Next Article in Journal
Biodiesel Production from Waste Frying Oil (WFO) Using a Biomass Ash-Based Catalyst
Previous Article in Journal
Design and Performance of CuNi-rGO and Ag-CuNi-rGO Composite Electrodes for Use in Fuel Cells
Previous Article in Special Issue
Synthesis of Propiolic and Butynedioic Acids via Carboxylation of CaC2 by CO2 under Mild Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Properties, Industrial Applications and Future Perspectives of Catalytic Materials Based on Nickel and Alumina: A Critical Review

1
Dipartimento di Ingegneria Civile, Chimica e Ambientale, Università di Genova, Via Opera Pia 15, 16145 Genova, Italy
2
Consorzio Interuniversitario Nazionale di Scienza e Tecnologia dei Materiali (INSTM), UdR Genova, Via Dodecaneso 33, 16146 Genova, Italy
3
Dipartimento di Chimica e Chimica Industriale, Università di Genova, Via Dodecaneso 33, 16146 Genova, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(8), 552; https://doi.org/10.3390/catal14080552
Submission received: 31 July 2024 / Revised: 14 August 2024 / Accepted: 19 August 2024 / Published: 22 August 2024
(This article belongs to the Special Issue Feature Papers in "Industrial Catalysis" Section)

Abstract

:
The bulk and surface properties of materials based on nickel and aluminum oxides and hydroxides, as such or after reduction processes, are reviewed and discussed critically. The actual and potential industrial applications of these materials, both in reducing conditions and in oxidizing conditions, are summarized. Mechanisms for reactant molecule activation are also discussed.

1. Introduction

Metals represent a main family of catalytic materials, with particular relevance in the fields of hydrogenation and dehydrogenation reactions, including steam reforming, as well as total and partial oxidation [1]. As is well known, platinum group metals (PGMs), mainly platinum itself and palladium, have extraordinarily high catalytic and electrocatalytic activities, much more than typical transition metals such as nickel, cobalt, iron and copper. For this reason, they are applied in many chemical and electrochemical technologies, some of which are strictly needed for the next energetic transition to obtain decarbonization or defossilization for our civilization [2]. However, PGMs are relatively rare elements, and their costs are extremely high. For these reasons, they are becoming critical metals [3] and may be unsustainable for use in some technologies.
Among non-PGM transition metals, nickel is frequently the most active one. It is by far cheaper than PGMs and may represent, for this reason, the best or at least second-best choice after PGMs for many catalytic [4] and electrocatalytic technologies [5].
As is well known, while bulk metal powders can be used sometimes in catalysis, metals are supported by different carriers in most cases. This allows one to reduce the amount of metal used, but it also frequently induces new and different properties to the metal itself. Among the carriers largely used in heterogeneous catalysis, aluminas are relatively cheap, strong and reactive materials that allow for the strong dispersion of supported species and the production of metal nanoparticles whose catalytic activities are frequently significantly superior to those of micrometric or millimetric unsupported powders and whose thermal stability is also significantly superior to that of unsupported nanoparticles. For these reasons, alumina-supported nickel catalysts have, for many years, represented versatile and very active heterogeneous catalytic systems [6]. On the other hand, materials based on nickel and alumina may have very different behaviors depending on the type of alumina used and the relative amount of nickel and alumina, as well as the oxidation state, pretreatments, dispersion, etc. In this critical review, we will summarize the literature data, in part arising from our own experience, concerning materials based on nickel and alumina, both at the laboratory level and concerning the industrial applications of these materials. We will take into consideration catalytic systems where Ni species and alumina or metal aluminates play a major role as the active phase and as the “support”. In Table 1, the ΔH°298 (gas) values (calculated per mol of reactant molecule from the enthalpy of formation data [7]) and the practical conditions of the main reactions considered here are summarized.

2. Solid-State Chemistry of Nickel and Aluminum-Based Catalytic Materials

2.1. Oxide and Hydroxide Phases

2.1.1. Ni Oxides and Hydroxides

Precipitation procedures from Ni salts may produce hydroxides [8], which are relevant materials in electrochemical technologies in particular. Two phases exist of the Ni(OH)2 compound: β-Ni(OH)2, the mineral theophrastite, is isostructural with brucite, Mg(OH)2, with Ni2+ in octahedral coordination; the second phase is actually a hydrated form of β-Ni(OH)2, α-Ni(OH)2·xH2O, with 0.41 ≤ x ≤ 0.7, consisting of brucite-type layers intercalated by water molecules. By oxidation, two polymorphs of the oxy-hydroxide of trivalent Ni form, which have a main role in electrochemistry as cathode materials of Ni–Cd and Ni–MH (nickel–metal hydride) batteries. They are denoted as β- and γ-NiO(OH). However, their structure is still incompletely known, taking also into account that they could be impure compounds [9].
A single nickel oxide phase exists in nature and is easily produced, NiO, whose mineralogical name is bunsenite. NiO crystallizes in the rock salt or periclase (MgO) structure with all octahedral Ni2+ ions and melts at 1900 °C [10]. It has a green-colored, p-type, wide bandgap (3.6–4.0 eV) semiconductor and shows antiferromagnetic behavior underneath the Neel temperature of 523 K. The antiferromagnetic behavior of NiO becomes superparamagnetic when the particle size is small, i.e., around 100 nm [10].
Higher valence binary nickel oxides also exist, although their real structure is not fully clear. In particular, “black” nickel oxide, denoted as Ni2O3 [11] or NiO1+x, can only be formed in a water-free environment at temperatures below 250 °C [12]. Both these phases would formally contain the trivalent Ni ion Ni3+ and would exist only in very strong oxidizing conditions. On the other hand, a number of trivalent nickel mixed oxides exist such as, e.g., the LaNiO3 perovskite [13] and NaNiO2, which both contain Ni3+ in octahedral coordination [14]. Among the high oxidation Ni oxide compounds, LiNiO2 and LixNiO2 oxides have been the object of investigation because of their role as electrodes in Li-ion batteries. Formally, in the parent layered LiNiO2 compound, which is nearly isostructural with LiCoO2 and LiMnO2, nickel is trivalent, becoming partially tetravalent by the removal of a part of the Li ions, which stay mostly in the interlayer region between the oxide layers, although cation mixing can also occur [15,16]. Indeed, although high-valence nickel compounds (Ni3+ and Ni4+ based) are mostly quite instable structures, they can exist in highly oxidizing environments and, in particular, in structures.

2.1.2. Aluminum Oxides and Hydroxides

The chemistry of aluminum oxides and hydroxides is a very complex one and has been the object of several reviews [17,18]. Techniques for the preparation of alumina nanoparticles, as well as their characterization, have also been the object of a more recent review [19].
In conditions where the solubility is exceeded in Al3+ water solutions, usually “gelatinous” XRD amorphous precipitates form first. Depending primarily on temperature and pH, as well as on the aging time, the nature of anions present and the copresence of organic components [20,21], different crystalline hydroxides or oxyhydroxides form [22]. At a low temperature and in excess water, the following hydroxides are preferentially formed: bayerite α-Al(OH)3 if 5.8 < pH < 9 or gibbsite γ-Al(OH)3 for pH < 5.8 and for pH > 9. At temperatures higher than about 80 °C, the oxyhydroxides become thermodynamically more stable than the trihydroxides and thus tend to form. Boehmite γ-AlOOH, or the low crystallinity form pseudoboehmite, is most easily formed at atmospheric pressure while the production of diaspore α-AlOOH usually needs higher pressures. The additional Al(OH)3 polymorphs norstrandite and doyleite, as well as the oxyhydroxide akdalaite, or tohdite, whose formula is Al5O7(OH), can be prepared only in particular conditions.
The thermal decomposition of hydroxides and oxyhydroxides gives rise to the many different alumina phases. The thermodynamically stable phase α-Al2O3 (corundum) can be directly synthesized by the decomposition of diaspore α-AlOOH at about 500 °C as a medium-surface area powder (65 m2/g [23]) or by the treatment of any Al oxide or hydroxide above 1000 °C. In the latter case, it is usually prepared as a stable low-surface-area ceramic abrasive and refractory material (Tmelt 2040 °C). In catalysis, low-surface-area α-Al2O3 powders are used as inert and thermally stable supports of catalysts [24]. Higher-surface-area corundum powders can also be prepared by a resistive hotspot-induced phase transformation of γ-Al2O3 [25].
The most commonly used material in catalysis is cubic γ-Al2O3, which is usually prepared by the calcination of gelatinous pseudoboehmite or by crystalline boehmite at around 450 °C. When treated at a slightly higher temperature, the crystal structure can become slightly distorted and is denoted as δ-Al2O3. Much less used is the similar polymorph denoted as η-Al2O3, which is usually prepared by the decomposition of bayerite at 250–300 °C. These three slightly different materials, both constituted by cation defective spinel structures [26], can have a surface area of 150–300 m2/g, are stable until near 500 °C, and represent the most common supports for metal and sulfide catalysts working at temperatures < 500 °C. They are also active acid or acid–base catalysts as well as very active adsorbents at both vapor/solid and liquid/solid interfaces. The three cation defective spinel structures, when treated at 700–900 °C, give rise to θ-Al2O3, which has a typical surface area of 70–110 m2/g and is stable until ~900 °C. It represents the most common support for metal catalysts for moderately high-temperature reactions. Its monoclinic structure, the same of β-Ga2O3, is also derived from the structure of spinel. A further heat treatment of θ-Al2O3 at about 1000 °C gives rise to the progressive formation of low-surface-area corundum powders or mixed θ,α-Al2O3 composite powders.
It can be taken into consideration that many other carrier materials are currently available for supported catalyst manufacturing, such as other metal oxides (silica, titania, zirconia, ceria, magnesia, etc.) and carbon materials. However, for vapor-phase reactions in particular, alumina and silica are usually the preferred catalyst supports by catalyst producers because of their moderate costs, their well-known chemistry and their well-proven stabilities. In particular, aluminas are characterized by their dispersion ability, which allows them to tune supported metal particle sizes with loading and support surface area optimization.

2.1.3. Mixed Ni-Al Hydroxide and Oxide Phases

The basification of Ni2+ and Al3+-containing water solutions produces the coprecipitation of mixed hydroxides in a wide range of Ni:Al ratios. In the absence of carbonate ions, Al can enter the structures of both β- and α-Ni(OH)2 [27,28]. With Ni:Al ratio 3 and in the presence of carbonate ions, a crystalline precipitate with the formula Ni6Al2CO3(OH)16·4H2O, analogous to the mineral takovite, can be obtained. This is the Ni analogue of hydrotalcite, Mg6Al2CO3(OH)16·4H2O, the best known among layered-double-hydroxide anionic clays [29,30]. With lower Ni:Al ratios, the precipitates look essentially amorphous, maybe with the presence of takovite too. This methodology for the preparation of Ni/Al2O3 catalysts, described in the scientific literature by J. Ross [31], allows one to manufacture, by further calcination, a homogeneous series of catalytic materials that are characterized in a large compositional range of spinel/rock-salt NiO/NiAl2O4/Al2O3 solid solutions [32,33].
By calcining the precipitate with Ni:Al ratio = 0.5, the spinel-phase NiAl2O4 can be prepared [33,34,35], which is the only thermodynamically stable intermediate phase in the NiO-Al2O3 phase diagram [36]. In contrast with what happens with spinel MgAl2O4, the Jahn Teller effect, due to the electronic structure of Ni2+, pushes it in octahedral coordination. Thus, Ni aluminate cation distribution is that of an “inverted” spinel, with nearly half of the Al3+ ions in tetrahedral coordination and both the second half of Al3+ ions and the large predominance of Ni2+ ions in octahedral coordination [37].
The partial solubility of alumina into a Ni aluminate spinel is known to produce Al-rich non-stoichiometric spinels that are thermodynamically stable at a high temperature [36]. These spinels are essentially solid solutions of the stoichiometric inverse spinel NiAl2O4 and the defective spinel γ-Al2O3, which have partial solubility. In metastable phases, a continuous series of Ni-deficient spinels can be prepared [33,38], including samples with composition NiAl4O7 [39] and the NiAl10O16 phase [40]. Interestingly, the presence of even small amounts (0.1 mol%) of NiO in spinel alumina hinders the phase transformation to α-Al2O3 [41], thus stabilizing spinel-type aluminas.
On the other hand, Al-deficient or Ni-rich spinels can also be produced, such as Ni2Al2O5 [39], thanks to the reciprocal solubility of the two phases NiAl2O4 and NiO, which are both based on the cubic close packing array of oxide ions, forming cation excess non-stoichiometric spinel phases. The metastability of these Ni-rich spinels is confirmed by the segregation of NiO from NiAl2O4 upon high-temperature treatment [39].
Interestingly, some solubility of alumina into NiO exists; this is very limited in terms of thermodynamics and occurs only at a high temperature [36], but it is more abundant in metastable phases, producing, e.g., a bunsenite-type (rock-salt structure) crystalline material containing significant amounts of Al ions, e.g., with composition NiAl0.35O1.53 [33].
Similar Ni-Al mixed oxide materials can also be prepared with other techniques, such as liquid-feed flame spray pyrolysis [42] and the sol–gel technique followed by calcination [43,44,45]. It can also be taken into consideration that other coprecipitation techniques can produce less-homogeneous materials, producing rock salt and spinel phases with a Ni:Al ratio less than 2 [46].
The solid state reaction of NiO with Al2O3 has also been investigated to produce bulk NiAl2O4. It has been found that the spinel phase forms only at high temperatures, i.e., above 800 °C [47] or 1000 °C [48], when alumina spinel phases are no longer stable.

2.1.4. Deposition of NiO onto Transitional Aluminas

NiO-Al2O3 catalysts, as catalytic materials or as precursors of metallic Ni/Al2O3 catalysts, are frequently prepared by impregnation techniques (dry impregnation, wet impregnation, deposition precipitation, etc.). Common dry impregnation techniques allow for the large dispersion of Ni2+ ions (and the balancing O2− anions) over the alumina surface until the amount corresponds to a full monolayer. For typical γ-Al2O3 with a surface area of about 200 m2/g, this limit corresponds roughly to 15–20 wt% NiO on alumina for well-dispersed samples. The techniques agree, showing no NiO formed until this loading is obtained (see Figure 1 for X-Ray Diffraction (XRD)). These highly dispersed species are essentially constituted by isolated Ni2+ complexes strongly interacting with the support. They cover the strongest acid–base sites of alumina located in the edges and defects and, in a second step, the less-active sites located on the main crystal faces [49,50]. This agrees with the reduction in the number of both of the strong acid sites [51] and basic sites [52] of γ-Al2O3 by the deposition of the Ni oxide species. When the monolayer coverage is full, the formation of a NiAl2O4-like surface layer is observed [53]. The XRD pattern of the samples undergoes slight changes from that of γ-Al2O3 (Figure 1) and becomes very similar to that of bulk NiAl2O4, although the positions of the peaks, the 400 and the 440 ones in particular (near 2θ = 46° and 67°, respectively), are shifted at lower 2θ values with repect to pure γ-Al2O3 but are still at significantly higher values than those observed in the pattern of NiAl2O4. Correspondigly, the calculated unit cell parameter is < 8.0 Å, which is only slightly higher than that of γ-Al2O3 (7.887 Å) but still definitely smaller than that of NiAl2O4 (a = 8.052 Å) [54]. On the other hand, spectroscopies, scheletal infrared in particular (Figure 2) and, even more, visible and near infrared spectroscopy, allow one to distinguish such a surface layer from bulk NiAl2O4 and NiO, as shown in Figure 3. The surface NiAl2O4-like layer protects the alumina from surface area loss and sintering, and it inhibits bulk phase transformation [55]. Treatment at 600 °C and above results in a fast drop in the surface area with alumina phase transformation occurring [56].
When the “monolayer” loading limit is overcome, further loading produces NiO particles to form, which is well evident by FESEM, XRD, FTIR, UV–visible spectroscopy and XPS techniques [57] (see Figure 1 for XRD). These NiO nanoparticles, although still interacting with the alumina support, are by far more easily reducible to Ni, as is evident by the Temperature Programmed Reduction (TPR) resembling the behavior of the unsupported Ni nanoparticles. As stated, while a NiAl2O4 surface layer forms, the formation of bulk nickel aluminate does not occur until at least 600 °C. The reaction between NiO and γ-Al2O3 is still slow and incomplete, even at 1000 °C [48], when γ-Al2O3 is already converted, usually into θ,α-Al2O3 composite powders.
Impregnated materials are frequently prepared to be later reduced in hydrogen for the final manufacturing of reduced–supported Ni catalysts. It can be taken into consideration that, even though coprecipitated Ni/Al2O3-based catalysts might be a little more active than impregnated ones [58], the less complete use of nickel in coprecipitated samples (where a large part of nickel remains catalytically inactive in the bulk) and the likely simpler preparation of catalysts by the co-impregnation of commercial alumina with nickel and dopants make impregnation techniques advantageous (also from the environmental point of view) with respect to more complex coprecipitation procedures at the industrial level.

2.1.5. Deposition of NiO over α-Al2O3 and Alkali Earth Aluminates

For reactions occurring at very high temperatures, i.e., >700 °C, transitional aluminas lack the stability to be used as supports. Thus, more stable ceramic supports must be used. Among them, corundum powders can be applied [59]. Also, in this case, the deposition of Ni results in the formation of a spinel NiAl2O4 layer at even high temperatures up to 1100 °C, protecting α-Al2O3 by sintering [60,61].
Alternatively, Mg2+ can be added to the catalyst recipe. This can be done using MgAl2O4 (spinel, Tmelt 2300 °C) powder as the support [59] or preparing NiO-MgO-Al2O3 mixed oxide materials by coprecipitation techniques or co-impregnation techniques. The addition of magnesium not only allows one to increase thermal stability [62] but may also allow for a better utilization of nickel, which will remain limited at the surface of the catalyst [53]. The MgO-Al2O3 system has strong similarities with the NiO-Al2O3 system [63], except for the fact that MgAl2O4 is a normal spinel while NiAl2O4 is, as mentioned, an inverse spinel. In general, large solid-state solubility occurs in the system NiO-MgO-Al2O3 [64].
Additionally, calcium aluminates are also largely used as supports for nickel catalysts for high-temperature applications, e.g., for steam reforming and autothermal reforming (see below). A number of different structures exist in the CaO-Al2O3 system [65]. The calcium-poorer structure CaAl12O19 (hibonite) is declared by Clariant as the support of some of their catalysts [59]; it has the magnetolumbite structure, similar to the β-alumina-type layered structures with spinel-type Al2O3 layers separated by CaO-type layers. Other hexa-aluminates, such as Ca,K- [59], Ba-, Sr- and La- hexa-aluminates, may be used as catalyst supports for Ni-based catalysts [66] in high-temperature applications. Some Ni-based Johnson Matthey catalysts are instead supported on mixed calcium aluminate phases (calcium aluminate cement [67]).
The surface properties of the materials based on alkali-earth aluminates, such as MgAl2O4 [68] and BaAl12O19 [69], are similar to those of alkali-earth-doped spinel aluminas, with reduced Lewis acidity, increased basicity with respect to transitional aluminas and a slightly reduced nickel dispersion ability [70].

2.2. Metallic Phases in the Ni-Al System

Nickel crystallizes in the face-centered cubic (fcc) structure with a melting point of 1453 °C. Cubic nickel is ferromagnetic until the Curie temperature of 354 °C. However, it has been reported that the melting point is strongly reduced for nanoparticles down to 700 °C for 2 nm size particles produced by the NaBH4 reduction of Ni nitrate [71]. It must, however, be considered that nanoparticles produced with this technique may be amorphous [72] and can contain relevant amounts of boron impurities [73]. The Curie temperature for nanoparticles decreases so that they are superparamagnetic at room temperature [72,74].
Interestingly, hexagonal (hcp) nickel, instead of cubic nickel, is sometimes found in spent catalysts. This metastable structure is likely stabilized by the presence of carbide species [75] and can also be prepared by introducing small amounts of Pt in Ni [76] or be produced as a surface layer of cubic nickel particles [77]. Such a metastable phase is assumed to be more active in the field of electrocatalysis for the hydrogen evolution reaction. It is also supposed to be more active than cubic phase for steam reforming reactions [72].
Aluminum crystallizes too in the fcc structure. It is paramagnetic and has a low melting temperature of 660 °C. Six intermetallic compounds can exist in the Al–Ni system, namely, Al3Ni, Al3Ni2, Al4Ni3, AlNi, Al3Ni5 and AlNi3, which are all conducting materials and show a very strong chemical interaction between Al and Ni [78,79,80]. On the other hand, it must be taken into consideration that the reducibility of Ni is much easier than that of aluminum; this is illustrated by the standard reduction potentials π°= −1.676 V for Al3+, with respect to π°= −0.257 V for Ni2+, as well as by the results of the Temperature Programmed Reduction experiments. In fact, TPR data show that NiO is reduced to Ni by hydrogen at 300–450 °C depending on particle size [81], although reduction can be not fully complete in such low-temperature conditions [82]. Instead, Al2O3 is unreducible by hydrogen, based on thermodynamics, until potentially very high temperatures [83,84]. The TPR experiments confirm that pure aluminas do not undergo any reduction until 1000 °C. This means that, starting from Ni-Al mixed oxides, surface Ni species can be reduced selectively. It must, however, be taken into consideration that Al is soluble in fcc Ni to a considerable extent, according to phase diagrams, up to about 8% mol at 400 °C and near 20% at 1360 °C [79]. Thus, the formation of Al-doped Ni metal particles after the reduction of Ni-alumina materials is unlikely but cannot be fully excluded. To our knowledge, never has this possibility been taken into consideration.

2.3. On the Reactivity of the Metallic Ni-Based Phases

Real metal catalysis occurs when the catalyst contains an active metallic phase in real reaction conditions. It must, however, be considered that, in many cases, the elements and compounds present in the reaction medium can react with metal phases, converting them at least partially to different compound phases.

2.3.1. Oxidation of Metallic Nickel

The oxidation of metallic nickel with oxygen-containing flows to NiO is particle-size dependent. While nickel nanoparticles with a size distribution between 10 and 100 nm start to be oxidized around 200 °C and may be fully oxidized at 700 °C [85], reaction rates are slower for polycrystalline nickel, and, consequently, higher temperatures are needed [86]. The surface oxidation of nickel polycrystals in O2 is already found at 300 °C [87,88], while the NiO layer formed at T > 600 °C results in a significant passivation of its corrosion behavior [89], i.e., in a slowing of its bulk oxidation. These data indicate that in oxidizing conditions and temperatures above 300 °C or more, nickel catalysts are expected to be largely or fully oxidized.

2.3.2. Sulfidation of Metallic Nickel

As is well known, hydrocarbons coming from oil refineries but also flows coming from biomass treatments such as biogases and biomass gasification and pyrolysis, contain, or may contain, sulfur in the form of H2S and alkyl sulfides. In the presence of these compounds, metallic nickel can become sulfurized [90] and its catalytic activity reduced [91]. In fact, sulfur deposition from H2S decomposition occurs relatively fast on the surfaces of nickel-based materials due to the relatively stronger bonding of sulfur compounds and the small activation energies required for the H2S decomposition [92] producing surface sulfide species that hinder catalytic activity [91]. However, prolonged sulfidation can produce bulk nickel sulfide species. Indeed, various nickel sulfide phases exist [93] such as NiS (two polymorphic phases), Ni3+xS2, Ni3S2, Ni7S6, Ni9S8, Ni3S4 and NiS2 (two polymorphs). Studies on Ni/Al2O3 sulfidation suggest that NiS2 is formed as an intermediate phase and Ni3S2 as a final sulfide species [94].
However, it must be taken into consideration that even supported nickel catalysts may have significant metal loadings (20–60 wt%), and, in the presence of small amounts of sulfur in the feed, their sulfurization can be only partial. In fact, heavily loaded nickel catalysts are frequently preferred to low-loading noble metal catalysts just because they may have a better resistance to deactivation by sulfur due to their much higher metal surface area.

2.3.3. Reactivity with Carbon: Carburization and Formation of Carbon Nanotubes and Ni(CO)4

Nickel metal surface reactivity is also relevant with respect to carbon when in contact with carbon monoxide and hydrocarbons. Several different forms of carbon can be produced at the surface of metallic nickel: surface carbon species, amorphous layers, filamentous carbon, bulk carbide and graphitic carbon [91]. A quite fast deactivation of unsupported nickel was reported upon CO2 methanation at atmospheric pressure, coproducing CO, mainly associated with the production of encapsulating carbon [95] (see Figure 4). In conditions of the Fischer–Tropsch synthesis, the carburization of nickel is less extensive than that of other metals such as iron and cobalt, resulting in a less-efficient hydrocarbon-chain-growth mechanism [96]. This is the main reason why nickel-based catalysts are very active for methanation and less useful for Fischer–Tropsch synthesis. In fact, only a partial carburization of nickel to Ni3C was found with pure CO at 265 °C [97]. This reactivity is determined mainly by an electronic effect on the CO dissociation barrier, i.e., the increase in the dissociation barrier from left to right in the periodic table, which has been correlated with the metal d band center [98,99,100] and results in the strong CO dissociation activity of iron, the medium dissociation activity of nickel and cobalt and the minimum dissociation activity of copper.
However, in the presence of hydrocarbons or alcohols, as well as at higher temperatures (>450 °C), the reactivity of nickel with carbon species results in the formation of carbon “whiskers” or nanotubes [43,101,102,103] (see Figure 5). As discussed by researchers from Topsoe [101,102], they form specifically on the external surface of “spearhead-like” and “drill –like” Ni particles with the (111) face largely exposed. These materials do not directly cause the deactivation of the catalyst but increase the density of the bed occupying the empty spaces, thus increasing pressure losses. This chemistry causes difficulties in steam reforming processes (mainly pressure drops). However, this activity is useful for the preparation of nanotube materials, although the metal-catalyzed synthesis of carbon nanotubes produces materials that still contain metal particles, i.e., impure metal nanotube materials, that are not always easy to purify from the catalyst [104,105]. The formation of carbon species from hydrocarbons in the absence of steam, i.e., the decomposition of methane and other hydrocarbons, allows for the coproduction of hydrogen [106] (see below), supposing that the fate of the resulting catalyst-containing carbon material is determined.
Another relevant aspect of Ni catalysts when working in the presence of CO is the possibility of the formation of very volatile carbonyl Ni(CO)4, with the abstraction of Ni from the catalyst. This may thermodynamically occur in the range of 120–200 °C [107]. The formation of nickel tetracarbonyl causes the corrosion of the catalyst and the transport of nickel out of the reactor, with catalytic activity loss and relevant process problems.

2.3.4. Nickel Reactivity with Hydrogen

Bulk nickel hydride can be formed from nickel and hydrogen only at extremely high pressures (>300 MPa) [108,109] with a stoichiometry near NiH. The smaller the nickel particles, the more hydrogen they absorb up to stoichiometries about NiH1.8 above 100 atm [109]. However, at reduced pressures, hydrogen is essentially adsorbed as surface and subsurface atomic species (see below).

2.3.5. Nickel Reactivity with HCl

HCl can be present in gases coming from biomass or waste gasification or pyrolysis, as well as in the product of hydro-de-chlorination processes. The deactivation of nickel-based catalysts by HCl in the presence of hydrogen has been reported [110,111], attributed to the agglomeration of nickel crystallites. It has been concluded that NiCl2 is formed in the presence of HCl and is reduced back to larger-crystal-size metallic nickel [112,113].

3. Surface Properties of Materials Based on Nickel and Alumina

3.1. Characterization of Unreduced Systems

3.1.1. Surface Properties of Aluminas

The surface properties of aluminas have been the object of many investigations and reviews [17,18,114]. Transition aluminas, i.e., essentially γ-, η- and θ-Al2O3, whose typical surface areas are, as mentioned, 150–250 m2/g stable until 400 °C for γ- and η-Al2O3 and 60–120 m2/g stable until 700 °C for θ-Al2O3, have quite similar chemical characteristics, with a strong ionicity and acid basicity of the surface coordinatively unsaturated Al3+ ions—oxide anions couple, dominated by the strong Lewis acidity of low coordination Al3+, most probably on trigonal coordination over the high-temperature activated surfaces [115]. Their surfaces are also usually highly hydroxylated, with a medium–low Brønsted acidity of the surface hydroxy groups.
In contrast, typical and α-Al2O3 powders have typical surface areas of <30 m2/g, are stable up to 1000 °C and exhibit weaker surface reactivity and acido-basicity.

3.1.2. Surface Properties of Nickel Oxide NiOx (x ≥ 1)

Nickel oxide has been the object of few surface characterization studies. The IR spectroscopy of adsorbed molecules [33,116,117] shows the presence of hydroxy groups (νOH 3680, 3600 cm−1) and of weakly acidic Ni2+ sites that are able to adsorb carbon monoxide (νCO 2150–2160 cm−1) and pyridine (ν8a 1595 cm−1) with mode vibrational perturbation. The dissociative adsorption of water on NiO surfaces producing hydroxy groups has been the object of computational studies [118,119]. The surface hydroxy groups have poor Brønsted acidity, being unable to protonate pyridine [33] and ammonia [120]. IR reflection studies on NiO/Ni monocrystals show that NiO adsorbs carbon dioxide at −150 °C in the form of monodentate carbonate species [121]. A more recent calorimetric and IR spectroscopic study shows that the surface reactivity of NiO towards CO2-producing surface carbonate species is not very strong, suggesting a moderate basicity of NiO [122]. On the other hand, using the optical basicity scale discussed by Bordes and Courtine [123], based on the spectroscopic data, NiO is quite as strong of a base as the other rock-salt-type bivalent metal oxides. The surface and the bulk basicity are, according to Busca [124], mainly associated with the size of the cation, which is relatively large, or to its polarizing power, which is relatively small. According to this approach, the NiO surface is expected to be quite basic. The cationic radius of the octahedrally coordinated Ni2+ is reported to be 0.69 Å, which is smaller than that of Mg2+ (0.72 Å), thus suggesting that NiO should be a little less basic than the isostructural MgO. It seems possible to conclude that stoichiometric NiO has a prevalent, moderately basic nature.
As mentioned, in strongly oxidizing conditions, NiO becomes oxygen-rich and non-stoichiometric. XPS studies show that bulk rock-salt NiO particles may contain up to near 40% Ni3+ at their surface if calcined at 400–800 °C [125].

3.1.3. Surface Properties of NiAl2O4 and NiAl2O4-Al2O3 Solid Solutions

The surface properties of stoichiometric bulk NiAl2O4 and NiAl2O4-Al2O3 solid solutions have been the object of investigations by the FTIR studies of the adsorption of probe molecules, such as carbon monoxide and pyridine [34,126]. The surface of stoichiometric NiAl2O4 is largely hydroxylated in normal conditions, whose hydroxy group has low Brønsted acidity similar to those of the transitional aluminas. Strong Lewis acidic Al3+ ions are present at the surface, together with medium Lewis acidity Ni2+ sites and active oxide ions too. Interestingly, the CO stretching frequency of Ni2+ carbonyls observed over Ni aluminates after the adsorption of CO is higher (2180–2185 cm−1) than that observed on pure NiO (2150–2160 cm−1), showing that the Ni2+ centers on these catalysts are more electron-withdrawing (i.e., stronger Lewis acids) than those at the surface of NiO. By increasing Al content on the surface of the NiAl2O4-Al2O3, the solid solutions become more and more similar to those of alumina, with a decreasing amount of Ni2+ ions at the surface, as expected indeed.

3.1.4. Surface Properties of Impregnated NiO/Al2O3 Materials

Several studies report on the surface properties of unreduced NiO-Al2O3 powders produced by the impregnation techniques. At loadings lower than that of the “monolayer coverage” (see above), i.e., with NiO < 25 wt%, the surface probe molecules provide evidence of surface hydroxy groups with poor Brønsted acidity and small amounts of strong Lewis acidic Al3+ (at least two types) as well as weak Lewis acidic Ni2+ ions at the surface [127], whose amount increases with the nickel loading. The surface of such materials appears to be similar to those of the NiAl2O4-Al2O3 solid solutions discussed above. The XPS of alumina-supported nickel oxide also revealed the presence of Ni3+ after calcination at 350 °C, even in the absence of NiO [128]. Experiments with samples with a higher nickel loading show the additional presence of NiO particles, as mentioned above, which also contain Ni3+ ions after calcination at 350 °C [128].

3.2. Characterization of Reduced Ni/Al2O3 Materials

Temperature Programmed Reduction (TPR) in H2-containing feeds is a practical technique for testing the reducibility of catalysts [129,130]. However, it must be taken into account that, while the general behavior is similar in all cases for similar samples, the exact temperature for reduction acts may significantly differ as an effect of the actual conditions of the experiment (feed composition, flowrate and resulting contact time, heating rate, etc.). It must also be taken into consideration that, at the end of the TPR experiments, temperatures such as >1000 °C are usually raised. Thus, in the experiment, the NiO/γ-Al2O3 catalyst usually undergoes alumina sintering, loss of surface area and phase transition, with NiO segregation, agglomeration and reduction. Consequently, at the end of the TPR analysis, the catalyst is strongly modified, and further experiments (such as TPR-TPO cycles) have no sense. The indicative, typical H2-TPR curves are reported in Figure 6.
As already stated, the TPR experiments show that, while the NiO particles are reduced in hydrogen starting at about 300 °C to metallic nickel particles, the reduction of the surface Ni2+ on NiAl2O4 starts at definitely higher temperatures, >600 °C [35,39]. Meanwhile, at about 800 °C, bulk reduction starts to be observed with the formation of metallic Ni and γ-Al2O3 [35]. Surface Ni2+ on Al-rich spinels starts to reduce at higher temperatures than for stoichiometric NiAl2O4, i.e., above 700 °C [39]. The reduction temperature of near-monolayer amounts of nickel oxide impregnated over the different transitional aluminas γ-, δ-, η- and θ-Al2O3 (precalcined at 500 °C) is similar, i.e., above 500 °C [131]. Increasing calcination temperature results in only a slight decrease in the surface area but a significant shift of the TPR peak at higher temperatures [132], which is likely due to the stabilization of the NiAl2O4 subsurface layer. For impregnated samples calcined at high temperature (>700 °C), the TPR peak of nickel reduction is observed above 800 °C with a threshold at about 650–700 °C but shifts slightly to lower temperatures the higher the loading [133,134] and shows an additional medium-temperature reduction peak (550 °C) when the monolayer coverage is approached [49,135]. This reasonably concludes that, for samples pretreated at moderately high temperatures, NiO reduction occurs first (usually 300–400 °C), and highly concentrated dispersed Ni2+ species reduce later (500–600 °C) while highly dispersed Ni2+ reduces even later (>700 °C).
Surface characterization studies performed by the IR spectroscopy of adsorbed CO (Figure 7) allow one to distinguish three types of surface Ni metal sites depending on Ni loading on reduced Ni/γ-Al2O3 [127,136]. For low Ni loading, nearly isolated metal sites that are able to adsorb CO in the form of polycarbonyls are observed, which are produced by the partial reduction of fully isolated Ni2+ surface ions. At medium loadings, together with such polycarbonyl species, on-top monocarbonyl species are observed, with a shifting behavior by increasing/decreasing CO loading typical of CO bonded on metal faces (coupling effect), suggesting that more concentrated Ni2+ ions at the surface of alumina can produce small metallic nanoparticles, which are also observed using Field Emission Scanning Electron Microscopy (FE-SEM) and XRD. When nickel loading is even higher, larger metal particles adsorbing CO as both terminal and bridging species (νCO 2000–1800 cm−1) are observed. Thus, over catalysts with up to near the formal “NiO monolayer” loading, Ni metal particles formed by reduction have a very small particle size, few or few teens of nm at most. When loading is higher, larger and more metal particles are produced by reduction (Figure 8).

4. Mechanistic Aspects in Catalysis by Ni and Alumina Materials

4.1. The Activation of Hydrogen on Ni/Al2O3 Catalysts

The adsorption and reactivity of hydrogen on aluminas have been the object of early studies, see Refs. [137,138,139], showing the formation of different adsorbed species at low temperatures that tend to disappear with increasing temperature [140]. Adsorbed species that are stable at high temperatures are also observed. In particular, Kramer and Andre [139] showed that atomic hydrogen can be spilt over from metallic centers to alumina and remain stable until 480 °C. More recent theoretical and experimental studies suggest that spilt-over hydrogen can be adsorbed over γ-Al2O3-forming hydride species that are likely bonded on trigonal Al3+ centers [141,142,143]. It must, however, be taken into account that these studies fully neglect the possible role of transition metal impurities that are unavoidably present on and in γ-Al2O3, such as, in particular, the non-negligible amount of Fe3+ and Cr3+ [144] that can either be reduced by hydrogen or be involved in H2 dissociation. On the other hand, hydrogen spillover is far more evident on reducible oxides than on alumina [145]. Indeed, quantitatively, the adsorption of hydrogen in non-reducible oxides including γ-Al2O3 is essentially negligible at temperatures up to 600 °C [146].
On the other hand, hydrogen activation over oxide-supported metal catalysts is mostly considered to occur on metallic centers [147], as well as when supports are active in adsorbing it. According to the DFT studies [148], the dissociation of the hydrogen molecule occurs over the topmost Ni atom, and the resulting H atoms may adsorb either on two free hollow sites (but the adjacent bridge sites must be free) or two bridge sites (the adjacent hollow sites must be free). On nickel powder, the TPR studies [149] evidenced three distinct forms of hydrogen permanently chemisorbed at −173 °C: γ-form, located in the subsurface layer and desorbed at about −87 °C, β-form adsorbed in the so-called second layer and evolved at about 54 °C and α-form, fixed directly on the nickel surface and desorbed at the 80–400 °C range. Alumina supports insignificantly affect hydrogen that is strongly adsorbed on nickel but significantly affect weakly adsorbed hydrogen. On the other hand, the nature of the supports also affects the metal particle size and stability of nanoparticles, as well as their electronic properties [150].
In conclusion, hydrogen activation occurs on nickel metal particles, while alumina support influences the size and nature of the supported nickel particles, producing very small nickel nanoparticles when the loading is low. These supported nanoparticles may be more reactive than the larger ones. A possible role of the alumina-uncovered surface for spilling-over hydrogen atoms cannot be excluded.

4.2. The Activation of Carbon Monoxide

CO adsorbs molecularly even at low temperatures on both metals and oxides. The molecular adsorption of CO on aluminas is quite a weak phenomenon, and it is definitely weaker than on zerovalent metal centers [151]. It adsorbs at a low temperature on electron-withdrawing centers on oxide surfaces, in particular, on the Al3+ Lewis acid sites of aluminas through the lone pair at its C atom [152,153], resulting in the strengthening of the CO bond and a shift of the stretching frequency to higher wavenumbers (up to 2220 cm−1 with respect to 2140 cm−1 for unbonded CO). It desorbs easily at room or slightly higher temperatures. At moderately high temperatures, CO adsorption on hydroxylated alumina may give rise to formate ions by reaction with hydroxy groups with an “associative” adsorption mechanism [154].
CO adsorption on metallic surfaces is generally far stronger than on oxides due to the possibility of back-bonding from the partially filled d orbitals of the metal atom to the empty antibonding 2π* orbitals of CO, which produces a quite strong metal–C bond and a weakening of the C-O bond, whose stretching frequency shifts to lower wavenumbers [155]. The adsorption of CO on Ni monocrystal faces has been the object of several studies, showing the possible formation of three forms: (i) terminal (on top), whose CO stretching frequency is in the range of 2120–2000 cm−1; (ii) bridging between two Ni atoms, with CO stretching at 1950–1800 cm−1; and (iii) triply (or multiply) bridging over three or more Ni atoms, with a further significant weakening of the CO bond down to the CO stretching range of 1900–1780 cm−1 [155,156,157,158]. The surface carbonyl species are usually stable upon outgassing at room and slightly higher temperatures, with the stability trend triply bridging > bridging > on top.
As already stated, the formation of a strong Ni-CO bond easily results in the formation of Ni(CO)4, a stable gaseous compound at temperatures below 200 °C, which can result in nickel corrosion and being transported away. At higher temperatures, as a result of this strong interaction, CO dissociation on the nickel surface can also occur, producing carbon and oxygen species [159]. CO dissociation occurs on Ni/η-Al2O3 with an increasing rate at 250–400 °C and with a deposition of increasing amounts of carbon [160], occurring faster on ensembles of several Ni atoms on sufficiently highly loaded samples [161]. The CO dissociation rate is fundamentally influenced by two main factors [162]: (i) an electronic effect on the dissociation barrier, i.e., the increase in the dissociation barrier from left to right in the periodic table (with the trend Fe > Ni > Co >> Cu), which has been correlated with the metal d band center [98,99,100], and (ii) an effect of particle and surface morphology, with lower dissociation barrier on steps, kinks and defects than on terraces. As a common feature in catalysis, activation energy for CO dissociation is correlated inversely with the adsorption strength. The produced carbon species from CO dissociation can finally give rise to encapsulating carbon, which is the main cause of catalyst deactivation [91], or to the formation of bulk Ni3C.
Both carbon species produced by CO dissociation and surface carbonyl species can be the active surface intermediates in carbon oxide hydrogenation reactions, such as methanation and Fischer–Tropsch synthesis. Most of the data strongly converge for a predominant role of CO dissociation over the metal surface and carbide mechanisms for both CO and CO2 hydrogenation to methane over nickel catalyst reactions when carried out with syngas as a reactant [163].

4.3. The Activation of Unsaturated Hydrocarbons

The activation of unsaturated hydrocarbons can occur over oxide surfaces thanks to three different mechanisms: (i) protonation of the C=C bond by Brønsted acidic hydroxy groups; (ii) π-bonding of the C=C bond over d0 and dn cationic sites; (iii) activation at allylic bonds by C-H dissociation over Lewis acid–base pair sites. Previous studies show that, on alumina, the last interaction actually occurs at room temperature over well-activated and dehydroxylated surfaces [164], which likely represents the key interaction involved in the double-bond isomerization activity of alumina. However, it has also been shown that the protonic acidity of alumina’s surface OHs is not negligible, and the protonation of butenes can occur at high temperatures (>400 °C) as the main step in the olefin skeletal isomerization activity of alumina [164].
As in the case of CO activation, the adsorption and activation of unsaturated hydrocarbons (acetylenics, dienes, olefins and aromatics) on metal centers also primarily occurs thanks to the electron donation from the full π-orbitals of the organic molecule to the empty d-type orbitals of the transition metal and the simultaneous electron back-donation from the full d-type orbitals of the transition metal to the empty π*-antibonding orbitals of the unsaturated organic molecule. According to the DFT calculations and STM experiments, this so-called π-bonding is the stable form of ethylene adsorbed on the flat and stepped Ni(111) crystal faces, [165] as well as on the nickel clusters [166]. However, according to calculations, this primary interaction could evolve towards more strongly adsorbed species producing multiple σ-bonding. As an example, the most stable acetylene adsorbed form on nickel should be the so-called µ-bridge, where the acetylene molecule interacts with four carbon atoms and retains a single C-C bond [165,167]. Similarly, benzene and aromatics adsorb on metallic nickel first through π-bonding, finally producing multiple σ-bonds with more nickel metal atoms [168]. At moderately high temperatures, e.g., >400 °C, all hydrocarbons will decompose on nickel surfaces producing carbon species [169].
In the presence of hydrogen, these surface hydrocarbon intermediates should later undergo the attack of atomic hydrogen species adsorbed on the nearest sites, finally giving raise to the hydrogenation reactions.
It is, however, interesting to remark that Ni2+ ion, being a d8 cation, is able to produce π-bonding, e.g., with acetylene and ethylene. This is demonstrated by the activity of NiCl2/Al2O3 powders as selective adsorbents of acetylene [170] and by the high catalytic activity of Ni2+ supported on acidic supports, such as silica-rich amorphous alumina [171] for the ethylene dimerization process to butene where the interaction of Ni2+ with the olefin π bonds is a key step [172].

4.4. The Activation of CO2

The activation of CO2 is relevant with respect to two types of catalysts based on Ni/Al2O3:CO2 and COx methanation and methane dry reforming. The two reactions have different thermodynamic properties (the first being exothermic, the second endothermic) and are realized in very different conditions: 300–400 °C and 20–30 atm in H2 for methanations and >600 °C at an atmospheric pressure for dry reforming.
CO2 adsorption occurs quite weakly at room temperature on hydroxylated alumina, mainly in the form of bicarbonates and of linear-coordinated CO2, while easy desorption is observed at moderate temperatures. However, it can become stronger in the form of carbonate species for high-temperature dehydroxylated alumina [114] or when alkali and alkali earth ions [173] or rare earth ions [174] are present on doped aluminas or on metal aluminates [175]. On the other hand, as seen above, carbon dioxide may adsorb in the form of carbonate species over nickel oxide [121,122], which is expected to show a higher basicity than alumina. Indeed TPR data indicate that CO2 adsorption on NiO/Al2O3 is weak, with a desorption temperature at about 100 °C [176], which is in agreement with the lower-surface basicity of NiO/Al2O3, at least in terms of the number of basic sites with respect to alumina [52].
CO2 adsorption on metal surfaces has been the object of early spectroscopic studies on monocrystal surfaces [177,178,179]. These studies report strong spectroscopic evidence for the formation, in addition to the physisorbed CO2 that only forms at very low temperatures, of negatively charged bent CO2δ−, which, depending on the nature of the metal, may dissociate into CO and O or transform into CO32− + CO. Theoretical studies of CO2 adsorption on transition metals [180,181,182] indicate that the strength of the CO2 exothermic adsorption forming bent CO2δ−, the charge of this species, and the exothermic dissociation energy to produce CO-O, all follow the trend Cu < Ni < Co < Fe, i.e., the trend of the d-band position. A recent ambient-pressure X-ray Photoelectron Spectroscopy study [183] indicated that carbonate species, adsorbed CO and graphitic carbon form on both Ni(111) and Ni(100) surfaces in the presence of 0.2 Torr CO2. However, more than 90% of the adsorption species on the Ni(111) surface is the carbonate, whereas the Ni(100) surface is mainly covered by adsorbed CO and graphitic carbon [183].
The adsorption of CO2 has been studied over reduced Ni/Al2O3 by DRIFT spectroscopy. In addition to the bicarbonate species bonded on the alumina support, carbonate and carboxylate species have been observed, that are supposed to be coordinated at the Ni–alumina interface and over metallic nickel, respectively, and the TPR data indicate that these species are adsorbed stronger than bicarbonates on bare alumina [184]. Interestingly, no CO coordinated to the Ni, which might be originated from the CO2 dissociation. Consequently, it seems that the activation of CO2 at low/moderate temperatures for its further hydrogenation on Ni/Al2O3 would occur mostly on Ni metal centers or at the support/metal interface, although, in the case of the presence of basic dopants, the support might have a role too [185]. On the other hand, data indicate that metallic nickel may be oxidized by CO2 above 700 °C, producing mainly CO and partially oxidized nickel [186]. This reactivity might be involved in the dry reforming of the hydrocarbons that only occur near that temperature.

4.5. The Activation of Water

The adsorption of water on alumina results in the formation of hydroxy groups, by dissociation. A part of such hydroxides appears to have moderate acidity, being unable to protonate pyridine, but being able to protonate stronger bases such as amines and piperidine [17,18]. On the other hand, a part of the hydroxy groups shows moderate basicity/nucleophilicity, being able to produce hydrogencarbonate species reacting with CO2 at low temperatures [114] and formate species reacting with CO at moderately high temperatures [154]. Thus, water adsorption on alumina can result in its activation as a base or as an acid. The un-reducibility of alumina does not allow for activation in terms of redox chemistry. A similar acid–base activation of water is observed on NiO and on NiO/Al2O3 powders, where surface hydroxy groups are observed too (see above) and are thought to have a predominantly basic character.
Water adsorbs also on nickel metal at very low temperatures. No nickel oxidation was found by reaction with steam at 25 °C [88]. However, XPS data on Ni polycrystalline surfaces show that, at 300 °C, water vapor produces a thin NiOx surface layer (<2 nm thickness) that also contains Ni3+ ions, similar to that produced by oxygen [88]. Thus, an oxidizing layer can be produced by the reaction of metallic nickel with water at relatively low temperatures. Experiments of the steam reforming of ethanol, a molecule whose activation is easy at a low temperature, over unsupported nickel allowed us to conclude that water activation is needed for such reactions and that the temperature threshold for water activation on unsupported nickel is less than 400 °C [72].

4.6. The Activation of Oxygen

Alumina as such is essentially unable to activate oxygen [140], except maybe in the case of impurity centers [144]. As reported above, calcined NiO, as well as NiO-Al2O3-based materials, do present XPS features attributed to Ni3+ (see above). In the case of NiO, this produces p-type semiconducting NiO1+x materials where the excess oxide ion species are supposed to be on reticular positions and are balanced by nickel vacancies [187], which are likely mainly occurring near the surface. At the surface of NiO/Al2O3, a similar mechanism should occur, with the formation of additional surface oxide ions and the oxidation of Ni2+ to Ni3+. This is expected to be a mechanism of oxygen activation upon oxidation catalysis over nickel-oxide-based materials.
On the other hand, data show that polycrystalline Ni already undergoes surface oxidation by O2 at 25 °C, producing a NiO1+x layer containing both Ni2+ and Ni3+ [87]. Thus, metallic nickel is also able to activate oxygen at low temperatures. However, obviously nickel metal would no longer exist in an oxygen-containing atmosphere at moderately high temperatures, as discussed above.

4.7. The Activation of Saturated Hydrocarbons on Reduced and Oxidized Nickel

The activation of methane over heterogeneous catalyst surfaces has been the object of several studies [188]. Among the simple transition metal oxides, NiO has the highest catalytic activity for methane combustion, with an ignition temperature in the range of 450–500 °C. This is attributed by a reaction with surface oxide ions interacting with Ni3+ ions [189] and, thus, with an Eley−Rideal mechanism. Thus, oxygen activation is needed more than methane activation for its combustion. Similar activation mechanisms can be proposed for the catalytic combustion of higher paraffins, whose reactivity is expected to be higher than for methane.
In the absence of water and oxygen, the formation of carbon species from several light hydrocarbons on nickel foils occurs over the temperature range of 400–600 °C [169]. Carbon formation from paraffins is comparatively slower than from olefins and acetylene. In fact, methane activation on metal surfaces, including nickel, has been reported to start with a single C–H bond breaking as the molecule collides with the surface, leaving chemisorbed H and CH3 fragments [190], which occur at about 100 °C [191]. This step is quite a slow one, is very temperature-sensitive and is supposed to be followed by the sequenced dehydrogenation of carbon species. On the other hand, the production of hydrogen by methane decomposition on nickel catalysts occurs at relatively high temperatures, i.e., in the 400–600 °C range [192,193].
When higher paraffins are reacted, a similar dissociative activation step can be presumed to occur on nickel metal particles as a first step for dehydrogenation reactions. On the other hand, nickel catalysts do mostly show low selectivity in paraffin dehydrogenation due to the too high activity in C-C bond cracking finally producing decomposition and carbon deposition. The selectivity in isobutane dehydrogenation on Ni/Al2O3 is, however, increased by partial sulfidation [194] that likely poisons the too-active sites. As is reported below, however, nickel catalysts—in particular, Ni/Al2O3—are reported to act as excellent catalysts for cycloparaffin aromatization. Early studies suggest that carbonized Ni metal surfaces are active in cyclohexane aromatization [195]. On the other hand, authors suggest that unreduced dispersed Ni2+ on alumina may act as the active site for propane dehydrogenation on Ni/Al2O3 [196]. Thus, the real active sites for paraffin dehydrogenation on Ni-Al2O3 catalysts are still the object of debate and can be different in different conditions.
In conditions where both paraffins (most commonly methane) with water vapor (steam reforming), or with CO2 (dry reforming), or with oxygen (partial oxidation), or all together (tri-reforming), are present together, it is not evident whether alkane activation is needed and whether it occurs on nickel metal surface or over the oxide layer produced by water, CO2 or oxygen activation. In fact, the temperature ranges of methane, water and CO2 activation nearly superpose. Most authors believe that methane activation occurs on the nickel surface that is reacting further with adsorbed oxygen species in all cases [188], and the usual deactivation of such catalysts by carbonaceous materials or, at least, the evident formation of carbonaceous species agrees with this conclusion. Among the supported nickel catalysts, Ni/Al2O3 is reported to have the lowest onset methane conversion temperature [197], thus making alumina the best support for producing active nickel species.

5. Deactivation and Reactivation of Ni/Al2O3 Metal Catalysts

Different types of deactivation mechanisms may affect catalyst behavior [91]. Some of them are due to solid-state reactions slowly occurring at moderately high temperature, such as the loss of support surface area (intraparticle sintering), support phase transition (γ- to θ- to α-Al2O3) and the formation of bulk NiAl2O4. The sintering of supported metal particles usually occurs in parallel. These phenomena are essentially irreversible and do not allow for catalyst regeneration. In correct industrial conditions, they usually occur very slowly over years. When deactivation is excessive, the catalyst bed must be substituted by a fresh one.
Ni metal particle sintering can also occur without significant support sintering [198]. This phenomenon is also a thermal one but may also be favored by the adsorption of molecules, such as HCl (see above). In this case, at least partial regeneration can be attempted either by steaming or by the oxidation/reduction of the catalysts, allowing for nickel redispersion [199,200].
As already cited, Ni metal catalysts can react with species in the reactant medium and become deactivated by poisoning. To reduce this drawback, feed pretreatment can sometimes be realized: as an example, natural gas is virtually fully desulfurized before steam-reforming reactors to avoid nickel sulfidation. Thus, in most cases, the trace of contaminants still unavoidably present can cause deactivation very slowly (years on steam), and catalyst regeneration is consequently not needed.
When, however, the presence of poisons in the feed cannot be avoided, deactivation can be faster, and catalyst regeneration must be carried out. The surface or bulk sulfidation of nickel metal particles results in partial or total deactivation in most cases [198]. However, the adsorption and desorption of H2S are reversible phenomena, and this may allow for the equilibration of adsorbed and gas–phase species, resulting in limited deactivation. On the other hand, treatment in sulfur-free feeds can allow for surface cleaning by H2S desorption with the recovery of catalytic activity. Alternatively, oxidation may allow for the removal of sulfur as SO2 [201]. Steaming can also allow for the removal of sulfur [202,203], although, in these cases, residual sulfate species may remain on the surface. In oxidizing conditions, sulfur can be present in the feed, mainly as SO2, and this can also cause catalyst poisoning by sulfites and sulfates [204]. Oxidation and reduction treatments can remove sulfur poisons also in this case.
As said above, in the presence of carbon reactants (carbon oxides, hydrocarbons, alcohols, etc.), the formation of surface carbon deposits (surface carbide species, encapsulating carbon and carbon whiskers) also can occur at high temperature [205]. While carbon whiskers do not cause evident catalyst deactivation, at least initially, they causereactor overpressures with time [101,102]. Surface carbides and encapsulating carbon, instead, usually reduces catalytic activity down to total deactivation [95,198]. In this case, heating in air or oxygen results in the burning of carbon residues. Alternatively, steaming results in the gasification of surface carbon species. Finally, hydrogenation can remove carbon deposits also [91]. These procedures, followed by re-reduction treatments, can allow for the at-least-partial regeneration of the catalysts.
As is quite obvious, regeneration pretreatments must be carefully realized, avoiding excessive heating phenomena to not cause alumina sintering and phase transitions during regeneration.

6. Applications of Catalysts Based on Nickel and Alumina or Aluminates

6.1. Commercial Ni-Al2O3 Catalysts for Hydrogenation Reactions

In most hydrogenation reactions, noble metals such as Pt, Pd and Ru are far more active at the same loadings than nickel catalysts. However, thanks to its much lower cost, nearly 1000 times less than ruthenium and more than 2000 times less than platinum and palladium, nickel loading in commercial catalysts can be far higher. This allows for higher catalytic activity and the better management of sulfur poisoning. In fact, the same sulfur content in the reactant flow is more detrimental for <1 wt%-loaded noble metal catalysts than for 20–50 wt%-loaded supported Ni catalysts due to the much higher metal surface area. For these reasons, highly loaded Ni on alumina may definitely be competitive with extremely low-loading, supported, noble metal catalysts in several practical cases and is still much cheaper.
In this chapter, we will not take into consideration the typical hydrodesulfurization and hydrotreating catalysts that are usually based on alumina-supported MoS2 and WS2 where nickel may act as an activator (the so-called NiMo and NiW systems). This is because the main role in catalysis in this instance is due to the sulfides of molybdenum and tungsten.

6.1.1. Hydrogenation of Acetylenics and Dienes

Ni/Al2O3-based catalysts may be used instead of the more active Pd or Pt catalysts for hydrogenations of unsaturated hydrocarbons, such as to reduce acetylenics and aromatics in hydrocarbon mixtures coming from steam cracking or fluid catalytic cracking. In fact, Ni/Al2O3 [206] and Ni/NiAl2O4 [46] catalysts have interesting activities in acetylene-selective hydrogenation to ethylene. Ni/Al2O3, possibly containing copper, has excellent activity in the hydrogenation of acetylenics in C2, C3 and C4 cuts in the presence of butadiene [207] and can be preferential with respect to noble metal catalysts if the feed contains significant amounts of sulfur compounds. Clariant (previously Süd Chemie) offers multi-promoted NiO catalysts supported on silica alumina (0.6–2.6% NiO), denoted as the OleMax 100 series [208] for the treatment of FCC offgases. The hydrogenation of α-acetylenic compounds (1-butyne and vinylacetylene) may be performed with 100% yield to purify the steam cracking C4 cut in order to simplify the further extraction of 1,3-butadiene (e.g., by N-methyl pyrrolidone). This occurs [209] in the UOP-KLP process, using a single fixed bed of the KLP-60TM catalyst, Cu-Ni/Al2O3, in the liquid phase, as well as with the Hüls Selective Hydrogenation Process (SHP) using the Ni-based H-15 SHP catalyst from the UOP. Catalysts of 1–25 wt%Ni on θ- or γ-alumina have also been patented for the selective hydrogenation of phenylacetylene impurities in styrene/ethylbenzene mixtures [210] where they were found to resist deactivation better than the Pd-based catalysts.
Small amounts of straight-chain C10 to C13 di-olefins are selectively hydrogenated to the mono-olefins in the feed to the alkylation reactor in the process of the production of Linear Alkyl Benzene (LAB). The reaction is carried out in a liquid phase at 170–230 °C and 10–20 bar on nickel–alumina (Ni/Al2O3) catalysts [211,212].
Pyrolysis gasoline is purified from acetylenics and dienes by hydrogenation before the hydrodesulfurization occurs. Johnson Matthey offers the HTC NI 200 (12 wt% Ni on alumina [213]), which is available in three forms—oxidic (OX), pre-reduced and air passivated (RP) and pre-reduced, air passivated and pre-sulfide (RPS) [214]—and has stronger heavy metal and sulfur resistance with respect to Pd catalysts. Also, the Topsoe SH-501 catalyst [215], as well as the different Axen catalysts [216] offered for the first step of pyrolysis gasoline hydrogenation, are based on the Ni/alumina systems.

6.1.2. Dearomatization Reactions

Supported nickel catalysts may be used for the dearomatization of hydrocarbon flows in competition with Pt catalysts. Due to the wide pore structure and small nickel crystals providing high strength, activity, selectivity and poison resistance, together with lower gum/polymer formation and side reactions, nickel catalysts can offer comparable performance to noble metal catalysts. Clariant (previously Süd Chemie) offers Ni-based catalysts (the NiSAT® family, with NiO content from 43% to 77%) for naphtha dearomatization. The commercial nickel catalyst NiSAT 200 is reported to contain between 25 wt% and 50 wt% nickel monoxide, below 25 wt% calcium aluminate and more than 20 wt% silica [217]. The Johnson Matthey HTC Ni dearomatization [214,218] and solvent purification catalysts [219], from HTC Ni 200 to HTC Ni 900, are based on Ni/Al2O3 with Ni from 15 to 30 wt%; in particular, HTC Ni 500 is 21% Ni on alumina [220]. Also, the Axens LD 746 and AX 746 dearomatization catalysts are based on Ni/Al2O3 [216].

6.1.3. Low-Temperature Methanation of COx for Hydrogen Purification

Low-temperature methanation is usually realized in processes for preparing hydrogen for hydrogenation reactions or H2/N2 mixtures for ammonia synthesis [221,222]. They are realized to reduce as much as possible the concentration of COx in such gases because of the poisoning effect of COx towards hydrogenation catalysts. This reaction is realized in highly diluted COx in H2 mixtures (COx concentrations most commonly < 1% mol), with CO being the predominant COx compound in the feed, allowing the COx concentration to be reduced to less than 5ppm in the process gas leaving the methanator. As a consequence of the small number of reacting molecules, the evolved heat is low, and adiabatic reactors can be used. The reaction temperature is typically 170–350 °C. Ruthenium-based catalysts are most active at low temperatures. However, they have found only a limited area of application due to their high cost. Due to their sufficient activity, stability and relatively low cost, Ni-based catalysts are predominantly used for conventional low-temperature methanation [221,223,224]. Typical nickel loadings are in the range 20–45 wt% as NiO, over γ-Al2O3 or metal aluminate supports, or Cr2O3 [223]. Clariant produces catalysts with 43% and 25% NiO on a support (likely alumina) METH 135 and METH 134 in the oxidized form [59]. Topsoe produces the prereduced PK-7R catalyst, constituted by >23 wt% Ni [225] on an alumina carrier, which ensures that CO and CO2 are fully converted to methane at an operating temperature of 190 °C. Johnson Matthey sells catalyst precursors (Katalko 11 series) constituted by 35 wt% NiO supported by refractory oxides, such as alumina–silica–lime–magnesia, strengthened with calcium aluminate cement [226].

6.1.4. High-Temperature Methanation of COX for the Production of Substitute Natural Gas

High-temperature methanation is the conversion of a CO+CO2-rich syngas that comes from coal gasification or biomass conversion to produce a methane-rich gas to be used as a Substitute Natural Gas (SNG [227,228]). Due to the much higher COx concentration in hydrogen (usually in the 25–35% mol range) and their high conversion degree, the heat evolved is very high, and the temperature, at least in the first reactor, is also definitely high. Although real industrial applications of this technology have been very few in the previous century, several plants appear to having been recently put in operation or are under construction now—in particular, in China [229]. Only Ni-based catalysts are apparently used, taking into account that the extent of the reaction under industrial conditions is more heat and mass transfer-limited, rather than kinetic-limited. With these catalysts, selectivity to methane in the full process is very high, although the carbon deposition and formation of higher hydrocarbons occur to a limited extent. In the Topsoe’s TREMP™ process, the catalyst (MCR2X) [230] is a supported nickel catalyst containing 22 wt% Ni on a stabilized support, possibly based on MgAl2O4 [231], with a surface area decreasing from 50 m2/g (fresh) to 30 m2/g (used) [232] and with stable activity up to 700 °C. Other nickel-based catalysts, MCR-8 and PK-7R, are used in medium and low temperature steps. Catalysts for this process need to be sufficiently active at low temperatures, resistant against sintering at high temperatures, and to other phenomena producing deactivation, i.e., gum and Ni(CO)4 formation [233]. The Johnson Matthey CRG catalysts used in the process are apparently based on Ni and refractory supports such as alumina–lanthana–magnesia [234]. They catalyze the methanation reaction and, simultaneously, the Water–Gas Shift (WGS) reaction, thanks to water produced by the methanation reaction. Also, the catalysts provided by Clariant Catalysts (Munchen, Germany) [235], with a wide operating temperature range (230–700 °C), are based on nickel [236].

6.1.5. Methanation of CO2

The methanation of the captured CO2 using green hydrogen is considered a possible way to valorize and reuse it, in the frame of the so-called “power to gas” technologies [237,238]. Many projects and plants with various sizes have been undertaken in the last fifteen years [239,240]. The new processes currently under development for CO2 methanation [241,242] are strictly related to the ones described above for SNG production from CO-rich syngases. Although an enormous number of projects are under development, information on real industrial applications is scarce. One of the earliest examples of power-to-gas technology was Audi’s e-gas plant in Werlte, Germany, which was constructed by Etogas to produce renewable synthetic methane from hydrogen and carbon dioxide emissions and to achieve low-carbon mobility for Audi’s A3 Sportback g-tron vehicles. Clariant provides the catalyst to be applied in such an application [243]. No detailed information is available on the Clariant’s SNG 100 DCARB methanation catalyst just developed for power-to-gas processes though it seems that it is based on nickel [236]. In fact, taking into account that practical reaction temperatures in the catalyst bed are relatively high, the low-temperature activity of noble metals such as ruthenium is not necessary, at least in main methanation reactors. Thus Ni-based catalysts are most promising [244]. High-temperature stability is certainly a key feature. A very large number of different catalyst compositions have been proposed in the literature for CO2 methanation. A critical analysis of the scientific literature data concerning catalysts for CO2 methanation allowed us to propose that Ni/Al2O3 catalysts with medium metal loading that allow for small metal particles to cover almost entirely the support surface, with additional ceria and/or basic oxide dopants as stabilizer and activating components, would be very active toward methanation and not as much prone to produce carbon residues and gums [185]. Takovite-derived coprecipitated Ni/Al2O3 catalysts can also applied to the 3D printing technique to fabricate commercial spherical CO2 methanation catalysts [245].

6.1.6. Reverse Water–Gas Shift

The reverse water–gas shift reaction (rWGS) may represent an interesting approach to produce CO, or CO-containing syngases, from captured CO2 as a first step for further conversion [246], e.g., to synthetic and renewable fuels from the Fischer–Tropsch process, acetic acid, oxo-alcohols or methanol syntheses and methanation. Being rWGS, an endothermic reaction, which is shifted to products only above around 830 °C, when its ΔG° starts to become negative, relatively high temperatures are needed for high conversion. Typical water–gas shift catalysts based on Cu/ZnO/Al2O3, Fe3O4 or ZnAl2O4, which are active also for rWGS at low temperatures, cannot be applied at high temperatures [185]. Also, Ni/γ-Al2O3 catalysts at low nickel loading were found to be active and selective in rWGS instead of the competitive methanation reaction at low temperatures [50]. In fact, several companies such as Topsoe (Lyngby, Denmark) [247], Clariant Catalysts (Munich, Germany) [248], Johnson Matthey (Billingham, UK) [249] and Axens (Salindres, France) [250] are developing catalysts for rWGS based on nickel on refractory supports, which will likely include corundum and alkali-earth aluminates. It is, however, supposed that the real reaction occurring is a sequence of methanation and methane steam reforming instead of direct rWGS over these catalysts.

6.2. Ni/Al2O3-Based Catalysts for Other Hydrogenation Reactions

6.2.1. Application in Amine Syntheses

In addition to the above applications, which have industrial relevance in refinery and petrochemistry, Ni/Al2O3-based catalysts may have applications in the hydrogenation field. In particular, Ni/Al2O3 systems are cited among the industrial catalysts for the synthesis of amines through the gas-phase catalytic hydrogenation of nitriles and reductive amination reactions [131,251]. In particular, Ni/γ-Al2O3, as such or also containing MgO, is reported to act as a good catalyst for the hydrogenation of acetonitrile to ethylamine [252] and for the reductive amination of adehydes (including purely aliphatic ones), aryl–alkyl, dialkyl and diaryl ketones to primary amines [253]. The Johnson Matthey HT 500 catalyst, based on 21 wt% Ni/Al2O3, is an active catalyst for the synthesis of aniline by nitrobenzene hydrogenation [254].

6.2.2. Hydrogenation of Oxygenates

Several studies report on the good catalytic activity of nickel-based catalysts in the hydrogenation of oxygenated compounds in fine industrial chemistry [255]. Recent studies report on the very high activity and selectivity of Ni/Al2O3 catalysts in the hydrogenation of n-octanal to n-octanol [256] and for the liquid-phase hydrogenation of hydroxymethyl furfural to tetrahydrofuran-2,5-diyldimethanol (THFDM) in water [257]. Similar catalysts also exhibited high catalytic activity in the maleic anhydride hydrogenation with 92% selectivity to succinic acid and nearly 100% conversion at 140 °C [258]. The industrial HT-500 catalyst from Johnson Matthey is active in the hydrogenation of a number of oxygenated compounds, but its stability is strongly improved in the presence of acid components by the addition of 0.1–0.25 wt% ruthenium [220].
A number of chemicals must be purified of impurities by the hydrogenation of unsaturated impurities. 2-ethylhexanol is an important industrial material, in particular, for the production of its phthalate esters as lubricants and plasticizers and is usually obtained by the hydrogenation of 2-ethylhexenal. However, the quality of the product is greatly affected by the existence of unsaturated by-products (2-ethylhexanal and 2-ethylhexenol) and residual reactant (2-ethylhexenal). Ni/Al2O3 catalysts can be applied to purify crude ethylhexanol by the hydrogenation of such impurity molecules [259].

6.2.3. Hydrogenations and Hydrodeoxygenation of Renewables

As discussed elsewhere [2], the production of biofuels is expected to grow in the near future to contribute to the defossilization of transport technologies, in particular, in the field of shipping and aviation. To produce biofuels, the hydrogenation, hydrodeoxygenation (HDO) and/or deoxygenation of biomass-derived organics is needed. Among the promising processes and products, hydrogenated vegetable oils (HVOs) are an already-growing market, such as green diesel and sustainable aviation fuel (green jet fuel). The theoretical reaction stoichiometry for tri-stearate full hydrogenation is the following:
(C17H35COO)3C3H5 + 12H2 → 3C18H38 + C3H8 + 6H2O
where ΔH°298 = −888 kJ (liquid water) [260], showing the strong exothermicity of the process. The full hydrogenation of vegetable oils, mainly non-edible ones such as jatropha oil, castor oil and spent frying oil waste, is mainly realized today using alumina-supported Ni-Mo sulfides, which need sulfur in the feed to maintain stability. The process produces a mixture of hydrocarbons, mainly linear paraffins, and also some CO2 by decarboxylation. Nickel/alumina catalysts with high nickel loading (60 wt%) were found to have stable activity for this reaction at H2 pressure, 40 bar and temperature 350 °C [58,261]. Interestingly, Ni/MgO-Al2O3 materials (up to 20 wt% Ni) were also found to be active for this reaction [262] but also in the H2-free deoxygenation of vegetable oils, producing olefin-rich, diesel-like hydrocarbon mixtures by catalytic cracking at 450–500 °C [263]. The 10 wt% Ni/γ-Al2O3 catalyst exhibits good performance in the HDO of methyl laurate at 400 °C and 20 bar, mainly producing n-undecane [264].
Both 10% Ni/γ-Al2O3 and, even more, 10% Ni/α-Al2O3 were found to exhibit high activity in the hydrogenolysis of molecules similar to lignin monomers, such as diphenyl ether (DPE) and benzyl phenyl ether (BPE), but also show some activity in full lignin hydrolytic depolymerization [265]. In fact, an effective hydrolytic depolymerization of lignocellulosics was found using EDTA-activated Ni/Al2O3 [266] or by using Ni/Al2O3 in methanol [267].

6.2.4. Hydrogenolysis and Pyrolysis for the Conversion of Plastic Wastes to Useful Products

Another important point for the future is to convert wastes into chemicals, in particular, to recover organics from plastics [2]. Ni/Al2O3 catalysts were found active for the hydrotreating of plastic pyrolysis oils [111,268]. Similar catalysts are also reported to be very active for the pyrolysis and the production of hydrogen [269,270] from plastic wastes. Indeed, the pyrolysis of plastics will induce the dehydrogenation and hydrogenolysis of polyolefins together.

6.3. Catalysts for Hydrogen and Syngas Production

6.3.1. Industrial Steam Reforming of Methane, Natural Gas and Higher Hydrocarbons

Hydrogen is mostly produced today through the steam reforming of hydrocarbons, usually natural gas or methane [221,271,272,273], performed at 750–900 °C, 30–50 bar. Although Ni/γ-Al2O3 catalytic materials are even more active in short time frames in steam laboratory experiments [274], the industrial catalysts are mostly based on Ni supported on aluminate-based carriers, usually stabilized by the presence of alkali and/or alkali-earth cations to obtain the refractory character needed for stability at such a high temperature [275]. Typical methane steam-reforming catalysts contain 10–25 wt% Ni, 70–85 wt% Al2O3 or aluminate supports and up to 5 wt% K, Ba, Ca, etc., and have a limited surface area of 10–30 m2/g. The support phase is in fact a refractory oxide such as α-Al2O3, Mg aluminate spinel MgAl2O4, calcium aluminates such as CaAl12O19 and calcium-potassium aluminate CaK2Al22O34 [59]. The formation of carbon nanotubes is a main difficulty in particular when natural gases containing high amounts of higher hydrocarbons are fed. In this case, over-pressions can appear, as well as catalyst deactivation results. To limit this phenomenon, cesium- or potassium- doped catalyst may be put at the beginning of the reactor. As an example, the fresh Johnson Matthey catalyst Katalco 57-4Q catalyst (29 m2/g) contains 18 wt% NiO supported over a mixture of calcium aluminate phases, while its alkalized analogue Katalco 25-4Q additionally contains 1.8 wt% of K2O [276]. Similarly, the Clariant ReforMax 330 LDP Plus catalyst for methane SR is reported to contain 14 wt% NiO over CaAl12O9, while the alkalized ReforMax 210 LDP Plus catalyst contains 18 wt% NiO and 1.5 wt% K2O over CaK2Al22O34 and can also be used for the steam reforming of liquified petroleum gas, i.e., propane and butane. For naphtha steamreforming, Clariant commercializes the ReforMax 250 catalyst, with 23 wt% NiO and 6 wt% K2O over Mg aluminate [59].

6.3.2. Industrial Pre-Reforming of Natural Gas

Modern large hydrogen production plants contain a previous adiabatic fixed bed pre-reforming reactor, where higher molecular weight hydrocarbons are pre-reformed. The steam reforming of propane and butane occurs at lower temperatures (up to 650 °C), and this allows one to avoid their decomposition to carbon and hydrogen in the main tubular reactor [271]. These catalysts contain much more Ni (NiO 45–60%), have very high nickel surface areas (typically 20–30 m2/gcat, compared to <2 m2/gcat for a primary reformer catalyst) and usually a Mg aluminate support with other additional components (some silica and chromia). The Clariant Reformax 100 pre-reforming catalyst contains 56 wt% NiO [59].

6.3.3. Methane Dry Reforming and Biogas Reforming Reaction

The so-called “dry reforming” reaction, i.e., the reaction between methane and carbon dioxide to produce syngas and hydrogen, represents a possible future new way to produce hydrogen, consuming two greenhouse gases. In particular, this reaction can be interesting to produce renewable hydrogen when applied to biogas [277]. While Ni/γ-Al2O3 catalysts are excellent for this reaction, they tend to deactivate by carbon deposition. Additionally, the relatively high reaction temperature needed (>700 °C) may cause catalyst instability for long times on steam. Both drawbacks can be approached and improved by adding a basic component to the catalyst, such as, e.g., MgO or CaO, and adding small amounts of noble metals (e.g., Pt) [278,279,280,281].

6.3.4. Methane Partial Oxidation, Secondary Reforming and Autothermal Reforming

The partial oxidation of methane, i.e., the oxidation of methane in defect oxygen or defect air, represents another way for producing syngas and hydrogen. It is an exothermic reaction and, for this reason, can be interesting with respect to steam and dry reforming. In practice, methane partial oxidation is commercially realized today, together with steam reforming in secondary reformers, with air and a methane content of 15 mol%, realized in the processes for producing hydrogen gas for the ammonia synthesis. In this case, the gas entering the catalyst bed, after a first thermal combustion and reaction step, is at 1050–1100 °C, where further endothermic reforming reactions take place to achieve an equilibrated gas of 925–975 °C at the exit of the reactor [272]. Additionally, this reaction is realized in autothermal reformers, e.g., the processes for producing syngases for methanol and synthesis where the steam reforming and partial oxidation of methane occur together. In this case, the temperature of the combusted gases entering the catalyst bed, after the non-catalytic step, is 1250–1300 °C.
Ni/γ-Al2O3 catalysts are active for this reaction. However, due to the high temperature during the reaction, the Ni/θ-Al2O3 catalysts and Ni/α-Al2O3 catalysts may have more stability on steam [282,283]. In fact, commercial catalysts for secondary reforming and autothermal reforming from Clariant are based on 11.5 wt% NiO on CaAl12O19 (ReforMax 410 LDP catalyst) but a top layer 11.5 wt% NiO on α-Al2O3 (ReforMax 400 GG) is recommended. For autothermal reforming, Clariant proposes approx. 5–10% of ReforMax 420 (7.4 wt% NiO on α-Al2O3) on top of ReforMax 330 LDP catalyst (14 wt% NiO on CaAl12O19) [59].

6.3.5. Methane and Hydrocarbon Decomposition Catalysts for the Production of Hydrogen and Carbon Nanomaterials

As already stated, Ni-based catalysts may react with hydrocarbons (usually methane) producing solid carbon species and C-free gaseous hydrogen. This is the so-called turquoise hydrogen [106,284]. Nickel-based materials are among the most active catalysts for this reaction. This reaction occurs at 800–900 °C over the 10% Ni/Al2O3 catalyst stabilized by calcium [285] with the formation of carbon nanotube-like deposit over the catalyst. The encapsulation of metallic nickel particles finally stops the reaction due to catalyst deactivation. In the case of the Ni/MgO-Al2O3 catalysts, the catalyst loaded with 25 wt% Ni has more prolonged activity than the 15 and 50 wt% Ni-loaded catalysts [286].
This activity can allow for the coproduction of hydrogen and useful carbonaceous materials, although the separation of carbon species from the catalyst is a problem. In practice, the production of carbon nanotubes [104,105] occurs over Ni/Al2O3 in the 450–700 °C temperature range. The amount of carbon nanotubes (CNTs) formed before the deactivation of the catalyst increases with the nickel content in the catalyst [43,287]. On the other hand, continuous nanofiber layers can be produced using very-low-loaded (0.04 wt%) Ni/α-Al2O3 catalysts [288]. The addition of other catalyst components, such as an alloying metal or an oxide component, can improve the activity and modify the morphology of the produced carbon material [284].

6.3.6. Ammonia Cracking Catalysts

Ammonia decomposition is commonly applied today on a small scale to produce forming gas that is used in metallurgical processes. In the near future, ammonia may be used as a hydrogen carrier molecule [289], thus large-scale ammonia cracking could be developed. While noble metal-based catalysts are needed for low-temperature ammonia cracking, nickel/alumina or aluminate-based catalysts [290] are commercial for high temperature ammonia cracking at temperatures up to 900 °C [291]. Johnson Matthey sells the Katalco™ 27–2 high-temperature ammonia cracking catalysts that typically operate in the range of 700–950 °C, giving high hydrogen recovery due to the low residual ammonia levels at this temperature range [292]. This catalyst is based on Ni/calcium aluminate, with a Ni content of about 15 wt% [293].

6.3.7. Steam Reforming of Biomass Tar for Purification of Biomass Gasification-Derived Syngas

The gasification of biomasses represents a promising and already quite established approach to produce syngases [294], whose combustion or conversion may give rise to renewable energy, bio-hydrogen and liquid biofuels. However, gasification technologies present a number of drawbacks, among which is the production of tar molecules polluting the syngas. Biomass tar is a complex mixture of quite-heavy organic molecules produced together with other low-volatility oxygenates. These molecules can condense in the cool sections of the plant (like on heat exchangers), producing fouling, or may cause the coking and deactivation of catalysts (such as for the water–gas shift) [295] and anodes of fuel cells [296]. Catalytic steam reforming appears to be one of the most interesting approaches to destroy such molecules. Ni-based steam-reforming catalysts similar to those discussed above for methane steam reforming [297] appear to be active for this purpose but sensitive to sulfur poisoning due to the small amount of H2S usually present in the syngas. However, at relatively high temperatures, Ni-based tar steam-reforming catalysts maintain activity and are applied at the industrial level [298]. In particular, Ni/Al2O3 are active in the steam reforming of tar molecules, although care must be taken at controlling the temperature and during shut down and startups [299]. When doped with Mg or Ca, Ni/Al2O3 systems appear to be among the most active for the catalytic abatement of tars [300].

6.3.8. Steam Reforming, Partial Oxidation and Autothermal and Dry Reforming of Renewables for the Production of Renewable Hydrogen

Renewable hydrogen can be produced by the steam reforming of biomass-derived organic molecules. In particular, bioethanol steam reforming (ESR) [301,302] has been investigated deeply: it is an endothermic reaction and thus favored at relatively high temperatures and low pressures. It can be realized at 600–700 °C over metal catalysts such as supported Ni, Co, Pt or Rh, including Ni/Al2O3 [303], and is frequently improved by alloying [304] with excess water. It seems that this technology is still not realized at the industrial level but is seriously considered for commercial application [305]. It has been considered that ESR can be realized using existing hydrocarbon steam-reforming plants, provided effective catalysts have been developed [306]. Additionally, it has been shown that Ni/Ca aluminate commercial catalysts for natural gas steam reforming are excellent catalysts for ESR too in laboratory experiments [307].
Steam-reforming processes to produce hydrogen can also be applied to other biomass-derived substances [308] such as vegetable oils [309], wastes from agricultural manufacturing, glycerol and biomass-pyrolysis oil [310]. Also, the partial oxidation and autothermal reforming processes or dry reforming processes of biomass-derived organics have been investigated and can be realized over Ni/Al2O3 catalysts [311,312,313].

6.3.9. Ni/Al2O3 for Chemical Looping Processes

Chemical looping reforming (CLR) technologies imply the separation of the reforming step occurring over an “oxidized” catalyst surface by the reoxidation step of the catalyst using steam, CO2 or oxygen [314]. Ni/Al2O3-based materials have been studied for ethanol, glycerol and methane CLR. In the case of glycerol CLR and in the case of glycerol CL dry reforming, the addition of alkali earth, in particular Sr, improves the performances [315]. NiO/LaAl11O18 was reported to be an excellent material for chemical looping combustion, although its structure was converted into NiO-NiAl2O4-αAl2O3 in the process [316].

6.3.10. Ni/Alumina Catalysts in Dehydrogenation and Aromatization Reactions

Reduced nickel catalysts have limited application in paraffin dehydrogenation reactions because of the high activity of metallic nickel particles in causing C-C bond cracking and deactivation by carbon deposition at high temperatures [317]. However, early studies showed excellent activities of Ni/Al2O3 catalysts for the aromatization of cyclohexanes, a reaction realized at low temperatures, e.g., at 300 °C. [318]. In particular, good activity has been reported for a 20 wt% Ni/Al2O3 [319] catalyst and of NiAl2O4 [320] in the dehydrogenation of methylcyclohexane to toluene at 380–450 °C. This shows that the aromatization and dearomatization activity of quite cheap Ni/Al2O3 catalysts can be interesting in view of the use of the methylcyclohexane/toluene (MCH-TOL) system as a Liquid Organic Hydrogen Carrier (LOHC) [321,322,323], which is already available commercially [324].
A recent work reported that a carefully prepared unreduced 1% Ni/Al2O3 catalyst has high activity and selectivity in propane dehydrogenation to propene at 580 °C, where the active site is supposed to be a highly isolated Ni2+ ion on alumina [196].

6.4. Applications of Unreduced NiO-Al2O3 Catalysts

6.4.1. NiO and NiO-Al2O3 as Combustion Catalysts

As already stated, NiO shows the highest catalytic activity for methane combustion, among simple transition metal oxides, with an ignition temperature in the range of 450–500 °C, attributed by reaction with surface oxide ions interacting with Ni3+ ions [189]. NiO is considered among the useful catalysts for the combustion of Volatile Organic Compounds (VOCs) [325] such as, e.g., in the total oxidation of toluene and formaldehyde [326]. Also, NiO/Al2O3 was reported to have good activity in methanol [327], toluene [328] and isopropanol full oxidation [329].

6.4.2. The Oxidation of Ammonia and Hydrogen

The combustion of hydrogen and the oxidation of ammonia are becoming more and more interesting in view of the large use of hydrogen as a green fuel and of ammonia as a hydrogen transport molecule. Among the non-noble, metal-based catalysts, NiO shows very high activity in hydrogen combustion [330]. Instead, NiO does not seem to have high activity in ammonia oxidation to NO, although some of its mixed oxides do have [331]. Interesting tests have been realized concerning the Ammonia Selective Oxidation (SCO) to nitrogen. Here, 10–20 wt% NiO/Al2O3 [328,332,333] and NiOMgO-Al2O3 [334] catalysts have high activity in the oxidation of ammonia, with a main production of N2 (up to >90% yield) at about 550 °C but also some coproduction of NO.
Ni/Al2O3 has been tested in the combustion of H2-NH3 mixtures and shows the highest activity among tested transition metal catalysts. Over 5%Ni/Al2O3, the light-off of both H2 combustion and NH3 oxidation, is below 200 °C, with a limited coproduction of N2O and NOx [335].

6.4.3. Oxidation of Carbon Monoxide

The low-temperature oxidation of carbon monoxide is an important reaction for treating combustion and automotive flue gases. Bulk and supported NiO are among most active non-noble metal catalytic materials [336,337]. In particular, NiO/Al2O3 [338,339] shows good activity and stability. Interestingly, high activity in CO oxidation has been reported for a commercial NiO/Al2O3-based steam-reforming catalyst (G-65 from Clariant) whose composition is reported to be 34 wt% NiO on alumina at 180–210 °C with 7–25 kPa CO and 5–15 kPa O2 partial pressure in nitrogen [340], which is in agreement with the activity of a 10% NiO/γ-Al2O3 laboratory catalyst [329].

6.4.4. Oxidative Dehydrogenation of Alkanes

The oxidative dehydrogenation (ODH) of light alkanes is a useful reaction to produce olefins starting from natural-gas-derived paraffins with an exothermic process. A totl of 15 wt% NiO/Al2O3 catalysts appear to be very promising for the production of ethylene from ethane ODH [341] at 400 °C with ethylene selectivity of about 95%. The catalyst works in the oxidized state, with the best activity near a monolayer NiO coverage. Similarly, propene can be produced from propane ODH over a NiO/Ca12Al14O33 catalyst at 480–580 °C but with definitely lower selectivity and yield [342].

6.4.5. Ozone Decomposition and Use

NiO and NiO/Al2O3 are also reported among the most active ozone decomposition catalysts [329,343], as well as catalysts for ozonation reactions, i.e., the oxidation of CO and VOC with ozone [329].

7. Conclusions

The above review summarizes, hopefully in an almost complete way, the properties of the materials based on nickel and alumina and most of their actual and potential applications. As reported by Yu et al. [344], catalysts belonging to this system may also have a number of other less usual applications. It is evident that nickel and alumina represent together an extremely versatile system that is applicable in reducing and oxidizing conditions, offering stability in very different conditions, and is also cheap. A main drawback of this system is due to the toxicity of several nickel compounds that make cautions necessary in catalyst preparation and for recovery after their end of life.
The usefulness of Ni–alumina systems is due to the properties of nickel to act as a versatile catalytically active phase in many different conditions. A key property of nickel is to be quite easily reducible to metal but also quite easily oxidized to NiO up to Ni3+-containing NiO1+x. Together, the unique ability of alumina to disperse ionic species at its surface in a stable way, without allowing nickel ions to penetrate the bulk in mild conditions, and to interact strongly with surface metal centers is the other key feature of this system. This allows one to produce small metallic particles with high stability, with moderate nickel loading, or larger particles with high loadings. By changing the alumina polymorphs from γ-Al2O3 or η-Al2O3 to δ-Al2O3, θ-Al2O3 and α-Al2O3, coprecipitating NiAl2O4 and/or NiO-NiAl2O4-Al2O3 solid solutions, or adding alkali-, alkali earth- or rare earth oxides, it is also possible to tune the thermal stability of the system from relatively low temperatures to very high ones and the acido-basicity as well.
Nickel alumina materials may find increased interest in the near future in the frame of hydrogen technologies as hydrogenation and dehydrogenation catalysts. In particular, they can act as optimal catalysts, e.g., for both the hydrogenation and dehydrogenation step in the case of the MCH-TOL system being used as an LOHC or for the decomposition step when ammonia is used as a circular hydrogen carrier.
As shown above, nickel–alumina catalysts can also be used for a number of reactions based on renewables and can be of interest in the energy sector, whose future application will grow significantly. Taking into account that noble metals more active in catalysis, palladium and platinum, in particular, are becoming more and more critical due to their limited availability and increasing applications, alternatives are needed. Although nickel, too, has been defined as a critical metal, its still significant abundance, the relative wide resource distribution and the application of new technologies for extraction from poor resources [345] and for recovery from used devices suggest that the application of nickel in the catalysis field (as well as in the parent electrocatalysis field) may grow very much in the near future. Thus, interest for the nickel–alumina material, very relevant in the past hydrocarbon-based industrial chemistry and energy sectors, will grow further in the future renewable-based industrial chemistry and energy sectors.

Author Contributions

Conceptualization, methodology and writing—review and editing: G.B., E.S., P.R. and G.G. All authors have read and agreed to the published version of the manuscript.

Funding

G.G. and E.S. acknowledge the National Recovery and Resilience Plan (NRRP), Mission 4 Component 2 Investment 1.3—Call for tender No. 1561 of 11.10.2022 of Ministero dell’Uni-versità e della Ricerca (MUR); funded by the European Union—NextGenerationEU project title “Network 4 Energy Sustainable Transition—NEST” (project code PE0000021).

Data Availability Statement

The data in this review paper are obtained from the cited references.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Busca, G. Heterogeneous Catalytic Materials; Elsevier: Amsterdam, The Netherlands, 2014. [Google Scholar]
  2. Busca, G. Critical aspects of energetic transition technologies and the roles of materials chemistry and engineering. Energies 2024, 17, 3565. [Google Scholar] [CrossRef]
  3. Hughes, A.E.; Haque, N.; Northey, S.A.; Giddey, S. Platinum Group Metals: A Review of Resources, Production and Usage with a Focus on Catalysts. Resources 2021, 10, 93. [Google Scholar] [CrossRef]
  4. Ananikov, V.P. Nickel: The “Spirited Horse” of Transition Metal Catalysis. ACS Catal. 2015, 5, 1964–1971. [Google Scholar] [CrossRef]
  5. Serov, A. Nickel catalysts for affordable fuel cells. Nat. Catal. 2022, 5, 971–972. [Google Scholar] [CrossRef]
  6. Bartholomew, C.H.; Farrauto, R.J. Chemistry of nickel-alumina catalysts. J. Catal. 1976, 45, 41–53. [Google Scholar] [CrossRef]
  7. Perry, R.H.; Green, D.W. Perry’s Chemical Engineers’ Handbook, 7th ed.; Mc Graw Hill Pub.: New York, NY, USA, 1997; Chapter 4. [Google Scholar]
  8. Hall, D.S.; Lockwood, D.J.; Bock, C.; MacDougall, B.R. Nickel hydroxides and related materials: A review of their structures, synthesis and properties. Proc. R. Soc. A 2015, 471, 20140792. [Google Scholar] [CrossRef] [PubMed]
  9. Li, Y.F.; Li, J.L.; Liu, Z.P. Structure and Catalysis of NiOOH: Recent Advances on Atomic Simulation. J. Phys. Chem. C 2021, 125, 27033–27045. [Google Scholar] [CrossRef]
  10. Kannan, K.; Radhika, D.; Kumar Sadasivuni, K.; Raghava Reddy, K.; Raghu, A.V. Nanostructured metal oxides and its hybrids for photocatalytic and biomedical applications. Adv. Colloid Interface Sci. 2020, 281, 102178. [Google Scholar] [CrossRef]
  11. Hussein, N.A.; Ali, N.A. Ni2O3 nanomaterial: Synthesis and characterization by simple chemical process. AIP Conf. Proc. 2023, 2834, 030021. [Google Scholar]
  12. Fingerle, M.; Tengeler, S.; Calvetz, W.; Jaegermann, W.; Mayer, T. Sputtered Nickel Oxide Thin Films on n-Si(100)/SiO2 Surfaces for Photo-Electrochemical Oxygen Evolution Reaction (OER): Impact of Deposition Temperature on OER Performance and on Composition before and after OER. J. Electrochem. Soc. 2020, 167, 136514. [Google Scholar] [CrossRef]
  13. Pereñíguez, R.; González-DelaCruz, V.M.; Holgado, J.P.; Caballero, A. Synthesis and characterization of a LaNiO3 perovskite as precursor for methane reforming reactions catalysts. Appl. Catal. B Environ. 2010, 93, 346–353. [Google Scholar] [CrossRef]
  14. Nagle-Cocco, L.A.V.; Bull, C.L.; Ridley, C.J.; Dutton, S.E. Pressure Tuning the Jahn−Teller Transition Temperature in NaNiO2. Inorg. Chem. 2022, 61, 4312–4321. [Google Scholar] [CrossRef]
  15. Li, Y.; Tan, Z.; Liu, Y.; Lei, C.; He, P.; Li, J.; He, Z.; Cheng, Y.; Wu, F.; Li, Y. Past, present and future of high-nickel materials. Nano Energy 2024, 119, 109070. [Google Scholar] [CrossRef]
  16. Genreith-Schriever, A.R.; Alexiu, A.; Phillips, G.S.; Coates, C.S.; Nagle-Cocco, L.A.V.; Bocarsly, J.D.; Sayed, F.N.; Dutton, S.E.; Grey, C.P. Jahn−Teller Distortions and Phase Transitions in LiNiO2: Insights from Ab Initio Molecular Dynamics and Variable-Temperature X-ray Diffraction. Chem. Mater. 2024, 36, 2289–2303. [Google Scholar] [CrossRef]
  17. Busca, G. The surface of transition aluminas. A critical Review. Catal. Today 2014, 226, 2–13. [Google Scholar] [CrossRef]
  18. Busca, G. Structural, surface and catalytic properties of aluminas. Adv. Catal. 2014, 57, 319–404. [Google Scholar]
  19. Yang, Y.; Miao, C.; Wang, R.; Zhang, R.; Li, X.; Wang, J.; Wang, X.; Yao, J. Advances in morphology-controlled alumina and its supported Pd catalysts: Synthesis and applications. Chem. Soc. Rev. 2024, 53, 5014–5053. [Google Scholar] [CrossRef]
  20. Violante, A.; Huang, P.M. Influence of inorganic and organic ligands on the formation of aluminum hydroxides and oxyhydroxides. Clays Clay Min. 1985, 33, 181–192. [Google Scholar] [CrossRef]
  21. Pereira Antunes, M.L.; de Souza Santos, H.; de Souza Santos, P. Characterization of the aluminum hydroxide microcrystals formed in some alcohol−water solutions. Mater. Chem. Phys. 2002, 76, 243–249. [Google Scholar] [CrossRef]
  22. Wefers, K.; Mitra, C. Oxides and Hydroxides of Aluminum; Alcoa Laboratories: New Kensington, PA, USA, 1987. [Google Scholar]
  23. Tsuchida, T. Preparation and reactivity of acicular α-Al2O3 from synthetic diaspore, β-Al2O3·H2O. Solid State Ion. 1993, 63–65, 464–470. [Google Scholar] [CrossRef]
  24. Marturano, M.; Aglietti, E.F.; Ferretti, O. α-Al2O3 catalyst supports for synthesis gas production: Influence of different alumina bonding agents on support and catalyst properties. Mater. Chem. Phys. 1997, 47, 252–256. [Google Scholar] [CrossRef]
  25. Deng, B.; Advincula, P.A.; Luong, D.X.; Zhou, J.; Zhang, B.; Wang, Z.; McHugh, E.A.; Chen, J.; Carter, R.A.; Kittrell, C.; et al. High-surface-area corundum nanoparticles by resistive hotspot-induced phase transformation. Nat. Commun. 2022, 13, 5027. [Google Scholar] [CrossRef]
  26. Levin, I.; Brandon, D. Metastable Alumina Polymorphs: Crystal Structures and Transition Sequences. J. Am. Ceram. Soc. 1998, 81, 1995–2012. [Google Scholar] [CrossRef]
  27. Dai, J.; Li, S.F.Y.; Xiao, T.D.; Wang, D.M.; Reisner, D.E. Structural stability of aluminum stabilized alpha nickel hydroxide as a positive electrode material for alkaline secondary batteries. J. Power Sources 2000, 89, 40–45. [Google Scholar] [CrossRef]
  28. Ogata, F.; Nakamura, T.; Toda, M.; Otani, M.; Kawasaki, N. Evaluation of Nickel–Aluminium Complex Hydroxide for Adsorption of Chromium(VI) Ion. Chem. Pharm. Bull. 2020, 68, 70–76. [Google Scholar] [CrossRef] [PubMed]
  29. Cavani, F.; Trifirò, F.; Vaccari, A. Hydrotalcite-type anionic clays: Preparation, properties and applications. Catal. Today 1991, 11, 173–301. [Google Scholar] [CrossRef]
  30. Debecker, D.P.; Gaigneaux, E.M.; Busca, G. Exploring, tuning and exploiting the basicity of hydrotalcites for applications in heterogeneous catalysis. Chem. Eur. J. 2009, 15, 3920–3935. [Google Scholar] [CrossRef] [PubMed]
  31. Kruissink, E.C.; Van Reijen, L.L.; Ross, J.R.H. Coprecipitated nickel-alumina catalysts for methanation at high temperature. Part 1.—Chemical composition and structure of the precipitates. J. Chem. Soc. Faraday Trans. 1981, 1, 649–663. [Google Scholar] [CrossRef]
  32. Clause, O.; Rebours, B.; Merlen, E.; Trifiró, F.; Vaccari, A. Preparation and characterization of nickel-aluminum mixed oxides obtained by thermal decomposition of hydrotalcite-type precursors. J. Catal. 1992, 133, 231–246. [Google Scholar] [CrossRef]
  33. Busca, G.; Lorenzelli, V.; Sanchez Escribano, V. Preparation, solid-state characterization, and surface chemistry of high surface area NixAl2-2xO3-2x mixed oxides. Chem. Mater. 1992, 4, 595–605. [Google Scholar] [CrossRef]
  34. Busca, G.; Lorenzelli, V.; Sanchez Escribano, V.; Guidetti, R. FT-IR study of the surface properties of the spinels NiAl2O4 and CoAl2O4 in relation to those of transitional aluminas. J. Catal. 1991, 131, 167. [Google Scholar] [CrossRef]
  35. Morales-Marína, A.; Ayastuya, J.L.; Iriarte-Velasco, U.; Gutiérrez-Ortiz, M.A. Nickel aluminate spinel-derived catalysts for the aqueous phase reforming of glycerol: Effect of reduction temperature. Appl. Catal. B Environ. 2019, 244, 931–945. [Google Scholar] [CrossRef]
  36. Raghavan, V. Al-Fe-Ni-O (Aluminum-Iron-Nickel-Oxygen). J. Phase Equilib. Diffus. 2010, 31, 377–378. [Google Scholar] [CrossRef]
  37. O’Neill, H.S.C.; Dollase, W.A.; Ross, C.R. Temperature dependence of the cation distribution in nickel aluminate (NiAl2O4) spinel: A powder XRD study. Phys. Chem. Miner. 1991, 18, 302–319. [Google Scholar] [CrossRef]
  38. Gil-Calvo, M.; Jiménez-González, C.; de Rivas, B.; Gutiérrez-Ortiz, J.I.; López-Fonseca, R. Effect of Ni/Al molar ratio on the performance of substoichiometric NiAl2O4 spinel-based catalysts for partial oxidation of methane. Appl. Catal. B Environ. 2017, 209, 128–138. [Google Scholar] [CrossRef]
  39. Rogers, J.L.; Mangarella, M.C.; D’Amico, A.D.; Gallagher, J.R.; Dutzer, M.R.; Stavitski, E.; Miller, J.T.; Sievers, C. Differences in the Nature of Active Sites for Methane Dry Reformingand Methane Steam Reforming over Nickel Aluminate Catalysts. ACS Catal. 2016, 6, 5873–5886. [Google Scholar] [CrossRef]
  40. Bassoul, P.; Gilles, J.C. Structure and microstructure of the metastable B phase (NiAl10O16): I. Preparation and structural study by X-ray diffraction. J. Solid State Chem. 1985, 58, 383–388. [Google Scholar] [CrossRef]
  41. Odaka, A.; Yamaguchi, T.; Fujita, T.; Taruta, S.; Kitajima, K. Cation dopant effect on phase transformation and microstructural evolution in M2+-substituted γ-alumina powders. J. Mater. Sci. 2008, 43, 2713–2720. [Google Scholar] [CrossRef]
  42. Azurdia, J.A.; Marchal, J.; Shea, P.; Sun, H.; Pan, X.Q.; Laine, R.M. Liquid-Feed Flame Spray Pyrolysis as a Method of Producing Mixed-Metal Oxide Nanopowders of Potential Interest as Catalytic Materials. Nanopowders along the NiO-Al2O3 Tie Line Including (NiO)0.22(Al2O3)0.78, a New Inverse Spinel Composition. Chem. Mater. 2006, 18, 731–739. [Google Scholar] [CrossRef]
  43. Piao, L.; Li, Y.; Chen, J.; Chang, L.; Lin, J.Y.S. Methane decomposition to carbon nanotubes and hydrogen on an alumina supported nickel aerogel catalyst. Catal. Today 2002, 74, 145–155. [Google Scholar] [CrossRef]
  44. Al-Nakoua, M.A.; El-Naas, M.H.; Abu-Jdayil, B. Characterization and testing of sol–gel catalysts prepared as thin layers in a plate reactor. Fuel Process. Technol. 2011, 92, 1836–1841. [Google Scholar] [CrossRef]
  45. Komeili, S.; Taeb, A.; Ravanchi, M.T.; Fard, M.R. The properties of nickel aluminate nanoparticles prepared by sol–gel and impregnation methods. Res. Chem. Intermed. 2016, 42, 7909–7921. [Google Scholar] [CrossRef]
  46. Peña, J.A.; Herguido, J.; Guimon, C.; Monzón, A.; Santamaría, J. Hydrogenation of Acetylene over Ni/NiAl2O4 Catalyst: Characterization, Coking, and Reaction Studies. J. Catal. 1996, 159, 313–322. [Google Scholar] [CrossRef]
  47. Ghule, A.V.; Ghule, K.; Punde, T.; Liu, J.Y.; Tzing, S.H.; Chang, J.Y.; Chang, H.; Ling, J.C. In situ monitoring of NiO–Al2O3 nanoparticles synthesis by thermo-Raman spectroscopy. Mater. Chem. Phys. 2010, 119, 86–92. [Google Scholar] [CrossRef]
  48. Li, C.Y.; Zhang, H.J.; Chen, Z.Q. Reaction between NiO and Al2O3 in NiO/γ-Al2O3 catalysts probed by positronium Atom. Appl. Surf. Sci. 2013, 266, 17–21. [Google Scholar] [CrossRef]
  49. Garbarino, G.; Chitsazan, S.; Phung, T.K.; Riani, P.; Busca, G. Preparation of supported catalysts: A study of the effect of small amounts of silica on Ni/Al2O3 catalysts. Appl. Catal. A Gen. 2015, 505, 86–97. [Google Scholar] [CrossRef]
  50. Riani, P.; Spennati, E.; Villa Garcia, M.; Sanchez Escribano, V.; Busca, G.; Garbarino, G. Ni/Al2O3 catalysts for CO2 methanation: Effect of silica and nickel loading. Int. J. Hydrogen Energy 2023, 48, 24976–24995. [Google Scholar] [CrossRef]
  51. Jurczyk, K.; Kania, W. Acid-base properties of modified γ-alumina. Appl. Catal. 1987, 34, 1–12. [Google Scholar]
  52. Jurczyk, K.; Kania, W. Base Properties of Modified y-Alumina. Appl. Catal. 1989, 56, 253–261. [Google Scholar] [CrossRef]
  53. Margossian, T.; Larmier, K.; Kim, S.M.; Krumeich, F.; Fedorov, A.; Chen, P.; Muller, C.R.; Coperet, C. Molecularly Tailored Nickel Precursor and Support Yield a Stable Methane Dry Reforming Catalyst with Superior Metal Utilization. J. Am. Chem. Soc. 2017, 139, 6919–6927. [Google Scholar] [CrossRef]
  54. Busca, G.; Riani, P.; Garbarino, G.; Ziemacki, G.; Gambino, L.; Montanari, E.; Millini, R. The state of nickel in spent Fluid Catalytic Cracking catalysts. Appl. Catal. A Gen. 2014, 486, 176–186. [Google Scholar] [CrossRef]
  55. Li, M.; Fang, S.; Hu, Y.H. Self-stabilization of Ni/Al2O3 Catalyst with a NiAl2O4 Isolation Layer in Dry Reforming of Methane. Catal. Lett. 2022, 152, 2852–2859. [Google Scholar] [CrossRef]
  56. Williams, A.; Butler, G.A.; Hammonds, J. Sintering of Nickel-Alumina Catalysts. J. Catal. 1972, 24, 352–355. [Google Scholar] [CrossRef]
  57. Garbarino, G.; Valsamakis, I.; Riani, P.; Busca, G. On the consistency of results arising from different techniques concerning the nature of supported metal oxide (nano)particles. The case of NiO/Al2O3. Catal. Commun. 2014, 51, 37–41. [Google Scholar] [CrossRef]
  58. Nikolopoulos, I.; Kogkos, G.; Tsavatopoulou, V.D.; Kordouli, E.; Bourikas, K.; Kordulis, C.; Lycourghiotis, A. Nickel—Alumina Catalysts for the Transformation of Vegetable Oils into Green Diesel: The Role of Preparation Method, Activation Temperature, and Reaction Conditions. Nanomaterials 2023, 13, 616. [Google Scholar] [CrossRef] [PubMed]
  59. Clariant, Catalysts and Adsorbents for Syngas. Available online: https://www.clariant.com/en/Business-Units/Catalysts/Syngas-Catalysts (accessed on 24 July 2024).
  60. Molina, R.; Poncelet, G. α-Alumina-Supported Nickel Catalysts Prepared from Nickel Acetylacetonate: A TPR Study. J. Catal. 1998, 173, 257–267. [Google Scholar] [CrossRef]
  61. Bolt, P.; Habraken, F.H.P.M.; Geus, J.W. On the Role of a NiAl2O4 Intermediate Layer in the Sintering Behavior of Ni/α-Al2O3. J. Catal. 1995, 151, 300–306. [Google Scholar] [CrossRef]
  62. Azancot, L.; Bobadilla, L.F.; Santos, J.L.; Córdoba, J.M.; Centeno, M.A.; Odriozola, J.A. Influence of the preparation method in the metal-support interaction and reducibility of Ni-Mg-Al based catalysts for methane steam reforming. Int. J. Hydrogen Energy 2019, 44, 19827–19840. [Google Scholar] [CrossRef]
  63. Yang, J.; Xiao, H. Preparation and Oxidation Behavior of Metallic Nickel Containing MgAlON Composite. High Temp. Mater. Proc. 2018, 37, 563–569. [Google Scholar]
  64. Jacob, K.T.; Alcock, C.B. Activities and their relation to cation distribution in NiAl2O4-MgAl2O4 spinel solid solutions. J. Solid State Chem. 1977, 20, 79–88. [Google Scholar] [CrossRef]
  65. Qi, C.; Spagnoli, D.; Fourie, A. Structural, electronic, and mechanical properties of calcium aluminate cements: Insight from first-principles theory. Construct. Build. Mater. 2020, 264, 120259. [Google Scholar] [CrossRef]
  66. Roussière, T.; Schelkle, K.M.; Titlbach, S.; Wasserschaff, G.; Milanov, A.; Cox, G.; Schwab, E.; Deutschmann, O.; Schulz, L.; Jentys, A.; et al. Structure–Activity Relationships of Nickel–Hexaaluminates in Reforming Reactions Part I: Controlling Nickel Nanoparticle Growth and Phase Formation. ChemCatChem 2014, 6, 1438–1446. [Google Scholar] [CrossRef]
  67. Johnson Matthey, Methanation Catalysts. Available online: https://matthey.com/documents/161599/440146/JM+Methanation+product+brochure+%28c2020%29.pdf/56eaef18-4782-776f-ed2a-c332f5f16659?t=1653488109779 (accessed on 20 July 2024).
  68. Rossi, P.F.; Busca, G.; Lorenzelli, V.; Waqif, M.; Saur, O.; Lavalley, J.C. Surface basicity of mixed oxides: Magnesium and zinc aluminates. Langmuir 1991, 7, 2677–2681. [Google Scholar] [CrossRef]
  69. Busca, G.; Cristiani, C.; Forzatti, P.; Groppi, G. Surface characterization of Ba-β-alumina. Catal. Lett. 1995, 31, 65–74. [Google Scholar] [CrossRef]
  70. Garbarino, G.; Finocchio, E.; Lagazzo, A.; Valsamakis, I.; Riani, P.; Sanchez Escribano, V.; Busca, G. Steam reforming of ethanol–phenol mixture on Ni/Al2O3: Effect of magnesium and boron on catalytic activity in the presence and absence of sulphur. Appl. Catal. B Environ. 2014, 147, 813–826. [Google Scholar] [CrossRef]
  71. Van Teijlingen, A.; Davis, S.A.; Hall, S.R. Size-dependent melting point depression of nickel nanoparticles. Nanoscale Adv. 2020, 2, 2347–2351. [Google Scholar] [CrossRef]
  72. Riani, P.; Garbarino, G.; Lucchini, M.A.; Canepa, F.; Busca, G. Unsupported versus alumina-supported Ni nanoparticles as catalysts for steam/ethanol conversion and CO2 methanation. J. Mol. Catal. A Chem. 2014, 383–384, 10–16. [Google Scholar] [CrossRef]
  73. Riani, P.; Garbarino, G.; Infantes-Molina, A.; Rodríguez-Castellón, E.; Canepa, F.; Busca, G. Hydrogen from steam reforming of ethanol over cobalt nanoparticles: Effect of boron impurities. Appl. Catal. A Gen. 2016, 518, 67–77. [Google Scholar] [CrossRef]
  74. Peng, B.; Zhang, X.; Aarts, D.G.A.L.; Dullens, R.P.A. Superparamagnetic nickel colloidal nanocrystal clusters with antibacterial activity and bacteria binding ability. Nat. Nanotechnol. 2018, 13, 478–482. [Google Scholar] [CrossRef]
  75. Rodríguez-González, V.; Marceau, E.; Beaunier, P.; Che, M.; Train, C. Stabilization of hexagonal close-packed metallic nickel for alumina-supported systems prepared from Ni(II) glycinate. J. Solid State Chem. 2007, 180, 22–30. [Google Scholar] [CrossRef]
  76. Su, J.; Wang, Q.; Fang, M.; Wang, Y.; Ke, J.; Shao, Q.; Lu, J. Metastable Hexagonal-Phase Nickel with Ultralow Pt Content for anEfficient Alkaline/Seawater Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2023, 15, 51160–51169. [Google Scholar] [CrossRef]
  77. Li, Z.; Wen, X.; Chen, F.; Zhang, Q.; Zhang, Q.; Gu, L.; Cheng, J.; Wu, B.; Zheng, N. Hexagonal Nickel as a Highly Durable and Active Catalyst for Hydrogen Evolutione. ACS Catal. 2021, 11, 8798–8806. [Google Scholar] [CrossRef]
  78. Shi, D.; Wen, B.; Melnik, R.; Yao, S.; Li, T. First-principles studies of Al–Ni intermetallic compounds. J. Solid State Chem. 2009, 182, 2664–2669. [Google Scholar] [CrossRef]
  79. Paul, A.R.; Mukherjee, M.; Singh, D. A Critical Review on the Properties of IntermetallicCompounds and Their Application in the Modern Manufacturing. Cryst. Res. Technol. 2022, 57, 2100159. [Google Scholar] [CrossRef]
  80. Sampath, S.; Ravi, V.P.; Sundararajan, S. An Overview on Synthesis, Processing and Applications of Nickel Aluminides: From Fundamentals to Current Prospects. Crystals 2023, 13, 435. [Google Scholar] [CrossRef]
  81. Villarroel-Rocha, J.; Gil, A. Modeling the temperature-programmed reduction of metal oxide catalysts by considering the particle-size distribution effect. Chem. Eng. J. 2024, 487, 150722. [Google Scholar] [CrossRef]
  82. Manukyan, K.V.; Avetisyan, A.G.; Shuck, C.E.; Chatilyan, H.A.; Rouvimov, S.; Kharatyan, S.L.; Mukasyan, A.S. Nickel Oxide Reduction by Hydrogen: Kinetics and Structural Transformations. J. Phys. Chem. C 2015, 119, 16131–16138. [Google Scholar] [CrossRef]
  83. Braaten, O.; Kjekshus, A.; Kvande, H. The Possible Reduction of Alumina to Aluminum Using Hydrogen. JOM 2000, 52, 47–53. [Google Scholar] [CrossRef]
  84. Helali, Z.; Jedidi, A.; Syzgantseva, O.A.; Calatayud, M.; Minot, C. Scaling reducibility of metal oxides. Theor. Chem. Acc. 2017, 136, 100. [Google Scholar] [CrossRef]
  85. Song, P.; Wen, D.; Guob, Z.X.; Korakianitis, T. Oxidation investigation of nickel nanoparticles. Phys. Chem. Chem. Phys. 2008, 10, 5057–5065. [Google Scholar] [CrossRef]
  86. Unutulmazsoy, Y.; Merkle, R.; Fischer, D.; Mannhart, J.; Maier, J. The oxidation kinetics of thin nickel films between 250 and 500 °C. Phys. Chem. Chem. Phys. 2017, 19, 9045–9052. [Google Scholar] [CrossRef] [PubMed]
  87. Payne, B.P.; Grosvenor, A.P.; Biesinger, M.C.; Kobe, B.A.; McIntyre, N.S. Structure and growth of oxides on polycrystalline nickel surfaces. Surf. Interface Anal. 2007, 39, 582–592. [Google Scholar] [CrossRef]
  88. Payne, B.P.; Biesinger, M.C.; McIntyre, N.S. The study of polycrystalline nickel metal oxidation by water vapour. J. Electron Spectrosc. Relat. Phenom. 2009, 175, 55–65. [Google Scholar] [CrossRef]
  89. Yavuz, A.; Yilmaz, N.F.; Kalkan, M.F. Enhancement in the Corrosion Resistance of Nickel Metal Via a Straightforward Thermal Oxidation Method. J. Bio-Tribo-Corros. 2023, 9, 22. [Google Scholar] [CrossRef]
  90. Kumar, S.N.; Appari, S.; Vardhan, B.; Kuncharam, R. Techniques for Overcoming Sulfur Poisoning of Catalyst Employed in Hydrocarbon Reforming. Catal. Surv. Asia 2021, 25, 362–388. [Google Scholar] [CrossRef]
  91. Argyle, M.D.; Bartholomew, C.H. Heterogeneous Catalyst Deactivation and Regeneration: A Review. Catalysts 2015, 5, 145–269. [Google Scholar] [CrossRef]
  92. Hernandez, J.M.; Lim, D.H.; Nguyen, H.V.P.; Yoon, S.P.; Han, J.; Nam, S.W.; Yoon, C.W.; Kim, S.K.; Ham, H.C. Decomposition of hydrogen sulfide (H2S) on Ni(100) and Ni3Al(100) surfaces from first-principles. Int. J. Hydrogen Energy 2014, 39, 12251–12258. [Google Scholar] [CrossRef]
  93. Wang, L.; Zhu, Y.; Li, H.; Li, Q.; Qian, Y. Hydrothermal synthesis of NiS nanobelts and NiS2 microspheres constructed of cuboids architectures. J. Solid State Chem. 2010, 183, 223–227. [Google Scholar] [CrossRef]
  94. Rochet, A.; Ribeiro Passos, A.; Legens, C.; Briois, V. Sulphidation study of a dried Ni/Al2O3 catalyst by time-resolved XAS-MS combined with in situ Raman spectroscopy and multivariate Quick-XAS data analysis. Catal. Struct. React. 2017, 3, 33–42. [Google Scholar] [CrossRef]
  95. Garbarino, G.; Riani, P.; Magistri, L.; Busca, G. A study of the methanation of carbon dioxide on Ni/Al2O3 catalysts at atmospheric pressure. Int. J. Hydrogen Energy 2014, 39, 11557–11565. [Google Scholar] [CrossRef]
  96. Enger, B.C.; Holmen, A. Nickel and Fischer-Tropsch Synthesis. Catal. Rev. 2012, 54, 437–488. [Google Scholar] [CrossRef]
  97. Struis, R.P.W.J.; Bachelin, D.; Ludwig, C.; Wokaun, A. Studying the Formation of Ni3C from CO and Metallic Ni at T = 265 °C in Situ Using Ni K-Edge X-ray Absorption Spectroscopy. J. Phys. Chem. C 2009, 113, 2443–2451. [Google Scholar] [CrossRef]
  98. Hammer, B.; Nørskov, J.K. Why gold is the noblest of all the metals. Nature 1995, 376, 238. [Google Scholar] [CrossRef]
  99. Hammer, B.; Nørskov, J.K. Theoretical surface science and catalysis—Calculations and concepts. Adv. Catal. 2000, 45, 71–129. [Google Scholar]
  100. Nørskov, J.K.; Abild-Pedersen, F.; Studt, F.; Bligaard, T. Density functional theory in surface chemistry and catalysis. Proc. Natl. Acad. Sci. USA 2011, 108, 937–943. [Google Scholar] [CrossRef] [PubMed]
  101. Sehested, J. Four challenges for nickel steam-reforming catalysts. Catal. Today 2006, 111, 103–110. [Google Scholar] [CrossRef]
  102. Wen, J.G.; Huang, Z.P.; Wang, D.Z.; Chen, J.H.; Yang, S.X.; Rena, Z.F.; Wang, J.H.; Calvet, L.E.; Chen, J.; Klemic, J.F.; et al. Growth and characterization of aligned carbon nanotubes from patterned nickel nanodots and uniform thin films. J. Mater. Res. 2001, 16, 3246–3253. [Google Scholar] [CrossRef]
  103. Helveg, S.; Sehested, J.; Rostrup-Nielsen, J.R. Whisker carbon in perspective. Catal. Today 2011, 178, 42–46. [Google Scholar] [CrossRef]
  104. Ziebro, J.; Skorupińska, S.; Kądziołka, G.; Michalkiewicz, B. Synthesizing Multi-walled Carbon Nanotubes over a Supported-nickel Catalyst, Fullerenes, Nanotubes and Carbon. Nanostructures 2013, 21, 333–345. [Google Scholar]
  105. Lim, X.X.; Low, S.C.; Oh, W.D. A critical review of heterogeneous catalyst design for carbon nanotubes synthesis: Functionalities, performances, and prospects. Fuel Process. Technol. 2023, 241, 107624. [Google Scholar] [CrossRef]
  106. Abbas, H.F.; Wan Daud, W.M.A. Hydrogen production by methane decomposition: A review. Int. J. Hydrogen Energy 2010, 35, 1160–1190. [Google Scholar] [CrossRef]
  107. Goldberger, W.M.; Othmer, D.F. Kinetics of Nickel Carbonyl Formation. Ind. Eng. Chem. Process Des. Dev. 1963, 2, 202–209. [Google Scholar] [CrossRef]
  108. Baranowski, B.; Filipek, S.M. 45 Years of nickel hydride—History and perspectives. J. Alloys Compd. 2005, 404–406, 2–6. [Google Scholar] [CrossRef]
  109. Rana, S.; Masli, N.; Monder, D.S.; Chatterjee, A. Hydriding pathway for Ni nanoparticles: Computational characterization provides insights into the nanoparticle size and facet effect on layer-by-layer subsurface hydride formation. Comp. Mat. Sci. 2022, 210, 111482. [Google Scholar] [CrossRef]
  110. Cecilia, J.A.; Jiménez-Morales, I.; Infantes-Molina, A.; Rodríguez-Castellón, E.; Jiménez-López, A. Influence of the silica support on the activity of Ni and Ni2P based catalysts in the hydrodechlorination of chlorobenzene. Study of factors governing catalyst deactivation. J. Mol. Catal. A Chem. 2013, 368–369, 78–87. [Google Scholar] [CrossRef]
  111. Hao, Q.; Yang, Z.; Wu, B.; Zhu, J.; Li, Z.; Liu, J.; Ma, L. Study on the deactivation of Ni-based catalyst in the hydrotreating process of waste plastic pyrolysis oil. J. Anal. Appl. Pyrol. 2022, 168, 105789. [Google Scholar] [CrossRef]
  112. Ohtsuka, Y. Influence of hydrogen chloride treatment on the dispersion of nickel particles supported on carbon. J. Mol. Catal. 1989, 54, 225–235. [Google Scholar] [CrossRef]
  113. Menini, C.; Park, C.; Brydson, R.; Keane, M.A. Low-Temperature (553 K) Catalytic Growth of Highly Ordered Carbon Filaments during Hydrodechlorination Reactions. J. Phys. Chem. B 2000, 104, 4281–4284. [Google Scholar] [CrossRef]
  114. Morterra, C.; Magnacca, G. A case study: Surface chemistry and surface structure of catalytic aluminas, as studied by vibrational spectroscopy of adsorbed species. Catal. Today 1996, 27, 497–532. [Google Scholar] [CrossRef]
  115. Busca, G. Catalytic materials based on silica and alumina: Structural features and generation of surface acidity. Prog. Mater. Sci. 2019, 104, 215–249. [Google Scholar] [CrossRef]
  116. Escalona Platero, E.; Coluccia, S.; Zecchina, A. CO and NO Adsorption on NiO: A Spectroscopic Investigation. Langmuir 1985, 1, 407–414. [Google Scholar] [CrossRef]
  117. Escalona Platero, E.; Scarano, D.; Zecchina, A.; Meneghini, G.; De Franceschi, R. Highly sintered nickel oxide: Surface morphology and FTIR investigation of CO adsorbed at low temperature. Surf. Sci. 1996, 350, 113–122. [Google Scholar] [CrossRef]
  118. Kitakatsu, N.; Maurice, V.; Hinnen, C.; Marcus, P. Surface hydroxylation and local structure of NiO thin films formed on Ni(111). Surf. Sci. 1998, 407, 36–58. [Google Scholar] [CrossRef]
  119. Zhao, W.; Bajdich, M.; Carey, S.; Vojvodic, A.; Nørskov, J.K.; Campbell, C.T. Water Dissociative Adsorption on NiO(111): Energetics and Structure of the Hydroxylated Surface. ACS Catal. 2016, 6, 7377–7384. [Google Scholar] [CrossRef]
  120. Tsyganenko, A.A.; Pozdnyakov, D.V.; Filimonov, V.N. Infrared study of surface species arising from ammonia adsorption on oxide surfaces. J. Mol. Struct. 1975, 29, 299–318. [Google Scholar] [CrossRef]
  121. Matsumoto, T.; Kubota, J.; Kondo, J.N.; Hirose, C.; Domen, K. Adsorption Structures of Carbon Dioxide on NiO(111) and Hydroxylated NiO(111) Studied by Infrared Reflection Adsorption Spectroscopy. Langmuir 1999, 15, 2158–2161. [Google Scholar] [CrossRef]
  122. Atzori, L.; Cutrufello, M.G.; Meloni, D.; Onida, B.; Gazzoli, D.; Ardu, A.; Monaci, R.; Sini, M.F.; Rombi, E. Characterization and catalytic activity of soft-templated NiO-CeO2 mixed oxides for CO and CO2 co-methanation. Front. Chem. Sci. Eng. 2021, 15, 251–268. [Google Scholar] [CrossRef]
  123. Bordes-Richard, E.; Courtine, P. Optical basicity: A scale of acidity/basicity of solids and its application to oxidation catalysis. In Metal Oxides: Chemistry and Applications; Fierro, J.L.G., Ed.; CRC Press LLC: Boca Raton, FL, USA, 2006; Volume 108, pp. 319–352. [Google Scholar]
  124. Busca, G. Bases and basic materials in industrial and environmental chemistry. Liquid versus solid basicity. Chem. Rev. 2010, 110, 2217–2249. [Google Scholar] [CrossRef] [PubMed]
  125. Huang, X.; Xing, X.; Xiong, W.; Li, H. Facile synthesis of Ni-based oxides nanocatalyst: Effect of calcination temperature on NRR properties of NiO. Carbon Lett. 2024, 34, 1197–1206. [Google Scholar] [CrossRef]
  126. Otero Areán, C.; Peñarroya Mentruit, M.; López López, A.J.; Parra, J.B. High surface area nickel aluminate spinels prepared by a sol–gel methoñd. Coll. Surf. A Physicochem. Eng. Asp. 2001, 180, 253–258. [Google Scholar] [CrossRef]
  127. Garbarino, G.; Campodonico, S.; Romero Perez, A.; Carnasciali, M.M.; Riani, P.; Finocchio, E.; Busca, G. Spectroscopic characterizatin of Ni/Al2O3 catalytic materials for the steam reforming of renewables. Appl. Catal. A Gen. 2013, 452, 163–173. [Google Scholar] [CrossRef]
  128. Salagre, P.; Fierro, J.L.G.; Medina, F.; Sueiras, J.E. Characterization of nickel species on several γ-alumina supported nickel simples. J. Mol. Catal. A Chem. 1996, 106, 125–134. [Google Scholar] [CrossRef]
  129. Li, G.; Hu, L.; Hill, J.M. Comparison of reducibility and stability of alumina-supported Ni catalysts prepared by impregnation and co-precipitation. Appl. Catal. A Gen. 2006, 301, 16–24. [Google Scholar] [CrossRef]
  130. Gericke, S.M.; Rissler, J.; Bermeo, M.; Wallander, H.; Karlsson, H.; Kollberg, L.; Scardamaglia, M.; Temperton, R.; Zhu, S.; Sigfridsson Clauss, K.V.; et al. In situ H2 reduction of Al2O3-supported Ni- and Mo-based catalysts. Catalysts 2022, 12, 755. [Google Scholar] [CrossRef]
  131. Cho, J.H.; An, S.H.; Chang, T.S.; Shin, C.H. Effect of an Alumina Phase on the Reductive Amination of 2-Propanol to Monoisopropylamine Over Ni/Al2O3. Catal. Lett. 2016, 146, 811–819. [Google Scholar] [CrossRef]
  132. He, L.; Ren, Y.; Yue, B.; Tsang, S.C.E.; He, H. Tuning Metal–Support Interactions on Ni/Al2O3 Catalysts to Improve Catalytic Activity and Stability for Dry Reforming of Methane. Processes 2021, 9, 706. [Google Scholar] [CrossRef]
  133. Zieliński, J. Morphology of nickel/alumina catalysts. J. Catal. 1982, 76, 157–163. [Google Scholar] [CrossRef]
  134. Gupta, S.; Deo, G. Effect of metal amount on the catalytic performance of Ni-Al2O3 catalyst for the Tri-reforming of methane. Int. J. Hydrogen Energy 2023, 48, 5478–5492. [Google Scholar] [CrossRef]
  135. Li, C.P.; Chen, Y.-W. Temperature-programmed-reduction studies of nickel oxide/alumina catalysts: Effects of the preparation method. Thermochim. Acta 1995, 256, 457–465. [Google Scholar] [CrossRef]
  136. García-Diéguez, M.; Finocchio, E.; Larrubia, M.A.; Alemany, L.J.; Busca, G. Characterization of alumina-supported Pt, Ni and PtNi alloy catalysts for the dry reforming of methane. J. Catal. 2010, 274, 11–20. [Google Scholar] [CrossRef]
  137. Amenomiya, Y. Adsorption of Hydrogen and H2-D2 Exchange Reaction on Alumina. J. Catal. 1971, 22, 109–121. [Google Scholar] [CrossRef]
  138. Weller, S.W.; Montagna, A.A. Studies of Alumina, I. Reaction with Hydrogen at Elevated Temperatures. J. Catal. 1971, 21, 303–311. [Google Scholar] [CrossRef]
  139. Kramer, R.; Andre, M. Adsorption of Atomic Hydrogen on Alumina by Hydrogen Spillover. J. Catal. 1979, 58, 287–295. [Google Scholar] [CrossRef]
  140. Cabrejas Manchado, M.; Guil, J.M.; Pérez Masiá, A.; Ruiz Paniego, A.; Trejo Menayo, J.M. Adsorption of H2, O2, CO, and CO2 on a γ-Alumina: Volumetric and Calorimetric Studies. Langmuir 1994, 10, 685–691. [Google Scholar] [CrossRef]
  141. Digne, M.; Sautet, P.; Raybaud, P.; Euzen, P.; Toulhoat, H. Use of DFT to achieve a rational understanding of acid–basic properties of γ-alumina surfaces. J. Catal. 2004, 226, 54–68. [Google Scholar] [CrossRef]
  142. Wischert, R.; Laurent, P.; Copéret, C.; Delbecq, F.; Sautet, P. γ-Alumina: The essential and unexpected role of water for the structure, stability, and reactivity of “defect” sites. J. Am. Chem. Soc. 2012, 134, 14430–14449. [Google Scholar] [CrossRef] [PubMed]
  143. Karim, W.; Spreafico, C.; Kleibert, A.; Gobrecht, J.; VandeVondele, J.; Ekinci, Y.; van Bokhoven, J.A. Catalyst support effects on hydrogen spillover. Nature 2017, 541, 68–71. [Google Scholar] [CrossRef]
  144. Garbarino, G.; Travi, I.; Pani, M.; Carnasciali, M.M.; Busca, G. Pure vs ultra-pure γ-alumina: A spectroscopic study and catalysis of ethanol conversion. Catal. Commun. 2015, 70, 77–81. [Google Scholar] [CrossRef]
  145. Shen, H.; Li, H.; Yang, Z.; Li, C. Magic of hydrogen spillover: Understanding and application. Green Energy Environ. 2022, 7, 1161–1198. [Google Scholar] [CrossRef]
  146. Beck, A.; Rzepka, P.; Marshall, K.P.; Stoian, D.; Willinger, M.G.; van Bokhoven, J.A. Hydrogen Interaction with Oxide Supports in the Presence and Absence of Platinum. J. Phys. Chem. C 2022, 126, 17589–17597. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, L.; Zhou, M.; Wang, A.; Zhang, T. Selective Hydrogenation over Supported Metal Catalysts: From Nanoparticles to Single Atoms. Chem. Rev. 2020, 120, 683–733. [Google Scholar] [CrossRef] [PubMed]
  148. Panczyk, T.; Szabelski, P.; Rudzinski, W. Hydrogen Adsorption on Nickel (100) Single-Crystal Face. A Monte Carlo Study of the Equilibrium and Kinetics. J. Phys. Chem. B 2005, 109, 10986–10994. [Google Scholar] [CrossRef]
  149. Znak, L.; Zieliński, J. Effects of support on hydrogen adsorption/desorption on nickel. Appl. Catal. A Gen. 2008, 334, 268–276. [Google Scholar] [CrossRef]
  150. Wang, Y.; Winter, L.R.; Chen, J.G.; Yan, B.Y. CO2 hydrogenation over heterogeneous catalysts at atmospheric pressure: From electronic properties to product selectivity. Green Chem. 2021, 23, 249–267. [Google Scholar] [CrossRef]
  151. Blomqvist, J.; Lehman, L.; Salo, P. CO adsorption on metal-oxide surfaces doped with transition-metal adatoms. Phys. Status Solid. 2012, 249, 1046–1057. [Google Scholar] [CrossRef]
  152. Blyholder, G. Molecular orbital view of chemisorbed carbon monoxide. J. Phys. Chem. 1964, 68, 2772–2777. [Google Scholar] [CrossRef]
  153. Lupinetti, A.J.; Fau, S.; Frenking, G.; Strauss, S.H. Theoretical Analysis of the Bonding between CO and Positively Charged Atoms. J. Phys. Chem. A 1997, 101, 9551–9559. [Google Scholar] [CrossRef]
  154. Gopal, P.G.; Schneider, R.L.; Watters, K.L. Evidence for Production of Surface Formate upon Direct Reaction of CO with Alumina and Magnesia. J. Catal. 1987, 105, 366–372. [Google Scholar] [CrossRef]
  155. Vannice, M.A. Activation of Carbon Monoxide on Metal Surfaces. In Catalysis Science and Technology; Andesron, J.R., Boudart, M., Eds.; Springer: Berlin/Heidelberg, Germany, 1982; Volume 3, p. 139. [Google Scholar]
  156. Abild-Pedersen, F.; Andersson, M.P. CO adsorption energies on metals with correction for high coordination adsorption sites—A density functional study. Surf. Sci. 2007, 601, 1747–1753. [Google Scholar] [CrossRef]
  157. Nguyen, T.T.; Sheppard, N. Advances in Infrared and Raman Spectroscopy; Hester, R.E., Clark, R.H.J., Eds.; Heyden: London, UK, 1982; Volume 5, pp. 86–95. [Google Scholar]
  158. Fielicke, A.; Gruene, P.; Meijer, G.; Rayner, D.M. The adsorption of CO on transition metal clusters: A case study of cluster surface chemistry. Surf. Sci. 2009, 603, 1427–1433. [Google Scholar] [CrossRef]
  159. Vázquez-Parga, D.; Jurado, A.; Roldan, A.; Viñes, F. A computational map of the probe CO molecule adsorption and dissociation on transition metal low Miller indices surfaces. Appl. Surf. Sci. 2023, 618, 156581. [Google Scholar] [CrossRef]
  160. Tøttrup, P.B. Kinetics of Decomposition of Carbon Monoxide on a Supported Nickel Catalyst. J. Catal. 1976, 42, 29–36. [Google Scholar] [CrossRef]
  161. Gałuszka, J.; Chang, J.R.; Amenomiya, Y. Disproportionation of carbon monoxide on supported nickel catalysts. J. Catal. 1981, 68, 172–181. [Google Scholar] [CrossRef]
  162. Liu, Z.P.; Hu, P. General trends in CO dissociation on transition metal surfaces. J. Chem. Phys. 2001, 114, 8244–8247. [Google Scholar] [CrossRef]
  163. Schmider, D.; Maier, L.; Deutschmann, O. Reaction Kinetics of CO and CO2 Methanation over Nickel. Ind. Eng. Chem. Res. 2021, 60, 5792–5805. [Google Scholar] [CrossRef]
  164. Trombetta, M.; Busca, G.; Rossini, S.A.; Piccoli, V.; Cornaro, U. FT-IR studies on light olefin skeletal isomerization catalysis. Part I: The interaction of C4 olefins with pure γ-alumina. J. Catal. 1997, 168, 334–348. [Google Scholar] [CrossRef]
  165. Vang, R.T.; Honkala, H.; Dahl, S.; Vestergaard, E.K.; Schnadt, J.; Lægsgaard, E.; Clausen, B.S.; Nørskov, J.K.; Besenbacher, F. Ethylene dissociation on flat and stepped Ni(1 1 1): A combined STM and DFT study. Surf. Sci. 2006, 600, 66–77. [Google Scholar] [CrossRef]
  166. Yilmazer, N.D.; Fellaha, M.F.; Onal, I. A density functional theory study of ethylene adsorption on Ni10(111), Ni13(100) and Ni10(110) surface cluster models and Ni13 nanocluster. Appl. Surf. Sci. 2010, 256, 5088–5093. [Google Scholar] [CrossRef]
  167. Medlin, J.W.; Allendorf, M.D. Theoretical Study of the Adsorption of Acetylene on the (111) Surfaces of Pd, Pt, Ni, and Rh. J. Phys. Chem. B 2003, 107, 217–223. [Google Scholar] [CrossRef]
  168. Jenkins, S.J. Aromatic adsorption on metals via first-principles density functional theory. Proc. R. Soc. A 2009, 465, 2949–2976. [Google Scholar] [CrossRef]
  169. Lobo, L.S.; Trimm, D.L. Carbon formation from light hydrocarbons on nickel. J. Catal. 1973, 29, 15–19. [Google Scholar] [CrossRef]
  170. Kodde, A.J.; Padin, J.; van der Meer, P.J.; Mittelmeijer-Hazeleger, M.C.; Bliek, A.; Yang, R.T. NiCl2 on γ-Alumina as Selective Adsorbents for Acetylene over Ethylene. Ind. Eng. Chem. Res. 2000, 39, 3108–3111. [Google Scholar] [CrossRef]
  171. Shimura, K.; Yoshida, S.; Oikawa, H.; Fujitani, T. Impacts of Ni-Loading Method on the Structure and the Catalytic Activity of NiO/SiO2-Al2O3 for Ethylene Oligomerization. Catalysts 2023, 13, 1303. [Google Scholar] [CrossRef]
  172. Koninckx, E.; Mendes, P.S.F.; Thybaut, J.W.; Broadbelt, L.J. Ethylene oligomerization on nickel catalysts on a solid acid support: From New mechanistic insights to tunable bifunctionality. Appl. Catal. A Gen. 2021, 624, 118296. [Google Scholar] [CrossRef]
  173. Horiuchi, T.; Hidaka, H.; Fukui, T.; Kubo, Y.; Horio, M.; Suzuki, K.; Mori, T. Effect of added basic metal oxides on CO2 adsorption on alumina at elevated temperatures. Appl. Catal. A Gen. 1998, 167, 195–202. [Google Scholar] [CrossRef]
  174. Garbarino, G.; Wang, C.; Valsamakis, I.; Chitsazan, S.; Riani, P.; Finocchio, E.; Flytzani-Stephanopoulos, M.; Busca, G. Acido-basicity of lanthana/alumina catalysts and their activity in ethanol conversion. Appl. Catal. B Environ. 2017, 200, 458–468. [Google Scholar] [CrossRef]
  175. Garbarino, G.; Riani, P.; García, V.; Finocchio, E.; Sánchez Escribano, V.; Busca, G. A study of ethanol conversion over zinc aluminate catalyst. Reac. Kinet. Mech. Cat. 2018, 124, 503–522. [Google Scholar] [CrossRef]
  176. Nobakht, A.R.; Rezaei, M.; Alavi, S.M.; Akbari, E.; Varbar, M.; Hafezi-Bakhtiari, J. CO2 methanation over NiO catalysts supported on CaO–Al2O3: Effect of CaO:Al2O3 molar ratio and nickel loading. Int. J. Hydrogen Energy 2023, 48, 38664–38675. [Google Scholar] [CrossRef]
  177. Solymosi, F. The bonding, structure and reactions of CO2 adsorbed on clean and promoted metal surfaces. J. Mol. Catal. 1991, 65, 337–358. [Google Scholar] [CrossRef]
  178. Freund, H.-J.; Roberts, M.W. Surface chemistry of carbon dioxide. Surf. Sci. Rep. 1996, 25, 225–273. [Google Scholar] [CrossRef]
  179. Taifan, W.; Boily, J.F.; Baltrusaitis, J. Surface chemistry of carbon dioxide revisited. Surf. Sci. Rep. 2016, 71, 595–671. [Google Scholar] [CrossRef]
  180. Liu, C.; Cundari, T.R.; Wilson, A.K. CO2 Reduction on Transition Metal (Fe, Co, Ni, and Cu) Surfaces: In Comparison with Homogeneous Catalysis. J. Phys. Chem. C 2012, 116, 5681–5688. [Google Scholar] [CrossRef]
  181. Liu, X.; Sun, L.; Deng, W.Q. Theoretical Investigation of CO2 Adsorption and Dissociation on Low Index Surfaces of Transition Metals. J. Phys. Chem. C 2018, 122, 8306–8314. [Google Scholar] [CrossRef]
  182. Jin, W.; Wang, Y.; Liu, T.; Ding, C.; Guo, H. CO2 chemisorption and dissociation on flat and stepped transition metal surfaces. Appl. Surf. Sci. 2022, 599, 154024. [Google Scholar] [CrossRef]
  183. Cai, J.; Han, Y.; Chen, S.; Crumlin, E.J.; Yang, B.; Li, Y.; Liu, Z. CO2 Activation on Ni(111) and Ni(100) Surfaces in the Presence of H2O: An Ambient-Pressure X-ray Photoelectron Spectroscopy Study. J. Phys. Chem. C 2019, 123, 12176–12182. [Google Scholar] [CrossRef]
  184. Schreiter, N.; Kirchner, J.; Kureti, S. A DRIFTS and TPD study on the methanation of CO2 on Ni/Al2O3 catalyst. Catal. Commun. 2020, 140, 105988. [Google Scholar] [CrossRef]
  185. Busca, G.; Spennati, E.; Riani, P.; Garbarino, G. Looking for an Optimal Composition of Nickel-Based Catalysts for CO2 Methanation. Energies 2023, 16, 5304. [Google Scholar] [CrossRef]
  186. Antill, J.E.; Warburton, J.B. Oxidation of Nickel by Carbon Dioxide. J. Electrochem. Soc. 1967, 114, 1215–1221. [Google Scholar] [CrossRef]
  187. Napari, M.; Huq, T.N.; Hoye, R.L.Z.; MacManus-Driscoll, J.L. Nickel oxide thin films grown by chemical depositiontechniques: Potential and challenges in next-generationrigid and flexible device applications. InfoMat 2021, 3, 536–576. [Google Scholar] [CrossRef]
  188. Horn, R.; Schlögl, R. Methane Activation by Heterogeneous Catalysis. Catal. Lett. 2015, 145, 23–39. [Google Scholar] [CrossRef]
  189. Liu, F.; Sang, Y.; Ma, H.; Li, Z.; Gao, Z. Nickel oxide as an effective catalyst for catalytic combustion of methane. J. Nat. Gas Sci. Eng. 2017, 41, 1–6. [Google Scholar] [CrossRef]
  190. Lee, M.B.; Yang, Q.Y.; Ceyer, S.T. Dynamics of the Activated Dissociative Chemisorption of CH4 and Implication for the Pressure Gap in Catalysis—A Molecular-Beam High-Resolution Electron-Energy Loss Study. J. Chem. Phys. 1987, 87, 2724–2741. [Google Scholar] [CrossRef]
  191. Kubokawa, M. The activated adsorption of methane on reduced nickel. Proc. Imp. Acad. 1938, 14, 61–66. [Google Scholar] [CrossRef]
  192. García-Sancho, C.; Guil-Lopez, R.; Sebastian-Lopez, A.; Navarro, R.M.; Fierro, J.L.G. Hydrogen production by methane decomposition: A comparative study of supported and bulk ex-hydrotalcite mixed oxide catalysts with Ni, Mg and Al. Int. J. Hydrogen Energy 2018, 43, 9607–9621. [Google Scholar] [CrossRef]
  193. Gubanov, M.A.; Ivantsov, M.I.; Kulikova, M.V.; Kryuchkov, V.A.; Nikitchenko, N.V.; Knyazeva, M.I.; Kulikova, A.B.; Pimenov, A.A.; Maksimova, A.L. Methane Decomposition Nickel Catalysts Based on Structured Supports. Petrol. Chem. 2020, 60, 1043–1051. [Google Scholar] [CrossRef]
  194. Resasco, D.E.; Marcus, B.K.; Huang, C.S.; Durante, V.A. Isobutane dehydrogenation over sulfided nickle catalysts. J. Catal. 1994, 146, 40–55. [Google Scholar] [CrossRef]
  195. Tsai, M.C.; Friend, C.M.; Muetterties, E.L. Dehydrogenation processes on nickel and platinum surfaces. Conversion of cyclohexane, cyclohexene, and cyclohexadiene to benzene. J. Am. Chem. Soc. 1982, 104, 2539–2543. [Google Scholar] [CrossRef]
  196. Ma, R.; Gao, J.; Kou, J.; Dean, D.P.; Breckner, C.J.; Liang, K.; Zhou, B.; Miller, J.T.; Zou, G. Insights into the Nature of Selective Nickel Sites on Ni/Al2O3 Catalysts for Propane Dehydrogenation. ACS Catal. 2022, 12, 12607–12616. [Google Scholar] [CrossRef]
  197. Diskin, A.M.; Cunningham, R.H.; Ormerod, R.M. The oxidative chemistry of methane over supported nickel catalyst. Catal. Today 1998, 46, 147–154. [Google Scholar] [CrossRef]
  198. Sehested, J.; Gelten, J.A.P.; Remediakis, J.N.; Bengaard, H.; Nørskov, J.K. Sintering of nickel steam-reforming catalysts: Effects of temperature and steam and hydrogen pressures. J. Catal. 2004, 223, 432–443. [Google Scholar] [CrossRef]
  199. Clavenna, L.R.; Davis, S.M.; Beasley, B.E. Process for the Reactivation of Nickel-Alumina Catalysts. U.S. Patent 5356845, 18 October 1994. [Google Scholar]
  200. Hashemnejad, S.M.; Parvari, M. Deactivation and Regeneration of Nickel-Based Catalysts for Steam-Methane Reforming. Chin. J. Catal. 2011, 32, 273–279. [Google Scholar] [CrossRef]
  201. Dega, F.B.; Abatzoglou, N. H2S Poisoning and Regeneration of a Nickel Spinellized Catalyst Prepared from Waste Metallurgical Residues, During Dry Autothermal Methane Reforming. Catal. Lett. 2019, 149, 1730–1742. [Google Scholar] [CrossRef]
  202. Rostrup-Nielsen, J.R. Some principles relating to the regeneration of sulfur-poisoned nickel catalyst. J. Catal. 1971, 21, 171–178. [Google Scholar] [CrossRef]
  203. Li, L.; Howard, C.J.; King, D.L.; Gerber, M.A.; Dagle, R.A.; Stevens, B.J. Regeneration of Sulfur Deactivated Ni-Based Biomass Syngas Cleaning Catalysts. Ind. Eng. Chem. Res. 2010, 49, 10144–10148. [Google Scholar] [CrossRef]
  204. Wachter, P.; Gaber, C.; Raic, J.; Demuth, M.; Hochenauer, C. Experimental investigation on H2S and SO2 sulphur poisoning and regeneration of a commercially available Ni-catalyst during methane tri-reforming. Int. J. Hydrogen Energy 2021, 46, 3437–3452. [Google Scholar] [CrossRef]
  205. Vogt, E.T.C.; Fu, D.; Weckhuysen, B.M. Carbon Deposit Analysis in Catalyst Deactivation, Regeneration, and Rejuvenation. Angew. Chem. Int. Ed. 2023, 62, e202300319. [Google Scholar] [CrossRef]
  206. Uemura, Y.; Hatate, Y. Effect of Nickel Concentration Profile on Selectivity of Acetylene Hydrogenation. J. Chem. Eng. Jpn. 1989, 22, 287–291. [Google Scholar] [CrossRef]
  207. Ryu, J.Y. Ni Catalyst, Process for Making Catalysts and Selective Hydrogenation Process. EP1786747A2, 14 February 2005. [Google Scholar]
  208. Available online: https://www.clariant.com/en/Solutions/Products/2014/06/16/09/13/OleMax-100-Series (accessed on 26 July 2024).
  209. Krupa, S.; Foley, T.; McColl, S. UOP KLP 1.3-butadiene from acetylene process. In Handbook of Petrochemicals Production Processes; Meyers, R.A., Ed.; McGraw-Hill: New York, NY, USA, 2005; pp. 3.11–3.14. [Google Scholar]
  210. Bosman, H.J.M.; Grootendorst, E.J.; Postma, L.H.; Smeets, T.M. Process for the Hydrogenation of Phenyl Acetylene in a Styrene-containing Medium with the Aid of a Catalyst. U.S. Patent 6747181B1, 8 June 2004. [Google Scholar]
  211. Rajeshwer, D.; Sreenivasa Rao, G.; Krishnamurthy, K.R.; Padmavathi, G.; Subrahmanyam, N.; Rachh, J.D. Kinetics of Liquid—Phase Hydrogenation of Straight Chain C10 to C13 Di-Olefins Over Ni/Al2O3 Catalyst. Int. J. Chem. React. Eng. 2006, 4, A17. [Google Scholar] [CrossRef]
  212. Kravtsov, A.V.; Zuev, V.A.; Kozlov, I.A.; Milishnikov, A.V.; Ivanchina, E.D.; Uriev, E.M.; Ivashkina, E.N.; Fetisova, V.A.; Shnidorova, I.O. Development of control system for nickel-containing catalyst in dienes hydrogenation. Pet. Coal 2009, 51, 248–254. [Google Scholar]
  213. Hoffer, W.; Bonné, R.L.C.; van Langeveld, A.D.; Griffiths, C.; Lok, C.M.; Moulijn, J.A. Enhancing the start-up of pyrolysis gasoline hydrogenation reactors by applying tailored ex situ presulfided Ni/Al2O3 catalysts. Fuel 2004, 83, 1–8. [Google Scholar] [CrossRef]
  214. Available online: https://matthey.com/products-and-markets/chemicals/pyrolysis-gasoline-catalysts (accessed on 26 July 2024).
  215. Available online: https://www.topsoe.com/our-resources/knowledge/our-products/catalysts/sh-501 (accessed on 26 July 2024).
  216. Available online: https://www.axens.net/solutions/catalysts-adsorbents-grading-supply/selective-hydrogenation-catalysts (accessed on 26 July 2024).
  217. Fremerey, P.; Jess, A.; Moos, R. Why does the Conductivity of a Nickel Catalyst Increase during Sulfidation? An Exemplary Study Using an In Operando Sensor Device. Sensors 2015, 15, 27021–27034. [Google Scholar] [CrossRef] [PubMed]
  218. Available online: https://matthey.com/products-and-markets/chemicals/de-aromatisation-catalysts (accessed on 26 July 2024).
  219. Available online: https://matthey.com/products-and-markets/chemicals/solvent-purification (accessed on 26 July 2024).
  220. Watson, M.J.; Partridge, M.G.; Appleton, P.; Jackson, S. Chemical Process and Catalyst. WO2010018405A1, 18 February 2010. [Google Scholar]
  221. Aasberg-Petersen, K.; Dybkjær, I.; Ovesen, C.V.; Schjødt, N.C.; Sehested, J.; Thomsen, S.G. Natural gas to synthesis gas—Catalysts and catalytic processes. J. Nat. Gas Sci. Eng. 2011, 3, 423–459. [Google Scholar] [CrossRef]
  222. Pearce, B.B.; Twigg, M.V.; Woodward, C. Catalyst Handbook, 2nd ed.; Twigg, M.V., Ed.; Wolfe Pub.: London, UK, 1989; pp. 340–383. [Google Scholar]
  223. Golosman, E.Z.; Efremov, V.N. Industrial Catalysts for the Hydrogenation of Carbon Oxides. Catal. Ind. 2012, 4, 267–283. [Google Scholar] [CrossRef]
  224. Miguel, C.V.; Mendes, A.; Madeira, L.M. Intrinsic kinetics of CO2 methanation over an industrial nickel-based catalyst. J. CO₂ Util. 2018, 25, 128–136. [Google Scholar] [CrossRef]
  225. Available online: https://www.topsoe.com/our-resources/knowledge/our-products/catalysts/pk-7r (accessed on 27 July 2024).
  226. Available online: https://matthey.com/en/products-and-markets/chemicals/methanation-catalysts (accessed on 27 July 2024).
  227. Kopyscinski, J.; Schildhauer, T.J.; Biollaz, S.M.A. Production of synthetic natural gas (SNG) from coal and dry biomass e a technology review from 1950 to 2009. Fuel 2010, 89, 1763–1783. [Google Scholar] [CrossRef]
  228. Røstrup-Nielsen, J.R.; Pedersen, K.; Sehested, J. High temperature methanation. Sintering and structure sensitivity. Appl Catal. A Gen. 2007, 330, 134–138. [Google Scholar] [CrossRef]
  229. Li, J.; Tian, Y.; Yan, X.; Yang, J.; Wang, Y.; Xu, W.; Xie, K. Approach and potential of replacing oil and natural gas with coal in China. Front. Energy 2020, 14, 419–431. [Google Scholar] [CrossRef]
  230. Available online: https://www.topsoe.com/our-resources/knowledge/our-products/catalysts/mcr-2 (accessed on 27 July 2024).
  231. Torcida, M.F.; Curto, D.; Martín, M. Design and optimization of CO2 hydrogenation multibed reactors. Chem. Eng. Res. Des. 2022, 181, 89–100. [Google Scholar] [CrossRef]
  232. Dagle, R.A.; Wang, Y.; Xia, G.G.; Strohm, J.J.; Holladay, J.; Palo, D.R. Selective CO methanation catalysts for fuel processing applications. Appl. Catal. A Gen. 2007, 326, 213–218. [Google Scholar] [CrossRef]
  233. Nguyen, T.T.M.; Wissing, L.; Skjøth-Rasmussen, M.S. High temperature methanation: Catalyst considerations. Catal. Today 2013, 215, 233–238. [Google Scholar] [CrossRef]
  234. Ross, J.R.H. Nickel Catalysts for C1 Reactions: Recollections from a Career in Heterogeneous Catalysis. Top. Catal. 2021, 64, 896–906. [Google Scholar] [CrossRef]
  235. Available online: https://www.clariant.com/en/Corporate/News/2016/07/VESTA-Oncethrough-Methanation-New-Technology-with-Wide-H2CO-Flexibility-Successfully-Passes-Pilot-Te (accessed on 27 July 2024).
  236. Romano, L.; Ruggeri, F. Methane from syngas—Status of Amec Foster Wheeler VESTA technology development. Energy Procedia 2015, 81, 249–254. [Google Scholar] [CrossRef]
  237. Götz, M.; Lefebvre, J.; Mörs, F.; Koch, A.M.; Graf, F.; Bajohr, S.; Reimert, R.; Kolb, R. Renewable Power-to-Gas: A Technological and Economic Review. Renew. Energy 2016, 85, 1371–1390. [Google Scholar] [CrossRef]
  238. Vega Puga, E.; Moumin, G.; Neumann, N.C.; Roeb, M.; Ardone, A.; Sattler, C. Holistic View on Synthetic Natural Gas Production: A Technical, Economic and Environmental Analysis. Energies 2022, 15, 1608. [Google Scholar] [CrossRef]
  239. Thema, M.; Bauer, F.; Sterner, M. Sterner Power-to-Gas: Electrolysis and methanation status review. Renew. Sustain. Energy Rev. 2019, 112, 775–778. [Google Scholar] [CrossRef]
  240. Pintér, G. The development of global power-to-methane potentials between 2000 and 2020: A comparative overview of international projects. Appl. Energy 2024, 353, 122094. [Google Scholar] [CrossRef]
  241. Tripodi, A.; Conte, F.; Rossetti, I. Carbon Dioxide Methanation: Design of a Fully Integrated Plant. Energy Fuels 2020, 34, 7242–7256. [Google Scholar] [CrossRef]
  242. Uddin, Z.; Yu, B.Y.; Lee, H.Y. Evaluation of alternative processes of CO2 methanation: Design, optimization, control, techno-economic and environmental analysis. J. CO2 Util. 2022, 60, 101974. [Google Scholar] [CrossRef]
  243. Available online: https://www.clariant.com/en/Business-Units/Catalysts/Energy-Transition/Carbon-Capture-and-Utilization (accessed on 27 July 2024).
  244. Mohd Ridzuan, N.D.; Shaharun, M.S.; Anawar, M.A.; Ud-Din, I. Ni-Based Catalyst for Carbon Dioxide Methanation: A Review on Performance and Progress. Catalysts 2022, 12, 469. [Google Scholar] [CrossRef]
  245. Hanh My Bui, H.M.; Großmann, P.F.; Gros, T.; Blum, M.; Berger, A.; Fischer, R.; Szesni, N.; Tonigold, M.; Hinrichsen, O. 3D printed co-precipitated Ni-Al CO2 methanation catalysts by Binder Jetting: Fabrication, characterization and test in a single pellet string reactor. Appl. Catal. A Gen. 2022, 643, 118760. [Google Scholar]
  246. Bown, R.M.; Joyce, M.; Zhang, Q.; Ramirez Reina, T.; Duya, M.S. Identifying Commercial Opportunities for the Reverse Water Gas Shift Reaction. Energy Technol. 2021, 9, 2100554. [Google Scholar] [CrossRef]
  247. Thor Wismann, S.; Larsen, K.E.; Mølgaard Mortensen, P. Electrical Reverse Shift: Sustainable CO2 Valorization for Industrial Scale. Angew. Chem. Int. Ed. 2022, 61, e202109696. [Google Scholar] [CrossRef] [PubMed]
  248. Available online: https://www.hydrocarbonengineering.com/clean-fuels/26052022/ineratec-and-clariant-join-forces-for-a-cleaner-future/ (accessed on 30 October 2023).
  249. Available online: https://matthey.com/en/news/2022/hycogen (accessed on 30 October 2023).
  250. Available online: https://www.hydrocarbonprocessing.com/news/2023/08/e-fuels-axens-paul-wurth-ifpen-agree-to-co-develop-reverse-water-gas-shift-tech/ (accessed on 30 October 2023).
  251. Gomez, S.; Peters, J.A.; Maschmeyer, T. The Reductive Amination of Aldehydes and Ketones and the Hydrogenation of Nitriles: Mechanistic Aspects and Selectivity Control. Adv. Synth. Catal. 2002, 344, 1037–1057. [Google Scholar] [CrossRef]
  252. Medina, F.; Dutartre, R.; Tichit, D.; Coq, B.; Dung, N.T.; Salagre, P.; Sueiras, J.E. Characterization and activity of hydrotalcite-type catalysts for acetonitrile hydrogenation. J. Mol. Catal. A Chem. 1997, 119, 201–212. [Google Scholar] [CrossRef]
  253. Hahn, G.; Kunnas, P.; de Jonge, N.; Kempe, R. General synthesis of primary amines via reductive amination employing a reusable nickel catalyst. Nat. Catal. 2019, 2, 71–77. [Google Scholar] [CrossRef]
  254. McCullagh, A.M.; Brennan, C.; Davidson, A.L.; Lennon, D.; Ballas, C.E.; How, C. The application of an alumina-supported Ni catalyst for the hydrogenation of nitrobenzene to aniline. Catal. Today 2024, 442, 114933. [Google Scholar] [CrossRef]
  255. Chen, B.; Dingerdissen, U.; Krauter, J.G.E.; Lansink Rotgerink, H.G.J.; Möbus, K.; Ostgard, D.J.; Panster, P.; Riermeier, T.H.; Seebald, S.; Tack, T.; et al. New developments in hydrogenation catalysis particularly in synthesis of fine and intermediate chemicals. Appl. Catal. A Gen. 2005, 280, 17–46. [Google Scholar] [CrossRef]
  256. Farahani, M.D.; Valand, J.; Mahomed, A.S.; Friedrich, H.B. A Comparative Study of NiO/Al2O3 Catalysts Prepared by Different Combustion Techniques for Octanal Hydrogenation. Catal. Lett. 2016, 146, 2441–2449. [Google Scholar] [CrossRef]
  257. Perret, N.; Grigoropoulos, A.; Zanella, M.; Manning, T.D.; Claridge, J.B.; Rosseinsky, M.J. Catalytic Response and Stability of Nickel/Alumina for the Hydrogenation of 5-Hydroxymethylfurfural in Water. ChemSusChem 2016, 9, 521–531. [Google Scholar] [CrossRef]
  258. Li, J.; Qian, L.P.; Hu, L.Y.; Yue, B.; He, H.Y. Low-temperature hydrogenation of maleic anhydride to succinic anhydride and γ-butyrolactone over pseudo-boehmite derived alumina supported metal (metal = Cu, Co and Ni) catalysts. Chin. Chem. Lett. 2016, 27, 1004–1008. [Google Scholar] [CrossRef]
  259. Yang, R.; Li, X.; Wu, J.; Zhang, X.; Zhang, Z.; Cheng, Y.; Guo, J. Hydrotreating of crude 2-ethylhexanol over Ni/Al2O3 catalysts: Surface Ni species-catalytic activity correlation. Appl. Catal. A Gen. 2009, 368, 105–112. [Google Scholar] [CrossRef]
  260. Smejkal, Q.; Smejkalová, L.; Kubicka, D. Thermodynamic balance in reaction system of total vegetable oil hydrogenation. Chem. Eng. J. 2009, 146, 155–160. [Google Scholar] [CrossRef]
  261. Gousi, M.; Andriopoulou, C.; Bourikas, K.; Ladas, S.; Sotiriou, M.; Kordulis, C.; Lycourghiotis, A. Green diesel production over nickel-alumina co-precipitated catalysts. Appl. Catal. A Gen. 2017, 536, 45–56. [Google Scholar] [CrossRef]
  262. Morgan, T.; Santillan-Jimenez, E.; Harman-Ware, A.E.; Ji, Y.; Grubb, D.; Crocker, M. Catalytic deoxygenation of triglycerides to hydrocarbons over supported nickel catalysts. Chem. Eng. J. 2012, 189–190, 346–355. [Google Scholar] [CrossRef]
  263. Shi, F.; Wang, H.; Chen, Y.; Zheng, Z.; Zheng, Y. Green diesel-like hydrocarbon production by H2-free catalytic deoxygenation of oleic acid via Ni/MgO-Al2O3 catalysts: Effect of the metal loading amount. J. Environ. Chem. Eng. 2023, 11, 110520. [Google Scholar] [CrossRef]
  264. Chen, S.; Zhou, G.; Xie, H.; Jiao, Z.; Zhang, X. Hydrodeoxygenation of methyl laurate over the sulfur-free Ni/γ-Al2O3 catalysts. Appl. Catal. A Gen. 2019, 569, 35–44. [Google Scholar] [CrossRef]
  265. Jiang, W.; Cao, J.P.; Huan, Z.X.; Hu, X.; Tang, W.; Chen, C.X.; Wang, H.Y.; He, Z.M.; Zhao, X.Y.; Bai, H.C. Relatively Electron-Rich Ni Nanoparticles Supported on α-Al2O3 for High-Efficiency Hydrogenolysis of Lignin and Its Derivatives under Mild Conditions. ACS Sustain. Chem. Eng. 2023, 11, 17646–17661. [Google Scholar] [CrossRef]
  266. Hou, M.; Chen, H.; Li, Y.; Wang, H.; Zhang, L.; Bi, Y. Reductive Catalytic Fractionation of Lignocellulose over Ni/Al2O3 Catalyst Prepared by an EDTA-Assisted Impregnation Method. Energy Fuels 2022, 36, 1929–1938. [Google Scholar] [CrossRef]
  267. Van den Bosch, S.; Renders, T.; Kennis, S.; Koelewijn, S.-F.; Van den Bossche, G.; Vangeel, T.; Deneyer, A.; Depuydt, D.; Courtin, C.M.; Thevelein, J.M.; et al. Integrating lignin valorization and bio-ethanol production: On the role of Ni-Al2O3 catalyst pellets during lignin-first fractionation. Green Chem. 2017, 19, 3313–3326. [Google Scholar] [CrossRef]
  268. Brandon, C.; Vance, B.C.; Najmi, S.; Kots, P.A.; Wang, C.; Jeon, S.; Stach, E.A.; Zakharov, D.N.; Marinkovic, N.; Ehrlich, S.N.; et al. Structure–Property Relationships for Nickel Aluminate Catalysts in Polyethylene Hydrogenolysis with Low Methane Selectivity. JACS Au 2023, 3, 2156–2165. [Google Scholar]
  269. Alshareef, R.; Nahil, M.A.; Williams, P.T. Hydrogen Production by Three-Stage (i) Pyrolysis, (ii) Catalytic Steam Reforming, and (iii) Water Gas Shift Processing of Waste Plastic. Energy Fuels 2023, 37, 3894–3907. [Google Scholar] [CrossRef]
  270. Aminu, I.; Nahil, M.A.; Williams, P.T. High-yield hydrogen from thermal processing of waste plastics. Proc. Inst. Civil Eng.—Waste Res. Manag. 2022, 175, 3–13. [Google Scholar] [CrossRef]
  271. Røstrup-Nielsen, J.R.; Sehested, J.; Nørskov, J.K. Hydrogen and Synthesis Gas by Steam and CO2 Reforming. Adv. Catal. 2002, 47, 65–139. [Google Scholar] [CrossRef]
  272. Fowles, M.; Carlsson, M. Steam Reforming of Hydrocarbons for Synthesis Gas Production. Top. Catal. 2021, 64, 856–875. [Google Scholar] [CrossRef]
  273. Zhang, H.; Sun, Z.; Hu, Y.H. Steam reforming of methane: Current states of catalyst design and process upgrading. Renew. Sustain. Energy Rev. 2021, 149, 111330. [Google Scholar] [CrossRef]
  274. Garbarino, G.; Pugliese, F.; Cavattoni, T.; Busca, G.; Costamagna, P. A Study on CO2 Methanation and Steam Methane Reforming over Commercial Ni/Calcium Aluminate Catalysts. Energies 2020, 13, 2792. [Google Scholar] [CrossRef]
  275. Røstrup-Nielsen, J.R. Catalytic Steam Reforming. In Catalysis Science and Technology; Anderson, J.R., Boudart, M., Eds.; Springer: Berlin, Germany, 1984; Volume 5, pp. 1–118. [Google Scholar]
  276. Carlsson, M. Carbon Formation in Steam Reforming and Effect of Potassium Promotion. Johns. Matthey Technol. Rev. 2015, 59, 313–318. [Google Scholar] [CrossRef]
  277. Goula, M.A.; Charisiou, N.D.; Papageridis, K.N.; Delimitis, A.; Pachatouridou, E.; Iliopoulou, E.F. Nickel on alumina catalysts for the production of hydrogen rich mixtures via the biogas dry reforming reaction: Influence of the synthesis method. Int. J. Hydrogen Energy 2015, 40, 9183–9200. [Google Scholar] [CrossRef]
  278. Rashid, M.U.; Wan Daud, W.M.A.; Abbas, H.F. Dry reforming of methane: Influence of process parameters—A review. Renew. Sustain. En. Rev. 2015, 45, 710–744. [Google Scholar]
  279. Aramouni, N.A.K.; Touma, J.G.; Tarboush, B.A.; Zeaiter, J.; Ahmad, M.N. Catalyst design for dry reforming of methane: Analysis review. Renew. Sustain. Energy Rev. 2018, 82, 2570–2585. [Google Scholar] [CrossRef]
  280. Torrez-Herrera, J.J.; Korili, S.A.; Gil, A. Recent progress in the application of Ni-based catalysts for the dry reforming of methane. Catal. Rev. 2021, 65, 1300–1357. [Google Scholar] [CrossRef]
  281. Qin, Z.; Chen, J.; Xie, X.; Luo, X.; Su, T.; Ji, H. CO2 reforming of CH4 to syngas over nickel-based catalysts. Environ. Chem. Lett. 2020, 18, 997–1017. [Google Scholar] [CrossRef]
  282. Liu, Z.W.; Jun, K.W.; Roh, H.S.; Park, S.E.; Oh, Y.S. Partial Oxidation of Methane over Nickel Catalysts Supported on Various Aluminas. Korean J. Chem. Eng. 2002, 19, 735–741. [Google Scholar] [CrossRef]
  283. Yang, H.; Zhang, Z.; He, Y.; Tang, Y.; Tan, L. Study of Ni Catalysts with Different Al2O3 Supports in the Partial Oxidation of Methane Reaction. Ind. Eng. Chem. Res. 2023, 62, 12120–12132. [Google Scholar] [CrossRef]
  284. Raza, J.; Khoja, A.H.; Anwar, M.; Saleem, F.; Naqvi, S.R.; Liaquat, R.; Hassan, M.; Javaid, R.; Qazi, U.Y.; Lumbers, B. Methane decomposition for hydrogen production: A comprehensive review on catalyst selection and reactor systems. Renew. Sustain. Energy Rev. 2022, 168, 112774. [Google Scholar] [CrossRef]
  285. Biondo, L.D.; Manera, C.; Aguzzoli, C.; Godinho, M. DoE-driven thermodynamic assessment of COX-free hydrogen production from methane decomposition. Catal. Commun. 2024, 187, 106874. [Google Scholar] [CrossRef]
  286. Salmones, J.; Wang, J.A.; Valenzuela, M.A.; Sánchez, E.; Garcia, A. Pore geometry influence on the deactivation behavior of Ni-based catalysts for simultaneous production of hydrogen and nanocarbon. Catal. Today 2009, 148, 134–139. [Google Scholar] [CrossRef]
  287. Li, Y.; Chen, J.; Chang, L. Catalytic growth of carbon fibers from methane on a nickel-alumina composite catalyst prepared from Feitknecht compound precursor. Appl. Catal. A Gen. 1997, 163, 45–57. [Google Scholar] [CrossRef]
  288. Kovalenko, G.A.; Rudina, N.A.; Perminova, L.V.; Skrypnik, O.V. Corundum Impregnation Conditions for Preparing Supported Ni Catalysts for the Synthesis of a Uniform Layer of Carbon Nanofibers. Kinet. Catal. 2010, 51, 762–770. [Google Scholar] [CrossRef]
  289. Wan, Z.; Tao, Y.; Shao, J.; Zhang, Y.; You, H. Ammonia as an effective hydrogen carrier and a clean fuel for solid oxide fuel cells. Energy Convers. Manag. 2021, 228, 113729. [Google Scholar] [CrossRef]
  290. Yu, X.; Yin, F.; Li, G.; Zhang, J.; Chen, B. Preparation of nickel aluminate supported Ni nanocatalyst and its catalytic activity for ammonia decomposition to produce hydrogen. Int. J. Hydrogen Energy, 2024; in press. [Google Scholar] [CrossRef]
  291. Ashcroft, J.; Goddin, H. Centralised and Localised Hydrogen Generation by Ammonia Decomposition. A technical review of the ammonia cracking process. Johns. Matthey Technol. Rev. 2022, 66, 375–385. [Google Scholar] [CrossRef]
  292. Available online: https://matthey.com/products-and-markets/chemicals/ammonia-cracking-catalysts (accessed on 26 July 2024).
  293. Angelette, L.M.; Belliveau, R.G.; Coopersmith, K.J.; Cooper, J.J.; Steedley, J.A.; Morrell, B.B. Evaluation of catalysts for decomposition of ammonia in hydrogen isotope purification systems. Fusion Eng. Des. 2021, 173, 112895. [Google Scholar] [CrossRef]
  294. Sansaniwal, S.K.; Pal, K.; Rosen, M.A.; Tyagi, S.K. Recent advances in the development of biomass gasification technology: A comprehensive review. Renew. Sustain. Energy Rev. 2017, 72, 363–384. [Google Scholar] [CrossRef]
  295. Cortazar, M.; Santamaria, L.; Lopez, G.; Alvarez, J.; Zhang, L.; Wang, R.; Bi, X.; Olazar, M. A comprehensive review of primary strategies for tar removal in biomass gasification. Energy Conv. Manag. 2023, 276, 116496. [Google Scholar] [CrossRef]
  296. Lorente, E.; Millan, M.; Brandon, N.P. Use of gasification syngas in SOFC: Impact of real tar on anode materials. Int. J. Hydrogen Energy 2012, 37, 7271–7278. [Google Scholar] [CrossRef]
  297. Røstrup-Nielsen, J.R.; Bøgild Hansen, J. Steam Reforming for Fuel Cells. In Fuel Cells, Technologies for Fuel Processing; Shekhawat, D., Spivey, J.J., Berry, D.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2011; Chapter 4; pp. 49–71. [Google Scholar]
  298. Andersson, K.J.; Rasmussen, M.S.S.; Højlund Nielsen, P.E. Industrial-scale gas conditioning including Topsoe tar reforming and purification downstream biomass gasifiers: An overview and recent examples. Fuel 2017, 203, 1026–1030. [Google Scholar] [CrossRef]
  299. Chitsazanm, S.; Sepehri, S.; Garbarino, G.; Carnasciali, M.M.; Busca, G. Steam reforming of biomass-derived organics: Interactions of different mixture components on Ni/Al2O3 based catalysts. Appl. Catal. B Environ. 2016, 187, 386–398. [Google Scholar] [CrossRef]
  300. Tang, W.; Cao, J.P.; He, Z.M.; Jiang, W.J.; Wang, Z.H.; Zhao, X.Y. Recent Progress of Catalysts for Reforming of BiomassTar/Tar Models at Low Temperatures—A Short Review. ChemCatChem 2023, 15, e202300581. [Google Scholar] [CrossRef]
  301. Phung, T.K.; Busca, G. Selective Bioethanol Conversion to Chemicals and Fuels via Advanced Catalytic Approaches. In Biorefinery of Alternative Resources: Targeting Green Fuels and Platform Chemicals; Nanda, S., Vo, D.-V.N., Sarangi, P.K., Eds.; Springer: Berlin/Heidelberg, Germany, 2020; pp. 75–103. [Google Scholar]
  302. Chen, W.H.; Biswas, P.P.; Hwai Chyuan Ong, H.C.; Hoang, A.T.; Nguyen, T.B.; Dong, C.D. A critical and systematic review of sustainable hydrogen production from ethanol/bioethanol: Steam reforming, partial oxidation, and autothermal reforming. Fuel 2023, 333, 126526. [Google Scholar] [CrossRef]
  303. Alipour, H.; Alavi, S.M.; Rezaei, M.; Akbari, E.; Varbar, M. The role of various preparation techniques and nickel loadings in ethanol steam reforming over mesoporous nanostructured Ni–Al2O3 catalysts. J. Energy Inst. 2024, 112, 101488. [Google Scholar] [CrossRef]
  304. Garbarino, G.; Cavattoni, T.; Riani, P.; Brescia, R.; Canepa, F.; Busca, G. On the Role of Support in Metallic Heterogeneous Catalysis: A Study of Unsupported Nickel–Cobalt Alloy Nanoparticles in Ethanol Steam Reforming. Catal. Lett. 2019, 149, 929–941. [Google Scholar] [CrossRef]
  305. Available online: https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attachment_data/file/866383/Phase_1_-_Wood_-_Novel_Renewable_Ethanol_Steam_Reformer.pdf (accessed on 14 May 2021).
  306. Oakley, J.H.; Hoadley, A.F.A. Industrial scale steam reforming of bioethanol: A conceptual study. Int. J. Hydrogen Energy 2010, 35, 8472–8485. [Google Scholar] [CrossRef]
  307. Garbarino, G.; Pampararo, G.; Phung, T.K.; Riani, P.; Busca, G. Heterogeneous Catalysis in (Bio)Ethanol Conversion to Chemicals and Fuels: Thermodynamics, Catalysis, Reaction Paths, Mechanisms and Product Selectivities. Energies 2020, 13, 3587. [Google Scholar] [CrossRef]
  308. Di Nardo, A.; Portarapillo, M.; Russo, D.; Di Benedetto, A. Hydrogen production via steam reforming of different fuels: Thermodynamic comparison. Int. J. Hydrogen Energy 2024, 55, 1143–1160. [Google Scholar] [CrossRef]
  309. Marquevich, M.; Farriol, X.; Medina, F.; Montané, D. Hydrogen Production by Steam Reforming of Vegetable Oils Using Nickel-Based Catalysts. Ind. Eng. Chem. Res. 2001, 40, 4757–4766. [Google Scholar] [CrossRef]
  310. Setiabudi, H.D.; Aziz, M.A.A.; Abdullah, S.; The, L.P.; Jusoh, R. Hydrogen production from catalytic steam reforming of biomass pyrolysis oil or bio-oil derivatives: A review. Int. J. Hydrogen Energy 2020, 45, 18376–18397. [Google Scholar] [CrossRef]
  311. Sepehri, S.; Rezaei, M.; Garbarino, G.; Busca, G. Facile synthesis of a mesoporous alumina and its application as a support of Ni-based autothermal reforming catalysts. Int. J. Hydrogen Energy 2016, 41, 3456–3464. [Google Scholar] [CrossRef]
  312. Park, S.-W.; Lee, D.; Kim, S.-I.; Kim, Y.J.; Park, J.H.; Heo, I.; Chang, T.S.; Lee, J.H. Effects of Alkali Metals on Nickel/Alumina Catalyzed Ethanol Dry Reforming. Catalysts 2021, 11, 260. [Google Scholar] [CrossRef]
  313. Nguyen, A.-T.; Ng, K.H.; Kumar, P.S.; Pham, T.P.T.; Setiabudi, H.D.; Yusuf, F.; Pham, L.K.H.; Show, P.L.; Hussain, I.; Vo, D.-V.N. Sustainable hydrogen production and CO2 mitigation from acetic acid dry reforming over Ni/Al2O3 catalyst. Int. J. Hydrogen Energy 2024, 67, 1044–1055. [Google Scholar] [CrossRef]
  314. Ramezani, R.; Di Felice, L.; Gallucci, F. A review of chemical looping reforming technologies for hydrogen production: Recent advances and future challenges. J. Phys. Energy 2023, 5, 024010. [Google Scholar] [CrossRef]
  315. Tavanarad, M.; Rezaei, M.; Meshkani, F. Preparation and evaluation of Ni/γ-Al2O3 catalysts promoted by alkaline earth metals in glycerol reforming with carbon dioxide. Int. J. Hydrogen Energy 2021, 46, 24991–25003. [Google Scholar] [CrossRef]
  316. Song, K.S.; Seo, Y.S.; Yoon, H.K.; Cho, S.J. Characteristics of the NiO/Hexaaluminate for Chemical Looping Combustion Korean. J. Chem. Eng. 2003, 20, 471–475. [Google Scholar]
  317. Wang, G.; Zhang, S.; Zhu, X.; Li, C. Honghong Shan Dehydrogenation versus hydrogenolysis in the reaction of light alkanes over Ni-based catalysts. J. Ind. Eng. Chem. 2020, 86, 1–12. [Google Scholar] [CrossRef]
  318. Shuikin, N.I.; Tulupova, E.D. Highly active nickel—Alumina catalyst for the dehydrogenation of the cyclohexane ring. Russ. Chem. Bull. 1960, 9, 668–674. [Google Scholar] [CrossRef]
  319. Yolcular, S.; Olgun, O. Ni/Al2O3 catalysts and their activity in dehydrogenation of methylcyclohexane for hydrogen production. Catal. Today 2008, 138, 198–202. [Google Scholar] [CrossRef]
  320. Wang, D.; Lei, Q.; Li, H.; Li, G.; Zhao, Y. Preparation of a Novel NiAlO Composite Oxide Catalyst for the Dehydrogenation of Methylcyclohexane. Catalysts 2022, 12, 958. [Google Scholar] [CrossRef]
  321. Niermann, M.; Beckendorff, A.; Kaltschmitt, M.; Bonhoff, K. Liquid Organic Hydrogen Carrier (LOHC)—Assessment based on chemical and economic properties. Int. J. Hydrogen Energy 2019, 44, 6631–6654. [Google Scholar] [CrossRef]
  322. Chu, C.; Wu, K.; Luo, B.; Cao, Q.; Zhang, H. Hydrogen storage by liquid organic hydrogen carriers: Catalyst, renewable carrier, and technology—A review. Carbon Res. Conv. 2023, 6, 334–351. [Google Scholar] [CrossRef]
  323. Munyentwali, A.; Tan, K.C.; He, T. Advancements in the development of liquid organic hydrogen carrier systems and their applications in the hydrogen economy. Prog. Nat. Sci. Mat. Int. 2024; in press. [Google Scholar] [CrossRef]
  324. Available online: https://ess.honeywell.com/us/en/solutions/sustainability/hydrogen-solutions/liquid-organic-hydrogen-carrier (accessed on 14 August 2024).
  325. Kamal, M.S.; Razzak, S.A.; Hossain, M.M. Catalytic oxidation of volatile organic compounds (VOCs). A review. Atmos. Environ. 2016, 140, 117–134. [Google Scholar] [CrossRef]
  326. Xia, Y.; Lin, M.; Ren, D.; Li, Y.; Hu, F.; Chen, W. Preparation of high surface area mesoporous nickel oxides and catalytic oxidation of toluene and formaldehyde. J. Porous Mater. 2017, 24, 621–629. [Google Scholar] [CrossRef]
  327. Ozkan, U.S.; Kueller, R.F.; Monteczuma, E. Methanol Oxidation over Nonprecious Transition Metal Oxide Catalysts. Ind. Eng. Chem. Res. 1990, 29, 1136. [Google Scholar] [CrossRef]
  328. Darvell, L.I.; Heiskanen, K.; Jones, J.M.; Ross, A.B.; Simell, P.; Williams, A. An investigation of alumina-supported catalysts for the selective catalytic oxidation of ammonia in biomass gasification. Catal. Today 2003, 81, 681–692. [Google Scholar] [CrossRef]
  329. Stoyanova, M.; Konova, P.; Nikolov, P.; Naydenov, A.; Christoskova, S.; Mehandjiev, D. Alumina-supported nickel oxide for ozone decomposition and catalytic ozonation of CO and VOCs. Chem. Eng. J. 2006, 122, 41–46. [Google Scholar] [CrossRef]
  330. Kozhukhova, A.E.; du Preez, S.P.; Bessarabov, D.G. Catalytic Hydrogen Combustion for Domestic and Safety Applications: A Critical Review of Catalyst Materials and Technologies. Energies 2021, 14, 4897. [Google Scholar] [CrossRef]
  331. Sadykov, V.A.; Isupova, L.A.; Zolotarskii, I.A.; Bobrova, L.N.; Noskov, A.S.; Parmon, V.N.; Brushtein, E.A.; Telyatnikova, T.V.; Chernyshev, V.I.; Lunin, V.V. Oxide catalysts for ammonia oxidation in nitric acid production: Properties and perspectives. Appl. Catal. A Gen. 2000, 204, 59–87. [Google Scholar] [CrossRef]
  332. Amblard, M.; Burch, R.; Southward, B.W.L. The selective conversion of ammonia to nitrogen on metal oxide catalysts under strongly oxidising conditions. Appl. Catal. B Environ. 1999, 22, L159–L166. [Google Scholar]
  333. Nassos, S.; Svensson, E.E.; Boutonnet, M.; Järås, S.G. The influence of Ni load and support material on catalysts for the selective catalytic oxidation of ammonia in gasified biomass. Appl. Catal. B Environ. 2007, 74, 92–102. [Google Scholar] [CrossRef]
  334. Chmielarz, L.; Kustrowski, P.; Rafalska-Łasocha, A.; Dziemba, R. Selective oxidation of ammonia to nitrogen on transition metal containing mixed metal oxides. Appl. Catal. B Environ. 2005, 58, 235–244. [Google Scholar] [CrossRef]
  335. Weissenberger, T.; Zapf, R.; Pennemann, H.; Kolb, G. Effect of the Active Metal on the NOx Formation during Catalytic Combustion of Ammonia SOFC Off-Gas. Catalysts 2022, 12, 1186. [Google Scholar] [CrossRef]
  336. Dey, S.; Dhal, G.C.; Mohan, D.; Prasad, R. Advances in transition metal oxide catalysts for carbon monoxide oxidation: A review. Advan. Compos. Hybrid Mat. 2019, 2, 626–656. [Google Scholar] [CrossRef]
  337. Dey, S.; Mehta, N.S. Oxidation of carbon monoxide over various nickel oxide catalysts in different conditions: A review. Chem. Eng. J. Adv. 2020, 1, 100008. [Google Scholar] [CrossRef]
  338. Han, S.W.; Kim, D.H.; Jeong, M.; Park, K.J.; Kim, Y.D. CO oxidation catalyzed by NiO supported on mesoporous Al2O3 at room temperature. Chem. Eng. J. 2016, 283, 992–998. [Google Scholar] [CrossRef]
  339. Jiajian, G.; Jia, C.; Li, J.; Zhang, M. Ni/Al2O3 catalysts for CO methanation: Effect of Al2O3 supports calcined at different temperatures. J. Energy Chem. 2013, 22, 919–927. [Google Scholar]
  340. Maharaj, C.; Lokhat, D.; Rawatlal, R. Oxidation of carbon monoxide over supported nickel oxide catalyst: Kinetic model development and identification. S. Afr. J. Chem. Eng. 2022, 39, 106–116. [Google Scholar] [CrossRef]
  341. Yousra Abdelbaki, Y.; de Arriba, A.; Solsona, B.; Delgado, D.; García-González, E.; Issaadi, R.; López Nieto, J.M. The nickel-support interaction as determining factor of the selectivity to ethylene in the oxidative dehydrogenation of ethane over nickel oxide/alumina catalysts. Appl. Catal. A Gen. 2021, 623, 118242. [Google Scholar] [CrossRef]
  342. Zhang, Q.; Jiang, X.; Li, Y.; Tan, Y.; Jiang, Q.; Liu, X.; Qiao, B. Oxidative Dehydrogenation of Propane over Supported Nickel Single-Atom Catalyst. Chin. J. Chem. 2023, 42, 370–376. [Google Scholar] [CrossRef]
  343. Li, X.; Ma, J.; He, H. Recent advances in catalytic decomposition of ozone. J. Environ. Sci. 2020, 94, 14–31. [Google Scholar] [CrossRef]
  344. Yu, X.; Wang, S.; Gao, H. Spinel NiAl2O4 Based Catalysts: Past, Present and Future. J. Environ. Sci. Eng. Tech. 2023, 11, 12–27. [Google Scholar] [CrossRef]
  345. Dilshara, P.; Abeysinghe, B.; Premasiri, R.; Dushyantha, N.; Ratnayake, N.; Senarath, S.; Ratnayake, S.A.; Batapola, N. The role of nickel (Ni) as a critical metal in clean energy transition: Applications, global distribution and occurrences, production-demand and phytomining. J. Asian Earth Sci. 2024, 259, 105912. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of γ-Al2O3 and 20 wt% and 50 wt% loaded NiO/Al2O3, prepared by impregnation. Arrows are the positions of the peaks of NiO (bunsenite). CoKα radiation.
Figure 1. XRD patterns of γ-Al2O3 and 20 wt% and 50 wt% loaded NiO/Al2O3, prepared by impregnation. Arrows are the positions of the peaks of NiO (bunsenite). CoKα radiation.
Catalysts 14 00552 g001
Figure 2. FTIR skeletal spectra (KBr pressed disks) of γ-Al2O3 and 5 wt%, 20 wt% and 50 wt% loaded NiO/Al2O3, prepared by impregnation, as compared with those of NiAl2O4 and NiO (bunsenite).
Figure 2. FTIR skeletal spectra (KBr pressed disks) of γ-Al2O3 and 5 wt%, 20 wt% and 50 wt% loaded NiO/Al2O3, prepared by impregnation, as compared with those of NiAl2O4 and NiO (bunsenite).
Catalysts 14 00552 g002
Figure 3. Visible and near infrared spectra of NiO, NiAl2O4 and 20 wt% NiO/Al2O3 “monolayer” catalyst.
Figure 3. Visible and near infrared spectra of NiO, NiAl2O4 and 20 wt% NiO/Al2O3 “monolayer” catalyst.
Catalysts 14 00552 g003
Figure 4. Field Emission Scanning Electron Microscopy image of bulk nickel (ex-NiO) after CO2 hydrogenation experiment. Carbonaceous matter is evident, mainly encapsulating the brighter Ni metal particles.
Figure 4. Field Emission Scanning Electron Microscopy image of bulk nickel (ex-NiO) after CO2 hydrogenation experiment. Carbonaceous matter is evident, mainly encapsulating the brighter Ni metal particles.
Catalysts 14 00552 g004
Figure 5. Field Emission Scanning Electron Microscopy image of a Ni/Al2O3-based catalyst (16% Ni) after ethanol steam reforming experiment (right). Carbonaceous matter is evident, mainly carbon nanotubes that contain some bright nickel particles in the interior.
Figure 5. Field Emission Scanning Electron Microscopy image of a Ni/Al2O3-based catalyst (16% Ni) after ethanol steam reforming experiment (right). Carbonaceous matter is evident, mainly carbon nanotubes that contain some bright nickel particles in the interior.
Catalysts 14 00552 g005
Figure 6. Typical H2-TPR curves for samples in the NiO-Al2O3 system.
Figure 6. Typical H2-TPR curves for samples in the NiO-Al2O3 system.
Catalysts 14 00552 g006
Figure 7. IR spectra of low-temperature adsorption of CO on patterns of γ-Al2O3 (top) and 20 wt% NiO/Al2O3, prepared by impregnation, oxidized (middle) and reduced (below). Bands at 2211, 2202 and 2188–2195 cm−1 are due to CO interacting with Al3+ ions. The band at 2183–2195 cm−1 is due to CO interacting with Ni2+. The bands at 2129, 2059, 2044 and 2020 cm−1 are due to polycarbonyls of dispersed Ni0. The band shifting from 2098 cm−1 to lower frequencies by decreasing coverage is due to terminal monocarbonyls over Ni metal particle surfaces.
Figure 7. IR spectra of low-temperature adsorption of CO on patterns of γ-Al2O3 (top) and 20 wt% NiO/Al2O3, prepared by impregnation, oxidized (middle) and reduced (below). Bands at 2211, 2202 and 2188–2195 cm−1 are due to CO interacting with Al3+ ions. The band at 2183–2195 cm−1 is due to CO interacting with Ni2+. The bands at 2129, 2059, 2044 and 2020 cm−1 are due to polycarbonyls of dispersed Ni0. The band shifting from 2098 cm−1 to lower frequencies by decreasing coverage is due to terminal monocarbonyls over Ni metal particle surfaces.
Catalysts 14 00552 g007
Figure 8. FE-SEM micrographs of spent Ni/Al2O3 catalysts after prereduction and CO2 methanation experiments, recorded using backscattered electrons. Left: 16% Ni (20 wt% NiO) loading; right: 39% (50 wt% NiO) Ni loading. Bright particles are Ni metal particles.
Figure 8. FE-SEM micrographs of spent Ni/Al2O3 catalysts after prereduction and CO2 methanation experiments, recorded using backscattered electrons. Left: 16% Ni (20 wt% NiO) loading; right: 39% (50 wt% NiO) Ni loading. Bright particles are Ni metal particles.
Catalysts 14 00552 g008
Table 1. Characteristics of reactions where Ni/Al2O3 catalysts might be applied.
Table 1. Characteristics of reactions where Ni/Al2O3 catalysts might be applied.
Reaction NameStoichiometryΔH°298 (Gas)Typical Reaction Conditions
(kJ/molR)PhaseT (°C)P (Bar)
Acetylene partial hydrogenation in C2 cutsHC≡CH + H2 → H2C=CH2−172Gas
Liq
100–250
25–100
5
20–35
C3–C5 steam cracking cut purificationDiene and acetylenic hydrogenation~−240Liq10–60<100
Pygas hydrotreatment
1 step
Diene and acetylenic hydrogenation~−150Liq70–20020–50
Phenylacetylene partial hydrogenation C6H5-C≡CH + H2 → C6H5-HC=CH2~−120Liq15–50<50
Linear diene hydrogenation for LAB synthesisR-CH=CH-CH=CH2
→ R-CH2CH2CH2CH3
−240Liq170–23010–20
Benzene saturationC6H6 + 3 H2 ⇆ C6H12−208Gas
Liq
500
200–350
30–50
CO methanationCO + 3 H2 → CH4 + H2O−206Gas250–35020–50
CO2 methanationCO2 + 4 H2 → CH4 + 2 H2O−165Gas300–40020–50
Reverse water–gas shiftCO2 + H2 → CO + H2O+41Gas>8501–30
Acetonitrile hydrogenation CH3C ≡ N + H2 → CH3CH2NH2−121Gas100–2001–20
n-octanal hydrogenation C7H15-CH=O + H2 → C7H15CH2OH −57Gas100–20020–50
Methane steam reforming CH4 + H2O → CO + 3H2 +206 Gas500–8001–40
Methane dry reformingCH4 + CO2
→ 2CO + 2H2
+247Gas500–7001–20
Methane partial oxidationCH4 + ½ O2 → CO + 2 H2−36Gas500–10001–40
Methane combustion to COCH4 + O2 → CO + 2 H2O−519Gas>2001–40
Methane decompositionCH4 → C + 2 H2 +76Gas>4001–10
Ammonia decompositionNH3 → ½ N2+ ³/₂ H2+46Gas>4001–10
Cyclohexane
dehydrogenation
C6H12 ⇆ C6H6 + 3 H2 +208Gas4001–3
Ammonia oxidation to NONH3 + 5/4 O2
NO + 3/2 H2O
−226Gas>6001–20
Selective oxidation of ammonia to N2NH3 + ¾ O2
→ ½ N2 + 3/2 H2O
−317Gas>6001–20
CO oxidationCO+ ½ O2 → CO2−283Gas>01–20
Hydrogen combustionH2 + ½ O2 → H2O−242Gas>1001–20
Ethane oxidative
dehydrogenation
H3C-CH3 + ½ O2 → H2=CH2 + H2O−105Gas500–7001–10
Ozone decompositionO3 → 3/2 O2−143Gas−50–0<1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Busca, G.; Spennati, E.; Riani, P.; Garbarino, G. Properties, Industrial Applications and Future Perspectives of Catalytic Materials Based on Nickel and Alumina: A Critical Review. Catalysts 2024, 14, 552. https://doi.org/10.3390/catal14080552

AMA Style

Busca G, Spennati E, Riani P, Garbarino G. Properties, Industrial Applications and Future Perspectives of Catalytic Materials Based on Nickel and Alumina: A Critical Review. Catalysts. 2024; 14(8):552. https://doi.org/10.3390/catal14080552

Chicago/Turabian Style

Busca, Guido, Elena Spennati, Paola Riani, and Gabriella Garbarino. 2024. "Properties, Industrial Applications and Future Perspectives of Catalytic Materials Based on Nickel and Alumina: A Critical Review" Catalysts 14, no. 8: 552. https://doi.org/10.3390/catal14080552

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

Article Metrics

Back to TopTop