

Song et al.100 reported higher activity and improved coke resistance on Ni–Fe catalysts. Fig. 13 presents the result of activity tests and O2-TPO of 15.7wt%Ni/Al2O3 and 9.5wt%Ni–4.2wt%Fe/Al2O3 catalyst after their low-temperature stability tests (723 K, 20 h). The peaks attributed to Cα, Cβ and Cγ are assigned according to the oxidation temperature. The formation of Cγ (inactive graphic carbon) was suppressed on the Fe–Ni catalyst. Furthermore, TEM measurements of spent catalysts confirmed that filamentous carbon was not formed on Fe–Ni/Al2O3, although it was formed on Ni/Al2O3.


image file: d2cy00066k-f13.tif
Fig. 13 The results of long-term stability tests over 15.7wt%Ni/Al2O3 and 9.5wt%Ni–4.2wt%Fe/Al2O3 at low temperature (450 °C, 20 h). (a) CH4 conversion (b) CO2 conversion (c) H2/CO ratio, and (d) O2-TPO spectra of spent catalysts. Reprinted with permission from ref. 100. Copyright 2020 American Chemical Society. https://pubs.acs.org/doi/10.1021/acs.iecr.0c01204.

Margossian et al.101 investigated the effect of Ni–Fe ratios and reduction temperature of NiFe nanoparticles (3–4 nm) supported by Mg(Al)O prepared by the colloid synthesis. They concluded that NiFe catalysts with a Ni/Fe ratio of 3 reduced at 923 K showed higher stability and activity than Ni catalysts.

In addition, oxide supports containing Fe have been investigated, such as Ni–Fe–Al mixed oxides,108 FexNiyMg1−xyO,109 Ni/MgFexAl2−xO4,110,111 La0.9Sr0.1Ni1−xFexO3,112 La0.6Sr0.2Ti0.85Ni0.15O3−δ113,114 and LaxFe1−xNi0.1O3−δ.115 For these catalysts, the formation of Ni–Fe alloys was confirmed by the precipitation of metallic particles because of hydrogen reduction. Higher resistance against carbon deposition was reported.

Co–Ni. Many papers describing the effects of Co–Ni alloys on DRM have been published during the last five years (Table 5). The Co–Ni alloy catalysts show high activity not only at high temperatures but also at low temperatures,116 high durability to carbon, and improved S tolerance.117 Many reports have described the existence of an optimal Co/Ni ratio, but the value varies depending on the catalyst. In addition, in situ XPS,116 and in situ scanning transmission X-ray microscopy (STXM)118,119 studies have been conducted recently.
Table 5 Summary of Co–Ni catalysts reported in recent years
Catalysts Preparation Reaction condition Activity Sel. Stability Remarks Ref.
10wt%Co–5wt%Ni/γ-Al2O3 Co-impregnation 650–700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 36 L h−1 gcat−1 X CH4, XCO2 = 67–71%, 700 °C H2 yield >50% Stable for 4 h E a = 36–40 kJ mol−1, the amorphous carbon was dominant on spent catalyst 126
5wt%Ni–5wt%Co/γ-Al2O3 Step-wise impregnation 350–700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]N2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]8, 300[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 = 76.0% at 700 °C H2/CO = 0.93 Stable for 60 h Active at low temp. It prevents side reactions 117
12wt%NiO–3wt%Co3O4/γ-Al2O3 Thermo gravimetric method 500–700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 12[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 = 53% Stable for 5 h with carbon deposition TGPR has advantages of the reduction of the number of synthesis stages 127
X CO2 = 58%
6wt%Ni@meshed–xwt%Co/γ-Al2O3 Wet-impregnation, 450–650 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 X CH4 = 70% Stable for 70 h with a little decrease of activity Mesh-like Co coating structure showed high activity and coking resistance 128
75% at 650 °C
NiCoOx/γ-Al2O3 (no data for Ni loading) Reverse micellar method 800 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]N2= 7[thin space (1/6-em)]:[thin space (1/6-em)]9.5[thin space (1/6-em)]:[thin space (1/6-em)]83.5 X CH4 = 86% Stable for 2 h In situ STXM were conducted 119, 120
X CO2 = 77%
5wt%Ni–5wt%Co/FSP-Al2O3 Co-impregnation 500–800 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]N2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1, 144 L h−1 gcat−1 Co; a high affinity for the removal of carbon species by oxidation, Ni; highly active for CH4 decomposition 129
5wt%Ni–5wt%Co/θ-Al2O3 700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]Ar = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1, 6000 h−1 X CH4 = 75% H2/co = 1.3 Deactivation occurred The rapid decrease of activity due to phase transition 130
X CO2 = 82%
2.5wt%Ni–2.5wt%Co/Al2O3–ZrO2 700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]N2 = 17[thin space (1/6-em)]:[thin space (1/6-em)]17[thin space (1/6-em)]:[thin space (1/6-em)]11, 18[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 = 91% H2/CO = 0.96 Stable for 28 h High metal dispersion of Ni–Co catThe catalyst was calcined at 973 K 131
2.5wt%Ni–2.5wt%Co/mesoporous SiO2 Deposition–precipitation technique 700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]He = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]8, 50[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4, XCO2 ≈ 90% H2/CO ≈ 1 Stable for 200 h The coexistence of nano-sized Co and Ni in the mesopore resulted in better stability 124
Ni–Co/SiO2 with various Ni/Co ratios, total 10 wt% Hydrothermal method with urea 750 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]He = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1, 60[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 ≈ 84% H2/CO ≈ 0.8 Stable for 100 h Co could enhance the metal–support interaction catalysts with Ni/Co = 7/3 showed high and stable activity, but others did not 132
X CO2 ≈ 86% at 1023 K
Co–Ni/SiO2, total 10 wt% 750 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]N2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1, 36[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 = 87% H2/CO = 0.84 Stable for 12 h Catalysts with Ni/Co = 4/1 showed the highest activity 133
X CO2 = 94% at 1023 K
Ni–Co/SBA-15 with various Ni/Co ratios, total 10 wt% Urea co-precipitation 800 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 72[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 ≈ 86–87% H2/CO ≈ 1 Stable for 50 h 9wt%Ni–1wt%Co was the bestCatalysts with Ni/Co >1 showed improved activity and stability due to the high dispersion of metals, the synergetic effect of Ni and Co, the confinement effect of SBA-15 mesoporous channels 134
X CO2 ≈ 93–94%
(xNiyCo)/SBA-15 total 5 wt% Modified co-impregnation 700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 30[thin space (1/6-em)]000–240[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 ≈ 40% H2/CO ≈ 0.6 Stable for 100 h with a slow deactivation 4.5wt%Ni–0.5wt%Co was the best composition, which showed high stability for 100 h 135
X CO2 ≈ 65%
Co–Ni/CeO2 with various Ni/Co ratios, total 10 wt% Incipient wetness co-impregnation 800 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]N2 = 3[thin space (1/6-em)]:[thin space (1/6-em)]3[thin space (1/6-em)]:[thin space (1/6-em)]4, 12[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 = 80% H2/CO > 0.6 Stable for 10 h The catalysts with Co/Ni = 0.8 (atomic ratio) showed the highest activity and stability 136
X CO2 = 85%
1.8wt%Co–1.2wt%Ni/Ce0.75Zr0.25O2 Deactivation due to coke deposition was faster on Ni–Co catalyst supported on Ce0.75Zr0.25O2 than that on a mixture of CeO2 and ZrO2 137
1.8wt%Co–1.2wt%Ni/Ce0.8Zr0.2O2/β-SiC Deposition precipitation with urea 750 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 12 L h−1 gcat−1 X CH4 ≈ 65–70% H2/CO = 0.82 Stable for 550 h By anchoring of Co–Ni/CZO on β-SiC, sintering, and oxidation of metals could be suppressed, resulting in high stability 138
X CO2 ≈ 70–75%
Ni–Co/Ce–Zr, Ce–La (9.2 atomic % of Ni) Deposition precipitation with urea 775 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 20–30 ppm sulfur Stable for sulfur The addition of Co could improve the sulfur tolerance 118
Co–Ni/H-ZrO2 with various Ni/Co ratios, total 5 wt% Co-impregnation 700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]Ar = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]9, 256 L h−1 gcat−1 X CH4 = 92.8% H2/CO = 0.8 Stable for 6 h with a little decrease of activity H-ZrO2 is a mesoporous ZrO2 hollow sphere 125
0.89wt%Co–3.83wt%Ni/H-ZrO2 was the best
X CO2 = 93%
Carbon deposition on Ni can be oxidized by oxygen on Co
4wt%Ni–1wt%Co/CF-La2O3 with various Ni/Co ratios Ultrasonic-assisted impregnation 740 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]N2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1, 14[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 = 88% H2/CO = 1.1 Stable for 10 h CF (carbon nanofiber) has a high surface area (1393 m2 g−1) 139
X CO2 = 83%
Resistance to Ni-sintering
12.5wt%Ni–2wt%Co/CeO2–ZnAl2O4 Co-precipitation 700 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 18[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 = 76% H2/CO = 0.99 Stable for 8 h with coke deposition Increasing surface area, smaller metal particles, stronger active phase/support interaction, higher coke resistance were observed by Co addition 122
2.8wt%Ni–2.8wt%Co/HAP (HAP; Ga10(PO4)6(OH)2) Successive incipient wetness impregnation 700–750 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2[thin space (1/6-em)]:[thin space (1/6-em)]N2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3 X CH4 = 73% H2/CO = 0.9 Stable for 160 h with an initial decrease of activity Carbon nanotubes were observed on spent catalysts 140
X CO2 = 79% at 750 °C
Ni–Co catalyst from La(CoxNi1−x)0.5Fe0.5O3, total 5 wt% Sol–gel self-combustion 750 °C, CH4[thin space (1/6-em)]:[thin space (1/6-em)]CO2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 12[thin space (1/6-em)]000 mL h−1 gcat−1 X CH4 = 70% Stable for 30 h Reduced perovskite catalysts with x = 0.10 or 0.30 showed structural stability, the highest activity, and the most stability for coke deposition 123
X CO2 = 80% at 750 °C


In general, Co has a strong affinity to oxygen; O species preferentially adsorb on Co, which has the effect of promoting the gasification of C species.120–123 Sheng et al.124 elucidated the mechanism of CO formation by the reaction of the generated carbon with active oxygen using XPS, XRD, CO2-TPD, FTIR, etc. (Fig. 14).


image file: d2cy00066k-f14.tif
Fig. 14 Schematic image of reaction mechanism on Ni–Co alloy. Reprinted with permission from ref. 124. Copyright 2019 American Chemical Society. https://pubs.acs.org/doi/10.1021/acsami.9b05822.

In addition to high resistance toward carbon deposition, the durability to sulphur has been reported by Jiang et al.117 As shown in Fig. 15, 0.24wt%Ni–0.24wt%Co catalysts supported on ZrO2–CeO2 showed drastically improved tolerance for sulphur. They suggested that the Ni–Co interaction is important for the development of sulphur tolerance and discussed the possibility that modification of the electronic state of Ni by Co in the small metal cluster inhibits the adsorption of H2S.


image file: d2cy00066k-f15.tif
Fig. 15 Activity of Ni–Co catalysts at 1048 K with sulphur feed. Co could improve the sulphur resistance. Reprinted with permission from ref. 117. Copyright 2021 Elsevier.

Wu et al.116 reported high activity, stability, and selectivity at high and low temperatures of 5 wt%Ni–5wt%Co supported on γ-Al2O3 catalysts prepared by stepwise impregnation method. The Ni–Co/γ-Al2O3 catalyst showed high activity and high H2/CO ratio, even at temperatures as low as 623 K. The catalyst maintained near-equilibrium conversion and high H2/CO ratio at 1073 K for 60 h. Strong metal–support interaction contributed to the suppression of sintering. The small metal particle size contributed to the suppression of carbon deposition. In situ XPS analysis of the Ni–Co catalyst in a reaction atmosphere at different temperatures showed that the metal was reduced as the reaction temperature increased, the cobalt content on the surface decreased, and the nickel content increased (Fig. 16).


image file: d2cy00066k-f16.tif
Fig. 16 The changes of surface Ni and co compositions of Ni–Co/Al2O3 with increasing temperature was changed, which was determined by in situ XPS. Reprinted with permission from ref. 116. Copyright 2019 American Chemical Society. https://pubs.acs.org/doi/10.1021/acscatal.8b02821.

Results of EXAFS analysis revealed the presence of lattice strain in the Ni–Co alloy, confirming strong interaction between Ni and Co atoms. The lattice-strained Ni–Co has a strong CO2 dissociation ability, which is responsible for the excellent activity of the DRM reaction, especially in the low-temperature region. In addition, H2-TPD experiments revealed that hydrogen desorption is enhanced greatly on Ni–Co catalysts, which is a factor suppressing the RWGS reaction.

Askari et al.118,119 investigated Ni–Co alloy catalysts using the technique of in situ STXM and showed a different result of in situ XPS study by Wu et al.116 Askari et al.118 first prepared NiCoOx/γ-Al2O3 catalyst by the modified reverse micellar method, of which compositions are determined by EDX as 56.4 atom% of O, 1.1 atom% of Co, 9.2 atom% of Ni, and 33.3 atom% of Al. Later, STXM studies of NiCoOx/γ-Al2O3 catalyst were conducted under Ar, 5%H2/Ar, and DRM atmosphere. These results are presented in Fig. 17 and 18. The fresh sample shows inhomogeneous distribution of Ni and Co, forming NiO and Co3O4 (Fig. 17(a)). Reduction of the catalysts under a 5 wt%H2/Ar atmosphere at increased temperature revealed that full reduction of the metal finished at 773 K. Furthermore, at 773 K, voids were observed as shown in Fig. 17(b), resulting from reconstruction because of oxygen removal from the lattice. At 973 K, although the metallic states were unchanged, the structure changed drastically. The core–shell structure was formed. The shell was made entirely of Ni. The core is made of Ni and Co Fig. 17(c). Then, STXM under DRM was conducted. Results obtained 1 min later and 120 min later are presented in Fig. 18. These elemental composition maps and the L3-edge spectra showed that the metallic states and the core–shell structure remained during the 120 min DRM reaction. However, the Ni concentration increment was observed at specific locations on the surface, as indicated by the red arrows. In addition, after 120 min under DRM conditions, the formation of Ni-rich branches on the particle surface was observed, which might be a result of filamentous carbon on the surface.


image file: d2cy00066k-f17.tif
Fig. 17 Elemental mapping of Ni–Co catalysts for (a) fresh calcinated catalyst, (b) catalyst during a reduction under 5%H2/Ar at 773 K, and (c) catalyst during a reduction under 5%H2/Ar at 973 K. reprinted with permission from ref. 118. Copyright 2020 American Chemical Society. https://pubs.acs.org/doi/10.1021/acscatal.9b05517.

image file: d2cy00066k-f18.tif
Fig. 18 Elemental mapping of Ni–Co catalysts for (a) catalyst after H2 reduction, (b) catalyst after 1 min of DRM reaction, (c) catalyst after 120 min of DRM reaction. Reprinted with permission from ref 118. Copyright 2020 American Chemical Society. https://pubs.acs.org/doi/10.1021/acscatal.9b05517.
Sn–Ni. Guharoy et al.140 and da Silva et al.141 investigated the effects of Sn addition to Ni catalysts. Both concluded that the optimal amount of Sn can suppress coke formation but that excess Sn lowered the activity considerably because Sn atoms occupy the C nucleation sites on Ni and increase the energy barrier for coke nucleation.140 In addition, DFT calculations and catalytic tests have indicated that Sn promotes the oxidation of key reaction intermediates.140
Cu–Ni. Investigations have also elucidated some aspects of Cu–Ni catalysts supported on different oxides supports. Song et al.142 investigated Ni–Cu/Mg(Al)O catalyst. Chatla et al.143 used experimentation and DFT calculation to assess Ni–Cu/Al2O3 catalyst. Rezaei et al.144 investigated Ni–Cu/Al2O3 catalysts using a microchannel reactor. In addition, Han et al.145 investigated the effects of Ni/Cu ratios in Ni–Cu nanoparticles on SiO2 supports.

Song et al.142 reported that the amounts of coke formation varied depending on the Cu/Ni ratios. An optimal value exists. The Cu–Ni/Mg(Al)O catalyst with the ratio of Cu/Ni = 0.25–0.5 showed high activity and high stability, whereas the catalysts with higher or lower Cu/Ni ratios deactivated rapidly because of coke formation. Results suggest that the decrease in CH4 decomposition and increase in CO2 dissociation attributable to Ni–Cu alloying contributed to suppression of coke deposition for the optimized Ni–Cu catalyst. However, Chatla et al.143 reported Ni–Cu/Al2O3 with the ratio Cu/Ni = 1/8 as the best. They concluded that formation of Ni–Cu alloys enhanced the reducibility of NiO and that it enhanced suppression of coke formation. Furthermore, DFT calculation revealed that the higher energy barrier for carbon adsorption makes it easier to remove of the deposited carbon species.

Han et al.145 prepared the Ni–Cu nanoparticles (approximately 5 nm) supported on SiO2 supports by electrostatic adsorption and investigated the effects of Cu/Ni ratios on activities and stabilities. They described that Ni–Cu/SiO2 catalyst showed higher activity and less coke formation than Ni/SiO2, but a little sintering was observed. The catalytic performance varied depending on the Cu/Ni ratios, as shown in Fig. 19. Results showed that the catalyst with the ratio of Cu/Ni = 1/8 exhibited the best performance. Additionally, they suggested the reaction mechanism on Ni–Cu catalysts based on their research and others (Fig. 19), i.e., first, CH4 molecules are activated on the Ni surface and CHx and H are formed, second, CO2 molecules are activated on Cu and CO and O are formed. Finally, CHx species from CH4 and O species from CO2 react on the interface between Ni and Cu, making CO and H2.


image file: d2cy00066k-f19.tif
Fig. 19 (Left) The effect of Cu/Ni ratios on CH4 conversion. (Right) The suggested reaction mechanism on N–Cu catalysts. Reprinted with permission from ref. 145. Copyright 2021 American Chemical Society. https://pubs.acs.org/doi/10.1021/acsanm.1c00673.

In addition, Wang et al.146 reported effects of Cu-substitution on carbon resistance using La2(Ni1−xCux)O4 perovskite prepared using a sol–gel method. The La2Ni0.8Cu0.2O4 after H2 treatment showed a negligible amount of carbon deposition. The CH4 conversion and CO2 conversion were, respectively, 73% and 80%. They concluded that the smaller metal particles of reduced perovskite and the divided Ni ensembles by Cu contributed to the suppression of coke formation.

Zn–Ni. Two groups147,148 investigated the effect of Zn for ZnO–Al2O3 composites supports. Sokolov et al.147 found that Zn-rich supports showed higher activity than Al-rich supports, which they attributed to formation of NiZn alloy or Ni3ZnC0.7. Nataj et al.148 found that although the addition of Zn decreased the surface area, ZnO interfered with formation of the NiAl2O4 spinel phase, thereby maintaining Ni reducibility. They concluded that catalysts with low Zn/Al ratios were less active and that they were deactivated because of large amounts of carbon deposition, whereas supports with Zn/Al = 1 and 2 remained highly active for a long time. That result was attributed to dilution of the Ni ensemble by Ni–Zn alloy formation.
Mo–Ni. Yao et al.149 investigated the effects of Mo addition to Ni/Al2O3 catalyst. They concluded that an addition of Mo engenders lowering of the activity because of the weak interaction between NiO species and Al2O3, the formation of MoNi4 phase by reduction of NiMoO4 species, and the low basicity of Mo-doped catalysts. In addition, Yao et al.150 prepared a yolk–shell structure (xNi@yMo-HSS, HSS; hollow silica structure). The Mo atoms form a SiMoOx shell rather than NiMo alloy. Also, SiMoOx species enhanced the electron cloud density of Ni and the acidity of the support, resulting in increased CH4 activation. However, the activity of NiMo@HSS was low because of the decrease in the electron cloud density of metal and increased metal particle sizes of NiMo alloy particle.

Song et al.151 prepared Ni–Mo nanocatalysts on single-crystalline MgO (NiMoCat) and reported that the catalyst maintained high activity over 850 h. They first prepared a highly crystalline MgO solid: 3.76wt%Ni and 1.76wt%Mo were loaded to MgO using a polyol-mediated reductive growth method in the presence of a size-limiting polyvinylpyrrolidone (PVP) polymer surfactant. Although Mo itself was not active against DRM, it can enhance the catalytic activity.

The NiMo catalyst maintained high activity during activity tests of over 850 h. No carbon deposition was observed. Moreover, the H2/CO ratio was almost 1. On the fresh catalyst, the average size of the supported metal was 2.88 nm, but it grew to 17.30 nm within 1 h at 1073 K with flowing of reaction gas. Even after the long-term activity test, the particle size remained the same value at around 17 nm. Fig. 20 presents TEM images and illustrations of Ni–Mo particles of fresh and spent catalysts. Song et al.151 concluded that the fine particles migrate to the high-energy step edges of crystalline MgO(111) during the reaction, forming stable and persistent particles with average size of 17 nm. Covering the high-energy step edges can be expected to prevent further sintering while eliminating risks of MgO participation in the catalytic reaction.


image file: d2cy00066k-f20.tif
Fig. 20 The schematic image of NanoCatalysts on single crystal edges (NOSCE) technique on NiMoCat. Reprinted with permission from ref. 151. Copyright 2020 AAAS.
In–Ni. Liu et al.152 prepared and investigated the Ni-In intermetallic alloy covered by SiO2. Even the sample with only 0.1wt%In showed excellent coking resistance in the DRM reaction. Results obtained from H2-TPR revealed that the addition of In led to decreased reduction temperature. The optimum In loading amount was 0.5 wt% considering the balance between carbon-resistance and DRM reactivity; low In loading tended to decrease the coking resistance, whereas high In loading decreased the catalytic activity because of the formation of InNi3C0.5 species. The increased electron density of Ni by electron transfer from In might weaken the ability of Ni to activate C–H bonds and reduce the CHx cracking process, resulting in the suppression of carbon deposition.

Károlyi et al.153 similarly investigated a 2wt%In–3wt%Ni catalyst supported on SiO2. They found that diluting the Ni surface can suppress carbon deposition.

5.2 Addition of noble metals to Ni catalysts

This chapter summarizes the effects of adding a small amount of noble metals (Ru, Rh, Re, Pt) to supported Ni catalysts on DRM.
Ru–Ni. Small loadings (<1 wt%) of Ru on Ni catalysts on various supports have been investigated. Álvarez et al.154 investigated the 0.5wt%Ru–15wtNi catalyst supported on MgO/Al2O3. Luisetto et al.155 investigated 0.5wt%Ru–10wt%Ni catalyst supported on γ-Al2O3. Also, Wysocka et al.156 investigated 1wt%Ru-7wt%Ni on various supports (SiO2, Al2O3, MgAl2O4, ZrO2). Álvarez et al.154 and Luisetto et al.155 found that the Ru can improve long-term stability by suppressing coke formation. Luisetto et al.155 explained that the property of Ru which favours carbon gasification contributes to inhibition of coke formation. Álvarez et al.154 claimed that Ru atoms were located on the surface of more active sites, such as step or edge sites because of the low solubility between Ni and Ru. They described that this effect of Ru might be the reason for the lessened activity of Ru-doped catalysts. Luisetto et al.155 stated that Ru can improve not only stability, but also the activity, by keeping the Ni metal reduced. Wysocka et al.156 reported that the activity order as Ru–Ni/Al2O3 > Ru–Ni/MgAl2O4 > Ru–Ni/ZrO2 > Ru–Ni/SiO2. This order is the same as that of Ni-loaded catalysts. Also, Ru–Ni/Al2O3 showed long-term stability for 6 h. They also described that introduction of Ru enhanced methane dissociation, causing the low H2/CO ratio.

As described in chapter 3 and in 4.2, the Ru–Ni section of chapter 4, Yoon et al.60 investigated carbon deposition using DFT on Ni and Ni–Ru alloys. They claimed that, to predict the catalytic efficiency and carbon deposition properly, one must investigate not only the dissociation into CH3, CH2, CH, C, and H, but also the CO gas evolution reaction (details were presented in chapter 3). The theoretical calculation showed that Ni–Ru catalysts had more stability than either Ni or Ni–Rh.

In addition, Zhou et al.157 prepared RuxNiyMg1−xyO catalysts using the solvothermal synthesis. The catalysts obtained using the direct reduction method were highly active and stable. Actually, Ru was found to change the type of carbon deposition from a persistent graphitic one that can only be removed by O2 to a soft system that can be readily removed by CO2. Furthermore, Ru increased the activation barrier of CH4 dissociation, thereby slowing the carbon deposition rate.

Rh–Ni. Mozammel et al.158 prepared RhNi, NiCo, and NiCoRh catalysts supported on mesoporous alumina (MAl). These alloy catalysts showed better stability and conversion than the respective mono-metal catalysts. Among these catalysts, 4.5wt%Rh–4.5wt%Ni/MAl showed the highest activity and stability because of the enhanced methane cracking and carbon gasification on alloy surface. They also found that nickel formed a homogeneous alloy phase with cobalt, but a heterogeneous bimetallic phase with rhodium. Alloying with cobalt moderated coking but alloying with rhodium promoted carbon gasification through hydrogen spillover effects. Furthermore, Rh-supported catalysts were less active than Rh-alloyed catalysts, indicating that Rh alloying increased the activity of monometallic catalysts by alloying.

Romano et al.159 reported that the addition of noble metals to Ni catalysts increased the reducibility of the metal, but it had no significant effect on TOF or stability. They prepared Ni, Rh, and Pd supported catalysts and Ni-based alloy catalysts with a few noble metals by the incipient wetness impregnation. They conducted screening tests. Fig. 21 shows H2-TPR results for the Ni monometallic catalysts and Ni-based alloy catalysts. In fact, Ni/Al2O3 shows a NiO reduction peak with a peak top at 711 K and a Ni–O–Al reduction peak with a peak top at 838 K. However, the reduction temperature of the alloy catalyst with a small amount of noble metal was much lower. Only one peak was observed for the Rh–Ni catalyst. Enhancement of the Ni reducibility by the addition of noble metals is attributable to the higher ability to dissociate hydrogen.


image file: d2cy00066k-f21.tif
Fig. 21 The H2-TPR spectra of various Ni monometallic catalysts and Ni-based alloy catalysts. Reprinted with permission from ref. 159. Copyright 2021 Elsevier.
Re–Ni. Zubenko et al.160 prepared perovskite precursors based on lanthanum ferrite (LaFeO3), which were subsequently doped with Ni and Re phases. Fig. 22 shows a schematic image of the possible structure of catalysts by co-exsolution from the ceramic precursor. For this study, Re-alloy nanoparticles (Ni–Fe, Re–Fe, Re–Ni–Fe) were formed from composite precursors under reducing conditions. They were found to be active and stable under dry reforming conditions. The Re nanoparticles were not highly active up to 1173 K, whereas the Ni nanoparticles were more active with the addition of Re, even at low temperatures. It was concluded that, because of the strong catalyst–support interaction, the deposition of carbon, the sintering of nanoparticles, and the evaporation of the Re-containing phase were limited.
image file: d2cy00066k-f22.tif
Fig. 22 The schematic image of the possible structure of co-exsolution of B and B′ from an ordered ceramic precursor image file: d2cy00066k-t5.tif by H2 reduction. Reprinted with permission from ref. 160. Copyright 2017 Elsevier.
Pt–Ni. Egelske et al.161 prepared a series of Ni–Pt catalysts supported on γ-Al2O3 with 5.0 wt%Ni and up to 2.8wt%Pt by the electroless deposition. Then they evaluated their catalytic performance for DRM. At temperatures below 873 K, the activity of 0.4 layers of Pt was maximal. In fact, it exceeded that of undeposited Ni. However, at 973 K, all samples were deactivated. At high temperatures (973 K), the alloy phase of the nanoparticles separated, which was consistent with the phase diagram of the bulk. The Pt-rich ensemble promoted the activation of CH4 to coke causing deactivation.

Vasiliades et al.162 intensively studied details of DRM activity and carbon deposition of Pt–Ni alloy catalysts. DRM was conducted at 1023 K using 5 wt%Ni, 0.5wt%Pt, and their binary alloy (5wt%Ni–0.5wt%Pt) supported on Ce0.8Pr0.2O2−δ support. The origin of the deposited carbon during the DRM was quantified using isotopically labelled 13CO2 as the feed gas. Results showed that the carbon deposition on supported Ni and Ni–Pt alloys was mainly attributable to the CH4 decomposition on the metal surface, whereas that on supported Pt was attributable to the CO disproportionation reaction. The Ni–Pt catalyst was found to have a significantly lower rate of carbon deposition despite a smaller decrease in the CH4 conversion rate compared to the monometallic Ni catalyst. The Ni–Pt catalyst showed overall excellent performance and stability after 50 h of use at 1023 K DRM, with a low amount of deposited carbon of 0.38 wt%.

5.3 Others

In recent years, catalysts supported with not only two-metal but also three-metal and Co-based alloy catalysts have been developed.
Trimetallic catalysts.
Co–Fe–Ni. Joo et al.52 reported that exsolved Co–Ni–Fe ternary alloy supported on perovskite PrBaMn1.7Co0.1Ni0.2O5+δ had enhanced activity compared to either Co–Ni or Ni supported catalysts. They calculated the d-band centres of Co–Ni–Fe, Co–Ni, and Ni, and concluded that the upshift of the d-band centre can be attributed to the higher activity of Co–Ni–Fe.
Co–Cu–Ni. Ghanbarabadi and Khoshandam126 investigated Ni–Co, Ni–Cu, bimetallic nanocatalysts, and Ni–Co–Cu trimetallic nanocatalysts supported on γ-Al2O3. Their order of activity was Ni–Co/γ-Al2O3, Ni–Co–Cu/γ-Al2O3, and Ni–Cu/γ-Al2O3. Moreover, results showed that Ni–Co was more active than Ni–Co–Cu. However, the TPO results demonstrated that the amount of coke formation in Ni–Co–Cu was reduced by the presence of Cu.
Co–Mn–Ni. Kim et al.163 investigated the tri-metallic, reduced LaNi0.34Co0.33Mn0.33O3 perovskite catalysts. The catalyst showed higher activity and stability with less coke formation than either LaNi0.5Co0.5 or LaNiO3. Results showed that Mn can enhance the stability and that Co can enhance the activity.
Co–Ru–Ni. Aramouni et al.164 investigated Co–Ni–Ru trimetallic catalysts supported on Al(Mg)O with various Co/Ni ratios and calcination temperatures. The high Ni/Co ratio was good for activity and stability. The high oxygen affinity of Co was effective for inhibiting carbon deposition. The addition of Ru also improved the stability and coke resistance of the tested Ni–Co catalysts, although the activity at high-temperature calcination was reduced slightly.
Fe–Cu–Ni. Jin et al.165 investigated the trimetallic NiFeCu catalysts supported on Mg(Al)O with various Cu ratios. Results showed that Cu addition improved the coke resistance of Ni–Fe catalysts, with Ni3Fe1Cu1 catalyst being the most durable.
Fe–Pd–Ni. Theofanidis et al.105 reported enhanced activity and stability of the Pd-added Ni–Fe/MgAl2O4 catalyst. Core–shell structures, of which the core is Fe–Ni and the shell is Fe–Ni–Pd, were observed after the reduction. Because of the structure, the segregation of Fe on the surface can be suppressed.
Co-based catalysts. Supported Co catalysts have also been studied in DRM because of their low cost and large-scale availability.166 Bradford and Vannice167 provide a summary of Co catalysts on various supports. Furthermore, San-José-Alonso et al.168 reported higher activity of Co or Co-rich Ni catalysts for DRM because of the high ability of Co for CH4 decomposition, but much carbon deposition was observed on these catalysts. Consequently, Co catalysts are promising catalysts, but they are necessary for additional development. However, Co catalysts have not been studied as much as Ni and noble metal catalysts have been. This section summarizes information related to Co-based alloy catalysts that have been reported in recent years.

Xie et al.169 reported synergetic effects of Co and Pt on DRM activity and H2/CO selectivity. Using the co-impregnation method, Co–Pt alloy, of which the Pt/Co molar ratio is 1/3, supported on CeO2 was prepared. During the reaction, the Pt and Co maintained their metallic state, but there was slight oxygen decoration, resulting in oxygen–metal site pairs (O*–*). Kinetic studies and DFT calculation revealed that the catalytic surface of PtCo/CeO2 with O*-modified formation promotes the activation of C–H bonds. In addition, PtCo/CeO2 catalysts were resistant to coke formation compared to Pt/CeO2. They can be regenerated easily by a mild CO2 treatment. In addition, Itkulova et al.166 investigated Co catalysts with a small amount of Pt supported on Al2O3–ZrO2 for DRM and bi-reforming. Results indicate that 5 wt%Co–Pt(95[thin space (1/6-em)]:[thin space (1/6-em)]5)/Al2O3–ZrO2 showed high stability and no sintering or coke formation observed during the 100 h activity tests.

Perovskite-based catalysts LaCoO3 and LaCu0.55Co0.45O3 were prepared and investigated by Touahra et al.170 The LaCu0.55Co0.45O3 catalyst showed higher catalytic performance and higher carbon resistance in DRM than the LaCoO3 catalyst. The high catalytic performance of the LaCu0.55Co0.45O3 catalyst was attributed to the strong Cu–Co interaction observed after reduction, which suppressed carbon formation and particle aggregation. In addition, the carbon species formed on the Co sites are well removed by the oxygen species generated from La2O2CO3.

6. Summary and perspectives

This paper summarizes recent research on alloys in SRM and DRM, mainly emphasizing Ni-based catalysts. The Ni catalyst is promising as an industrial catalyst because of its low cost and high activity, but the deactivation of the catalyst because of carbon deposition is an issue. Although noble metal catalysts are superior to Ni catalysts in terms of their activity and durability, they are difficult to industrialize because of their high costs. Therefore, the mainstream development is designed to improve the catalytic performance of Ni catalysts by adding various secondary metals. Actually, Ni-based alloys of many various types have been reported, including Ni-based alloy catalysts with base metals (Mn, Fe, Co, Cu, Zn, Mo, In, Sn, etc.) and Ni-based alloy catalysts with noble metals (Ru, Rh, Ag, Re, Ir, Pt, Au, etc.). Even noble metal doped catalysts are reported to improve performance, in many cases by a small amount of noble metal doping, which does not lead to high cost. Alloying Ni catalysts enables to succeed in developing an ideal catalyst that is both inexpensive and has excellent catalytic performance.

As for catalyst preparation, many catalysts were prepared with metals supported on oxide supports such as Al2O3, whereas others were catalysts with Ni or Ni-based alloy nanoparticles deposited on the surface by reduction of oxides (such as perovskite oxides) containing Ni and secondary metals in their composition. The alloy effects were reported to improve the dispersion and reducibility of the supported metal, to change the catalytic performance such as activity and selectivity, and to improve the durability against carbon deposition, sulphur poisoning, and sintering, and so on. Tables 3 (for SRM) and 4 (for DRM) summarize the Ni-based alloy catalyst effects reported herein. However, alloy effects often vary with catalysts, compositions, and reaction conditions. Some papers present different results. Therefore, it is noteworthy that Tables 3 and 4 show only representative features of a few alloy catalysts presented herein.

This section presents a brief summary of the effects of the addition of secondary metals by reviewing Tables 3 and 4.

(1) Properties of metal

Numerous reports have described that the addition of a secondary metal changes the properties, such as reducibility and dispersibility, of the supported metal. The improvement of reducibility is remarkably pronounced for Ni catalysts doped with noble metals, as shown in Fig. 21, where marked reduction of the reduction temperature was observed by adding Rh or Pd to Ni catalysts.159 The high hydrogen dissociation capacity of the noble metals might be a factor in the high activity of the catalysts. The addition of a second metal was also found to improve the Ni dispersion in several alloys (Mn–Ni,97,98 Co–Ni,116 Rh–Ni,76 Ir–Ni,83 Pt–Ni,84etc.). As discussed in chapter 3, the Ni particle size has a significant effect on the ease of carbon deposition. Actually, carbon deposition is less likely to occur on highly dispersed small Ni particles. Therefore, alloying to reduce the Ni particle size is very effective for inhibiting carbon deposition.

(2) Catalytic performance

Whereas the activity of the Ni catalyst decreases during the reaction, most papers surveyed in this report have described the improved stability of many alloy catalysts after long-term stability tests (Tables 3 and 4). However, regarding the activity, some reports (especially for precious metal) have described that the addition of a second metal increased the activity, although others did not. The DFT calculations of some alloy surfaces showed that the CH4 dissociation barrier increases compared to Ni (Au–Ni,61 Ir–Ni,83 Pt–Ni,42 Ru–Ni,157etc.). Some experimental reports showed that TOF increases or that activation energy increases (Sn–Ni,66 Ag–Ni,80 Pt–Ni,42etc.). In addition, a decrease in activity because of occupation of more active, or low-coordinated Ni sites by second metals (Au–Ni,48 Ag–Ni,82 Ru–Ni,154 Co–Ni,65 Sn–Ni,66,140etc.) was reported. Consequently, probably methane activity tends to decrease with the addition of the second metal. However, some reports have described that the overall activity increased because of the effect of improving the reducibility and dispersibility of the metal by adding a secondary metal (Au–Ni,86 and other precious metal addition). In addition, the change of selectivity was reported for some alloy catalysts for DRM (Cu–Ni,72 Rh–Ni,76 and Re–Ni81 for SRM; and Co–Ni116 and Ru–Ni155).

(3) Stability

In addition to the ability of alloy catalysts to inhibit carbon deposition, their abilities to inhibit sulphur poisoning and sintering were investigated. First, almost all reports have described that alloying improved the inhibition of carbon deposition and that it showed stable activity in long-term tests (Tables 3 and 4). For example, in the Fe–Ni system,100 the type of deposited carbon was evaluated by O2-TPO. Stable and difficult to react species of carbon was decreased by alloying (Fig. 13). Some papers have explained that the reason for the suppression of carbon deposition as the second metal blocks the more active site of Ni, such as edge sites or step sites, thereby preventing carbon nucleation. Other factors such as a decrease in the Ni particle size, dilution of the Ni ensemble, and decrease in stability of the C species on the alloy surface (e.g. Au–Ni,61 Cu–Ni143) were also reported. In addition, a mechanism was proposed for Co–Ni in which O is adsorbed preferentially onto Co; it reacts with C species formed on Ni, leading to CO gasification (Fig. 14).124 However, many reports have described that either too little or too much was not beneficial. For example, Sn–Ni can suppress carbon deposition, but too much Sn–Ni engenders a significant decrease in SRM activity (Fig. 12).73 Cu–Ni shows high activity and durability at the optimum Cu/Ni ratio, but too much or too little Cu–Ni engenders rapid deactivation of activity because of severe carbon deposition.143

Some studies have assessed durability against sulphur, such as Fe–Ni67 and Au–Ni88 for SRM, and Co–Ni118 for DRM (Fig. 15). Furthermore, the sintering was reportedly suppressed for Rh–Ni77 and Ir–Ni77 in SRM, and Co–Ni116 in DRM.

Although it is gratifying to see that almost all the alloy catalysts reportedly have improved durability against carbon deposition, apparently some room remains for additional research and development of catalysts with sulphur resistance, sintering suppression, and high activity at low temperatures. In addition, only a few studies have tested various alloy catalysts under the same conditions. It is still difficult to discuss which alloy catalyst has the best catalytic performance. Future studies should be undertaken to develop catalysts that not only have durability against carbon deposition, but which also have resistance to sulphur and sintering, and high activity even at low temperatures. To develop such materials, more research must be done to elucidate bi-metallic and tri-metallic catalysts, with consideration of the various properties of the respective metals. For such purpose, high-throughput screening system for multi-metal alloy catalysts and high-performance calculation will be important in future.

Conflicts of interest

There are no conflicts to declare.

Notes and references

  1. L. Liu and A. Corma, Chem. Rev., 2018, 118, 4981–5079 CrossRef CAS PubMed.
  2. M. Zhou, C. Li and J. Fang, Chem. Rev., 2021, 121, 736–795 CrossRef CAS PubMed.
  3. V. Ponec, Appl. Catal., A, 2001, 222, 31–45 CrossRef CAS.
  4. R. Ferrando, J. Jellinek and R. L. Johnston, Chem. Rev., 2008, 108, 845–910 CrossRef CAS PubMed.
  5. S. Furukawa and T. Komatsu, ACS Catal., 2017, 7, 735–765 CrossRef CAS.
  6. T. Zhang, A. G. Walsh, J. Yu and P. Zhang, Chem. Soc. Rev., 2021, 50, 569–588 RSC.
  7. R. T. Hannagan, G. Giannakakis, M. Flytzani-Stephanopoulos and E. C. H. Sykes, Chem. Rev., 2020, 120, 12044–12088 CrossRef CAS PubMed.
  8. Y. Xin, S. Li, Y. Qian, W. Zhu, H. Yuan, P. Jiang, R. Guo and L. Wang, ACS Catal., 2020, 10, 11280–11306 CrossRef CAS.
  9. D. Wu, K. Kusada, T. Yamamoto, T. Toriyama, S. Matsumura, S. Kawaguchi, Y. Kubota and H. Kitagawa, J. Am. Chem. Soc., 2020, 142, 13833–13838 CrossRef CAS PubMed.
  10. S. Furukawa, T. Komatsu and K. I. Shimizu, J. Mater. Chem. A, 2020, 8, 15620–15645 RSC.
  11. P. Liu and J. K. Nêrskov, Phys. Chem. Chem. Phys., 2001, 3, 3814–3818 RSC.
  12. M. Mohamedali, A. Henni and H. Ibrahim, ChemEngineering, 2018, 2, 1–23 CrossRef.
  13. J. N. Armor, Appl. Catal., A, 1999, 176, 159–176 CrossRef CAS.
  14. M. A. Peña, J. P. Gómez and J. L. G. Fierro, Appl. Catal., A, 1996, 144, 7–57 CrossRef.
  15. A. Iulianelli, S. Liguori, J. Wilcox and A. Basile, Catal. Rev.: Sci. Eng., 2016, 58, 1–35 CrossRef CAS.
  16. Q. Wang, X. Chen, A. N. Jha and H. Rogers, Renewable Sustainable Energy Rev., 2014, 30, 1–28 CrossRef.
  17. J. R. Rostrup-Nielsen, J. Catal., 1984, 85, 31–43 CrossRef CAS.
  18. J. G. Xu and G. F. Froment, AIChE J., 1989, 35, 88–96 CrossRef CAS.
  19. D. Li, Y. Nakagawa and K. Tomishige, Appl. Catal. A, 2011, 408, 1–24 CrossRef CAS.
  20. H. Wu, V. La Parola, G. Pantaleo, F. P. Puleo, A. M. Venezia and L. F. Liotta, Catalysts, 2013, 3, 563–583 CrossRef CAS.
  21. E. Meloni, M. Martino and V. Palma, Catalysts, 2020, 10, 352 CrossRef CAS.
  22. L. Chen, Z. Qi, S. Zhang, J. Su and G. A. Somorjai, Catalysts, 2020, 10, 858 CrossRef CAS.
  23. H. Zhang, Z. Sun and Y. H. Hu, Renewable Sustainable Energy Rev., 2021, 149, 111330 CrossRef CAS.
  24. C. Wang, Y. Wang, M. Chen, D. Liang, Z. Yang, W. Cheng, Z. Tang, J. Wang and H. Zhang, Int. J. Hydrogen Energy, 2021, 46, 5852–5874 CrossRef CAS.
  25. I. V. Yentekakis, P. Panagiotopoulou and G. Artemakis, Appl. Catal., B, 2021, 296, 120210 CrossRef CAS.
  26. Z. Bian, S. Das, M. H. Wai, P. Hongmanorom and S. Kawi, ChemPhysChem, 2017, 18, 3117–3134 CrossRef CAS PubMed.
  27. M. A. A. Aziz, H. D. Setiabudi, L. P. Teh, M. Asmadi, J. Matmin and S. Wongsakulphasatch, Chem. Eng. Technol., 2020, 43, 661–671 CrossRef.
  28. T. Yabe and Y. Sekine, Fuel Process. Technol., 2018, 181, 187–198 CrossRef CAS.
  29. V. V. Nedolivko, G. O. Zasypalov, A. V. Vutolkina, P. A. Gushchin, V. A. Vinokurov, L. A. Kulikov, S. V. Egazar'yants, E. A. Karakhanov, A. L. Maksimov and A. P. Glotov, Russ. J. Appl. Chem., 2020, 93, 765–787 CrossRef CAS.
  30. M. A. A. Aziz, H. D. Setiabudi, L. P. Teh, N. H. R. Annuar and A. A. Jalil, J. Taiwan Inst. Chem. Eng., 2019, 101, 139–158 CrossRef CAS.
  31. L. P. Teh, H. D. Setiabudi, S. N. Timmiati, M. A. A. Aziz, N. H. R. Annuar and N. N. Ruslan, Chem. Eng. Sci., 2021, 239, 116606 CrossRef.
  32. Z. Li, Q. Lin, M. Li, J. Cao, F. Liu, H. Pan, Z. Wang and S. Kawi, Renewable Sustainable Energy Rev., 2020, 134, 110312 CrossRef CAS.
  33. R. Singh, A. Dhir, S. K. Mohapatra and S. K. Mahla, Biomass Convers. Biorefin., 2020, 10, 567–587 CrossRef CAS.
  34. Z. Qin, J. Chen, X. Xie, X. Luo, T. Su and H. Ji, Environ. Chem. Lett., 2020, 18, 997–1017 CrossRef CAS.
  35. K. Wittich, M. Krämer, N. Bottke and S. A. Schunk, ChemCatChem, 2020, 12, 2130–2147 CrossRef CAS.
  36. E. Kikuchi, S. Tanaka, Y. Yamazaki and Y. Morita, Bull. Jpn. Pet. Inst., 1974, 16, 95–98 CrossRef CAS.
  37. J. R. Anderson and M. Boudart, Catalysis Science and Technology, 1984, vol. 5 Search PubMed.
  38. P. Ferreira-Aparicio, A. Guerrero-Ruiz and I. Rodríguez-Ramos, Appl. Catal., A, 1998, 170, 177–187 CrossRef CAS.
  39. I. D. Iglesias, G. Baronetti and F. Mariño, Energy Sources, Part A, 2017, 39, 129–133 CrossRef CAS.
  40. M. S. Fan, A. Z. Abdullah and S. Bhatia, ChemCatChem, 2009, 1, 192–208 CrossRef CAS.
  41. K. H. Delgado, L. Maier, S. Tischer, A. Zellner, H. Stotz and O. Deutschmann, Catalysts, 2015, 5, 871–904 CrossRef CAS.
  42. S. Roy, S. Hariharan and A. K. Tiwari, J. Phys. Chem. C, 2018, 122, 10857–10870 CrossRef CAS.
  43. P. M. Hundt, M. E. Van Reijzen, H. Ueta and R. D. Beck, J. Phys. Chem. Lett., 2014, 5, 1963–1967 CrossRef CAS PubMed.
  44. S. Nave, A. K. Tiwari and B. Jackson, J. Phys. Chem. A, 2014, 118, 9615–9631 CrossRef CAS PubMed.
  45. F. Abild-Pedersen, J. Greeley and J. K. Nørskov, Catal. Lett., 2005, 105, 9–13 CrossRef CAS.
  46. A. M. Ranjekar and G. D. Yadav, J. Indian Chem. Soc., 2021, 98, 100002 CrossRef.
  47. C. Vogt, J. Kranenborg, M. Monai and B. M. Weckhuysen, ACS Catal., 2020, 10, 1428–1438 CrossRef CAS.
  48. K. de Oliveira Rocha, C. M. P. Marques and J. M. C. Bueno, Chem. Eng. Sci., 2019, 207, 844–852 CrossRef CAS.
  49. A. Nilsson, L. G. M. Pettersson, B. Hammer, T. Bligaard, C. H. Christensen and J. K. Nørskov, Catal. Lett., 2005, 100, 111–114 CrossRef CAS.
  50. B. Hammer and J. K. Nørskov, Adv. Catal., 2000, 45, 71–129 CAS.
  51. B. Hammer and J. K. Nørskov, Surf. Sci., 1995, 343, 211–220 CrossRef CAS.
  52. S. Joo, K. Kim, O. Kwon, J. Oh, H. J. Kim, L. Zhang, J. Zhou, J. Q. Wang, H. Y. Jeong, J. W. Han and G. Kim, Am. Ethnol., 2021, 19104, 15912–15919 Search PubMed.
  53. K. Ray, R. Bhardwaj, B. Singh and G. Deo, Phys. Chem. Chem. Phys., 2018, 20, 15939–15950 RSC.
  54. S. Arora and R. Prasad, RSC Adv., 2016, 6, 108668–108688 RSC.
  55. D. L. Trimm, Catal. Rev.: Sci. Eng., 1977, 16, 155–189 CrossRef CAS.
  56. J. R. Rostrup-Nielsen and D. L. Trimm, J. Catal., 1977, 48, 155–165 CrossRef CAS.
  57. D. L. Trimm, Catal. Today, 1999, 49, 3–10 CrossRef CAS.
  58. J. Rostrup-Nielsen, Catal. Today, 2006, 111, 4–11 CrossRef CAS.
  59. J. R. Rostrup-Nielsen, J. Sehested and J. K. Noerskov, Adv. Catal., 2002, 47, 65–139 CAS.
  60. Y. Yoon, H. Kim and J. Lee, J. Power Sources, 2017, 359, 450–457 CrossRef CAS.
  61. F. Besenbacher, I. Chorkendorff, B. S. Clausen, B. Hammer, A. M. Molenbroek, J. K. Nørskov and I. Stensgaard, Science, 1998, 279, 1913–1915 CrossRef CAS PubMed.
  62. Y. Liu, Y. Chen, H. Yu, F. Guan, Z. Hou, D. Cui and Y. Zhang, Catal. Today, 2021, 369, 167–174 CrossRef CAS.
  63. C. H. Bartholomew, Catal. Rev.: Sci. Eng., 1982, 24, 67–112 CrossRef CAS.
  64. J. H. Kim, D. J. Suh, T. J. Park and K. L. Kim, Appl. Catal., A, 2000, 197, 191–200 CrossRef CAS.
  65. X. You, X. Wang, Y. Ma, J. Liu, W. Liu and X. Xu, ChemCatChem, 2014, 6, 3377–3386 CrossRef CAS.
  66. E. Nikolla, J. Schwank and S. Linic, J. Catal., 2009, 263, 220–227 CrossRef CAS.
  67. M. V. Tsodikov, S. S. Kurdymov, G. I. Konstantinov, V. Y. Murzin, O. V. Bukhtenko and Y. V. Maksimov, Int. J. Hydrogen Energy, 2015, 40, 2963–2970 CrossRef CAS.
  68. G. I. Konstantinov, S. S. Kurdyumov, Y. V. Maksimov, O. V. Bukhtenko and M. V. Tsodikov, Catal. Ind., 2018, 10, 1–8 CrossRef.
  69. A. Djaidja, H. Messaoudi, D. Kaddeche and A. Barama, Int. J. Hydrogen Energy, 2015, 40, 4989–4995 CrossRef CAS.
  70. H. Provendier, C. Petit and A. Kiennemann, C. R. Acad. Sci., Ser. IIc: Chim., 2001, 4, 57–66 CrossRef CAS.
  71. R. Thalinger, M. Gocyla, M. Heggen, R. Dunin-Borkowski, M. Grünbacher, M. Stöger-Pollach, D. Schmidmair, B. Klötzer and S. Penner, J. Catal., 2016, 337, 26–35 CrossRef CAS.
  72. T. J. Huang, T. C. Yu and S. Y. Jhao, Ind. Eng. Chem. Res., 2006, 45, 15–156 CrossRef.
  73. E. Nikolla, J. Schwank and S. Linic, J. Catal., 2007, 250, 85–93 CrossRef CAS.
  74. M. Nawfal, C. Gennequin, M. Labaki, B. Nsouli, A. Aboukaïs and E. Abi-Aad, Int. J. Hydrogen Energy, 2015, 40, 1269–1277 CrossRef CAS.
  75. S. C. Baek, K. W. Jun, Y. J. Lee, J. D. Kim, D. Y. Park and K. Y. Lee, Res. Chem. Intermed., 2012, 38, 1225–1236 CrossRef CAS.
  76. S. Katheria, G. Deo and D. Kunzru, Appl. Catal., A, 2019, 570, 308–318 CrossRef CAS.
  77. F. Morales-Cano, L. F. Lundegaard, R. R. Tiruvalam, H. Falsig and M. S. Skjøth-Rasmussen, Appl. Catal., A, 2015, 498, 117–125 CrossRef CAS.
  78. D. Li, T. Shishido, Y. Oumi, T. Sano and K. Takehira, Appl. Catal., A, 2007, 332, 98–109 CrossRef CAS.
  79. A. H. Dam, H. Wang, R. Dehghan-Niri, X. Yu, J. C. Walmsley, A. Holmen, J. Yang and D. Chen, ChemCatChem, 2019, 11, 3401–3412 CrossRef CAS.
  80. H. Wang, D. W. Blaylock, A. H. Dam, S. E. Liland, K. R. Rout, Y.-A. Zhu, W. H. Green, A. Holmen and D. Chen, Catal. Sci. Technol., 2017, 7, 1713–1725 RSC.
  81. Y. Xu, T. Hirano, H. Kunieda, Y. Hara and Y. Miyata, Catal. Sci. Technol., 2020, 10, 2004–2019 RSC.
  82. L. Wang, Bull. Mater. Sci., 2020, 43, 1–3 CrossRef.
  83. Z. Liu, F. Gao, Y. A. Zhu, Z. Liu, K. Zhu and X. Zhou, Chem. Commun., 2020, 56, 13536–13539 RSC.
  84. V. K. Jaiswar, S. Katheria, G. Deo and D. Kunzru, Int. J. Hydrogen Energy, 2017, 42, 18968–18976 CrossRef CAS.
  85. X. Li, D. Li, H. Tian, L. Zeng, Z. J. Zhao and J. Gong, Appl. Catal., B, 2017, 202, 683–694 CrossRef CAS.
  86. S. Palma, L. F. Bobadilla, A. Corrales, S. Ivanova, M. A. Centeno and J. A. Odriozola, Appl. Catal., B, 2014, 144, 846–854 CrossRef CAS.
  87. Y. H. Chin, D. L. King, H. S. Roh, Y. Wang and S. M. Heald, J. Catal., 2006, 244, 153–162 CrossRef CAS.
  88. F. M. Sapountzi, C. Zhao, A. Boréave, L. Retailleau-Mevel, D. Niakolas, C. Neofytidis and P. Vernoux, Catal. Sci. Technol., 2018, 8, 1578–1588 RSC.
  89. Z. T. Wang, M. T. Darby, A. J. Therrien, M. El-Soda, A. Michaelides, M. Stamatakis and E. C. H. Sykes, J. Phys. Chem. C, 2016, 120, 13574–13580 CrossRef CAS.
  90. Q. Shen, Y. Jiang, F. Xia, B. Wang, X. Lv, W. Ye and G. Yang, Pet. Sci. Technol., 2020, 38, 618–625 CrossRef CAS.
  91. R. Manabe, S. Okada, R. Inagaki, K. Oshima, S. Ogo and Y. Sekine, Sci. Rep., 2016, 6, 3–4 CrossRef PubMed.
  92. S. Okada, R. Manabe, R. Inagaki, S. Ogo and Y. Sekine, Catal. Today, 2018, 307, 272–276 CrossRef CAS.
  93. M. Torimoto, K. Murakami and Y. Sekine, Bull. Chem. Soc. Jpn., 2019, 92, 1785–1792 CrossRef CAS.
  94. M. Torimoto, S. Ogo, Y. Hisai, N. Nakano, A. Takahashi, Q. Ma, J. G. Seo, H. Tsuneki, T. Norby and Y. Sekine, RSC Adv., 2020, 10, 26418–26424 RSC.
  95. A. Takahashi, R. Inagaki, M. Torimoto, Y. Hisai, T. Matsuda, Q. Ma, J. G. Seo, T. Higo, H. Tsuneki, S. Ogo, T. Norby and Y. Sekine, RSC Adv., 2020, 10, 14487–14492 RSC.
  96. M. Torimoto, S. Ogo, D. Harjowinoto, T. Higo, J. G. Seo, S. Furukawa and Y. Sekine, Chem. Commun., 2019, 55, 6693–6695 RSC.
  97. A. J. Najfach, C. B. Almquist and R. E. Edelmann, Catal. Today, 2021, 369, 31–47 CrossRef CAS.
  98. Y. Ramezani, F. Meshkani and M. Rezaei, J. Chem. Sci., 2018, 130, 1–10 CrossRef CAS.
  99. K. Tomishige, D. Li, M. Tamura and Y. Nakagawa, Catal. Sci. Technol., 2017, 7, 3952–3979 RSC.
  100. Z. Song, Q. Wang, C. Guo, S. Li, W. Yan, W. Jiao, L. Qiu, X. Yan and R. Li, Ind. Eng. Chem. Res., 2020, 59, 17250–17258 CrossRef CAS.
  101. T. Margossian, K. Larmier, S. M. Kim, F. Krumeich, C. Müller and C. Copéret, ACS Catal., 2017, 7, 6942–6948 CrossRef CAS.
  102. S. M. Kim, P. M. Abdala, T. Margossian, D. Hosseini, L. Foppa, A. Armutlulu, W. Van Beek, A. Comas-Vives, C. Copéret and C. Müller, J. Am. Chem. Soc., 2017, 139, 1937–1949 CrossRef CAS PubMed.
  103. S. A. Theofanidis, V. V. Galvita, H. Poelman and G. B. Marin, ACS Catal., 2015, 5, 3028–3039 CrossRef CAS.
  104. G. Gunduz-Meric, S. Kaytakoglu and L. Degirmenci, Int. J. Hydrogen Energy, 2020, 45, 34547–34556 CrossRef CAS.
  105. S. A. Theofanidis, V. V. Galvita, M. Sabbe, H. Poelman, C. Detavernier and G. B. Marin, Appl. Catal., B, 2017, 209, 405–416 CrossRef CAS.
  106. L. Xu, H. Wen, X. Jin, Q. Bing and J. Liu, Appl. Surf. Sci., 2018, 443, 515–524 CrossRef CAS.
  107. K. Ray, A. S. Sandupatla and G. Deo, Int. J. Quantum Chem., 2021, 121, 1–7 CrossRef.
  108. Y. Kim, H. S. Lim, M. Lee and J. W. Lee, Catal. Today, 2021, 368, 86–95 CrossRef CAS.
  109. T. Zhang, Z. Liu, Y. A. Zhu, Z. Liu, Z. Sui, K. Zhu and X. Zhou, Appl. Catal., B, 2020, 264, 118497 CrossRef CAS.
  110. S. A. Theofanidis, V. V. Galvita, H. Poelman, N. V. R. A. Dharanipragada, A. Longo, M. Meledina, G. Van Tendeloo, C. Detavernier and G. B. Marin, ACS Catal., 2018, 8, 5983–5995 CrossRef CAS.
  111. H. Wang, N. V. Srinath, H. Poelman, C. Detavernier, P. Li, G. B. Marin and V. V. Galvita, Catal. Sci. Technol., 2020, 10, 6987–7001 RSC.
  112. S. Das, S. Bhattar, L. Liu, Z. Wang, S. Xi, J. J. Spivey and S. Kawi, ACS Catal., 2020, 10, 12466–12486 CrossRef CAS.
  113. S. Joo, O. Kwon, S. Kim, H. Y. Jeong and G. Kim, J. Electrochem. Soc., 2020, 167, 064518 CrossRef CAS.
  114. S. Joo, A. Seong, O. Kwon, K. Kim, J. H. Lee, R. J. Gorte, J. M. Vohs, J. W. Han and G. Kim, Sci. Adv., 2020, 6, 1–9 Search PubMed.
  115. S. Shah, S. Sayono, J. Ynzunza, R. Pan, M. Xu, X. Pan and K. L. Gilliard-AbdulAziz, AIChE J., 2020, 66, 1–12 CrossRef.
  116. Z. Wu, B. Yang, S. Miao, W. Liu, J. Xie, S. Lee, M. J. Pellin, D. Xiao, D. Su and D. Ma, ACS Catal., 2019, 9, 2693–2700 CrossRef CAS.
  117. C. Jiang, E. Loisel, D. A. Cullen, J. A. Dorman and K. M. Dooley, J. Catal., 2021, 393, 215–229 CrossRef CAS.
  118. A. Beheshti Askari, M. Al Samarai, B. Morana, L. Tillmann, N. Pfänder, A. Wandzilak, B. Watts, R. Belkhou, M. Muhler, M. Muhler and S. Debeer, ACS Catal., 2020, 10, 6223–6230 CrossRef CAS PubMed.
  119. A. Beheshti Askari, M. Al Samarai, N. Hiraoka, H. Ishii, L. Tillmann, M. Muhler and S. Debeer, Nanoscale, 2020, 12, 15185–15192 RSC.
  120. D. Harshini, Y. Kwon, J. Han, S. P. Yoon, S. W. Nam and T.-H. Lim, Korean J. Chem. Eng., 2010, 27, 480–486 CrossRef CAS.
  121. A. Movasati, S. M. Alavi and G. Mazloom, Fuel, 2019, 236, 1254–1262 CrossRef CAS.
  122. H. Wang, X. Dong, T. Zhao, H. Yu and M. Li, Appl. Catal., B, 2019, 245, 302–313 CrossRef CAS.
  123. S. Das, M. Sengupta, A. Bag, M. Shah and A. Bordoloi, Nanoscale, 2018, 10, 6409–6425 RSC.
  124. K. Sheng, D. Luan, H. Jiang, F. Zeng, B. Wei, F. Pang and J. Ge, ACS Appl. Mater. Interfaces, 2019, 11, 24078–24087 CrossRef CAS PubMed.
  125. T. J. Siang, S. Singh, O. Omoregbe, L. G. Bach, N. H. H. Phuc and D. V. N. Vo, J. Energy Inst., 2018, 91, 683–694 CrossRef CAS.
  126. H. Ghanbarabadi and B. Khoshandam, Int. J. Environ. Anal. Chem., 2021 DOI:10.1080/03067319.2020.1862812 , in press.
  127. K. Cao, M. Gong, J. Yang, J. Cai, S. Chu, Z. Chen, B. Shan and R. Chen, J. Catal., 2019, 373, 351–360 CrossRef CAS.
  128. J. Horlyck, C. Lawrey, E. C. Lovell, R. Amal and J. Scott, Chem. Eng. J., 2018, 352, 572–580 CrossRef CAS.
  129. X. Zhang, Z. Vajglova, P. Mäki-Arvela, M. Peurla, H. Palonen, D. Y. Murzin, S. A. Tungatarova, T. S. Baizhumanova and Y. A. Aubakirov, ChemistrySelect, 2021, 6, 3424–3434 CrossRef CAS.
  130. A. S. Al-Fatesh, J. K. Abu-Dahrieh, H. Atia, U. Armbruster, A. A. Ibrahim, W. U. Khan, A. E. Abasaeed and A. H. Fakeeha, Int. J. Hydrogen Energy, 2019, 44, 21546–21558 CrossRef CAS.
  131. Z. Bian and S. Kawi, J. CO2 Util., 2017, 18, 345–352 CrossRef CAS.
  132. G. Gunduz-Meric, S. Kaytakoglu and L. Degirmenci, React. Kinet., Mech. Catal., 2020, 129, 403–419 CrossRef CAS.
  133. J. Xin, H. Cui, Z. Cheng and Z. Zhou, Appl. Catal., A, 2018, 554, 95–104 CrossRef CAS.
  134. H. Wu, J. X. Liu, H. Liu and D. He, Fuel, 2019, 235, 868–877 CrossRef CAS.
  135. Y. Turap, I. Wang, T. Fu, Y. Wu, Y. Wang and W. Wang, Int. J. Hydrogen Energy, 2020, 45, 6538–6548 CrossRef CAS.
  136. M. A. Vasiliades, P. Djinović, A. Pintar, J. Kovač and A. M. Efstathiou, Catal. Sci. Technol., 2017, 7, 5422–5434 RSC.
  137. P. Djinović and A. Pintar, Appl. Catal., B, 2017, 206, 675–682 CrossRef.
  138. Y. Wei, M. Song, L. Yu and F. Meng, Process Saf. Environ. Prot., 2021, 145, 236–246 CrossRef CAS.
  139. T. S. Phan, A. R. Sane, B. Rêgo de Vasconcelos, A. Nzihou, P. Sharrock, D. Grouset and D. Pham Minh, Appl. Catal., B, 2018, 224, 310–321 CrossRef CAS.
  140. U. Guharoy, E. Le Saché, Q. Cai, T. R. Reina and S. Gu, J. CO2 Util., 2018, 27, 1–10 CrossRef CAS.
  141. F. de A. R. da Silva, R. C. R. dos Santos, R. S. Nunes and A. Valentini, Appl. Catal., A, 2021, 618, 118129 CrossRef CAS.
  142. K. Song, M. Lu, S. Xu, C. Chen, Y. Zhan, D. Li, C. Au, L. Jiang and K. Tomishige, Appl. Catal., B, 2018, 239, 324–333 CrossRef CAS.
  143. A. Chatla, M. M. Ghouri, O. W. El Hassan, N. Mohamed, A. V. Prakash and N. O. Elbashir, Appl. Catal., A, 2020, 602, 117699 CrossRef CAS.
  144. R. Rezaei, G. Moradi and S. Sharifnia, Energy Fuels, 2019, 33, 6689–6706 CrossRef CAS.
  145. K. Han, S. Wang, Q. Liu and F. Wang, ACS Appl. Nano Mater., 2021, 4, 5340–5348 CrossRef CAS.
  146. M. Wang, T. Zhao, M. Li and H. Wang, RSC Adv., 2017, 7, 41847–41854 RSC.
  147. S. Sokolov, J. Radnik, M. Schneider and U. Rodemerck, Int. J. Hydrogen Energy, 2017, 42, 9831–9839 CrossRef CAS.
  148. S. M. Masoom Nataj, S. M. Alavi and G. Mazloom, J. Chem. Technol. Biotechnol., 2019, 94, 1305–1314 CrossRef CAS.
  149. L. Yao, M. E. Galvez, C. Hu and P. Da Costa, Int. J. Hydrogen Energy, 2017, 42, 23500–23507 CrossRef CAS.
  150. L. Yao, M. E. Galvez, C. Hu and P. Da Costa, Ind. Eng. Chem. Res., 2018, 57, 16645–16656 CrossRef CAS.
  151. Y. Song, E. Ozdemir, S. Ramesh, A. Adishev, S. Subramanian, A. Harale, M. Albuali, B. A. Fadhel, A. Jamal, D. Moon, S. H. Choi and C. T. Yavuz, Science, 2020, 367, 777–781 CrossRef CAS PubMed.
  152. W. Liu, L. Li, S. Lin, Y. Luo, Z. Bao, Y. Mao, K. Li, D. Wu and H. Peng, J. Energy Chem., 2022, 65, 34–47 CrossRef.
  153. J. Károlyi, M. Németh, C. Evangelisti, G. Sáfrán, Z. Schay, A. Horváth and F. Somodi, J. Ind. Eng. Chem., 2018, 58, 189–201 CrossRef.
  154. M. A. Álvarez, L. F. Bobadilla, V. Garcilaso, M. A. Centeno and J. A. Odriozola, J. CO2 Util., 2018, 24, 509–515 CrossRef.
  155. I. Luisetto, C. Sarno, D. De Felicis, F. Basoli, C. Battocchio, S. Tuti, S. Licoccia and E. Di Bartolomeo, Fuel Process. Technol., 2017, 158, 130–140 CrossRef CAS.
  156. I. Wysocka, J. Hupka and A. Rogala, Catalysts, 2019, 9, 540 CrossRef CAS.
  157. H. Zhou, T. Zhang, Z. Sui, Y. A. Zhu, C. Han, K. Zhu and X. Zhou, Appl. Catal., B, 2018, 233, 143–159 CrossRef CAS.
  158. T. Mozammel, D. Dumbre, R. Hubesch, G. D. Yadav, P. R. Selvakannan and S. K. Bhargava, Energy Fuels, 2020, 34, 16433–16444 CrossRef CAS.
  159. P. N. Romano, J. F. S. de Carvalho Filho, J. M. A. R. de Almeida and E. F. Sousa-Aguiar, Catal. Today, 2021 DOI:10.1016/j.cattod.2021.08.006 , in press.
  160. D. Zubenko, S. Singh and B. A. Rosen, Appl. Catal., B, 2017, 209, 711–719 CrossRef CAS.
  161. B. T. Egelske, J. M. Keels, J. R. Monnier and J. R. Regalbuto, J. Catal., 2020, 381, 374–384 CrossRef CAS.
  162. M. A. Vasiliades, C. M. Damaskinos, K. K. Kyprianou, M. Kollia and A. M. Efstathiou, Catal. Today, 2020, 355, 788–803 CrossRef CAS.
  163. W. Y. Kim, J. S. Jang, E. C. Ra, K. Y. Kim, E. H. Kim and J. S. Lee, Appl. Catal., A, 2019, 575, 198–203 CrossRef CAS.
  164. N. A. K. Aramouni, J. Zeaiter, W. Kwapinski, J. J. Leahy and M. N. Ahmad, Fuel, 2021, 300, 120950 CrossRef CAS.
  165. F. Jin, Y. Fu, W. Kong, J. Wang, F. Cai, J. Zhang and J. Xu, Chem. Phys. Lett., 2020, 750, 137491 CrossRef CAS.
  166. S. S. Itkulova, Y. Y. Nurmakanov, S. K. Kussanova and Y. A. Boleubayev, Catal. Today, 2018, 299, 272–279 CrossRef CAS.
  167. M. C. J. Bradford and M. A. Vannice, Catal. Rev.: Sci. Eng., 1999, 41, 1–42 CrossRef CAS.
  168. D. San-José-Alonso, J. Juan-Juan, M. J. Illán-Gómez and M. C. Román-Martínez, Appl. Catal., A, 2009, 371, 54–59 CrossRef.
  169. Z. Xie, B. Yan, S. Kattel, J. H. Lee, S. Yao, Q. Wu, N. Rui, E. Gomez, Z. Liu, W. Xu, L. Zhang and J. G. Chen, Appl. Catal., B, 2018, 236, 280–293 CrossRef CAS.
  170. F. Touahra, R. Chebout, D. Lerari, D. Halliche and K. Bachari, Energy, 2019, 171, 465–474 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2022
Click here to see how this site uses Cookies. View our privacy policy here.