Next Article in Journal
Preparation, Characterization, and Catalytic Properties of Pd-Graphene Quantum Dot Catalysts
Next Article in Special Issue
Optimization of Autohydrolysis of Olive Pomaces to Obtain Bioactive Oligosaccharides: The Effect of Cultivar and Fruit Ripening
Previous Article in Journal
Photoactive Materials for Decomposition of Organic Matter Prior to Water Analysis—A Review Containing Original Research
Previous Article in Special Issue
Recent Advances in the Application of Enzyme Processing Assisted by Ultrasound in Agri-Foods: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Steam Reforming of Bioethanol Using Metallic Catalysts on Zeolitic Supports: An Overview

1
Chemical Engineering and Catalysis for Sustainable Processes (CECaSP) Laboratory, University of Calabria, Via Pietro Bucci, 87036 Rende, Italy
2
Department of Physical and Chemical Sciences, University of L’ Aquila, Via Vetoio (COPPITO 1–2), 67100 L′ Aquila, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(6), 617; https://doi.org/10.3390/catal12060617
Submission received: 4 May 2022 / Revised: 31 May 2022 / Accepted: 1 June 2022 / Published: 3 June 2022
(This article belongs to the Special Issue Catalysis in Biorefinery)

Abstract

:
Hydrogen is considered one of the energy carriers of the future due to its high mass-based calorific value. Hydrogen combustion generates only water, and it can be used directly as a fuel for electricity/heat generation. Nowadays, about 95% of the hydrogen is produced via conversion of fossil fuels. One of the future challenges is to find processes based on a renewable source to produce hydrogen in a sustainable way. Bioethanol is a promising candidate, since it can be obtained from the fermentation of biomasses, and easily converted into hydrogen via steam catalytic reforming. The correct design of catalysts and catalytic supports plays a crucial role in the optimization of this reaction. The best results have to date been achieved by noble metals, but their high costs make them unsuitable for industrial application. Very satisfactory results have also been achieved by using nickel and cobalt as active metals. Furthermore, it has been found that the support physical and chemical properties strongly affect the catalytic performance. In this review, zeolitic materials used for the ethanol steam reforming reaction are overviewed. We discuss thermodynamics, reaction mechanisms and the role of active metal, as well as the main noble and non-noble active compounds involved in ethanol steam reforming reaction. Finally, an overview of the zeolitic supports reported in the literature that can be profitably used to produce hydrogen through ethanol steam reforming is presented.

1. Introduction

Nowadays, the development of environmentally friendly fuels is considered one of the greatest technological challenges. The energy dependence on fossil fuels is considered one of the main threats both to world economic development (in particular of those countries that are not energy-independent) and to the protection of the planet’s environment.
Indeed, global warming and climate change are pushing scientific research to develop alternative fuels, environmentally friendly and renewable, with increased energy density [1,2]. Among the investigated alternatives, with their overall carbon neutrality, biofuels are considered valuable components in the energy mix of modern economies [3,4]. The first- and second-generation of biofuels are used in direct combustion processes (primary biofuels), whilst secondary biofuels can be obtained by thermochemical (combustion, pyrolysis and gasification) [5], chemical or biological conversion of residual, waste or side products [6]. Examples of the biofuels and sustainable energy sources are bio-oils, char, pellets [7], and biodiesel from lignocellulosic biomass [8]; green diesel and jet fuels from microalgae oils [9,10]; sugar-derived ethanol [11] and H2 [12]; 2-methylfuran from furfural [13,14,15,16]; H2 from municipal solid waste [17]; dimethyl ether from syngas (via biomass gasification) or CO2 hydrogenation and methanol dehydration [18,19,20,21,22,23,24].
Among the various chemicals from biomass, ethanol (EtOH) certainly plays a crucial role for its versatility, relative low production costs, availability and low toxicity [12,25,26]; it is also promising for the sustainable production of useful compounds, such as hydrocarbons/olefins [27,28,29] and hydrogen via aqueous phase [30] or ethanol steam reforming (ESR) [31,32,33].
Due to its high mass-based heating value (120 MJ/kg), hydrogen is considered a promising energy carrier since its combustion generates only water without any greenhouse gas emission. Hydrogen can be directly employed as fuel for electricity production and for automotive applications [34,35,36]. Furthermore, H2 can be used for ammonia synthesis [37] or in combination with CO and CO2 to produce synthetic fuels and chemicals, such as methane [38,39], methanol [40], and hydrocarbon mixtures [41]. Nowadays, more than 90% of worldwide hydrogen production comes from fossil fuels, but this molecule can be used for the chemical storage of intermittent renewable energy sources via water electrolysis [42,43].
In the overall reaction of the ESR, 1 mol of EtOH and 3 mol of water produce 6 mol of hydrogen in a strongly exothermic reaction [44,45,46,47]:
C2H5OH + 3H2O → 2CO2 + 6H2     ΔHr°298K = 174 kJ mol−1
Research interest in catalytic ESR increased in the last two decades [48]. To maximize hydrogen production, it is crucial to work with an excess of steam, also minimizing the ethanol dehydration (EDHy) and decomposition. Coke formation and metal sintering are the main problems for the stable operation of catalytic ESR. Coke formation is mainly due to the Boudouard reaction, oligomerization of ethylene or decomposition of methane formed during ESR. Coke can considerably reduce the catalyst activity, affecting the catalyst structure and occupying the catalytic sites [48,49].
Depending on the catalyst type and support, the ESR mechanism may follow different pathways as several other reactions may occur, affecting the overall H2 yield. Such a complex reaction pathway requires the tailoring of proper catalysts and support to address the reaction toward hydrogen production with a maximum yield and selectivity [50].
Various materials have been proposed to increase catalyst activity and stability: supported metals, such as rhodium (Rh), nickel (Ni) and cobalt (Co), have attracted attention as efficient catalysts at relatively low temperatures (300–400 °C) [51,52,53,54,55,56].
The use of economic materials for the catalytic system is desired form the perspective of sustainable industrial applications. Zeolites are effective microporous catalysts that can be used as support in ESR, due to their morphological properties (well-defined crystalline structure, high specific surface areas, uniform pores and good thermal stability) [57]. When using zeolites in metal catalyzed reactions, one of the main scientific challenges is the catalyst stability as the acid sites of zeolitic supports can favor ethanol dehydration instead of dehydrogenation with the consequent formation of ethylene that tends to form coke [58]. In turn, coke could deposit on the surface of the catalyst, affecting its performance and leading to progressive deactivation. Therefore, adequate strategies for limiting the formation of coke are the control of acidity and the improvement of the metal dispersion in order to reduce the sintering. Zeolite support can be used for the deposition of small Ni particles, but in presence of significant acidity (due to heteroatoms insertion in the framework, such as Al), it might favor coke formation, especially at low temperatures [59]. To overcome those difficulties, different techniques can be used to reduce the zeolite acidity (for example, post-synthesis dealumination) [60,61,62], also promoting the formation of small metal particles and reducing coke formation as a consequence [33]. In addition, the metal–support interaction strength could be tuned: a greater dispersion of the metal also prevents metal sintering under the operating conditions for ESR [63].
With respect to previous general reviews in the field [12,32,33,64], this paper proposes the systematic analysis of zeolite-based catalysts for ESR, mostly prepared using base metals as redox sites, highlighting the most important catalyst design features. Accordingly, in the first part, the thermodynamics of the process and potential reaction mechanisms occurring on metallic active sites are briefly discussed. Then, the mainly used active metals for the ESR are discussed, especially focusing on non-noble ones that can potentially improve its economic feasibility. Finally, the main zeolite-based catalysts in the open literature are discussed in relation to their potential profitability for H2 production.

2. Ethanol Steam Reforming

2.1. Thermodynamics of Ethanol Steam Reforming

In this section, the thermodynamics of hydrogen production via ESR is considered. The effect of reaction temperature and pressure on the outlet gas composition is investigated. As reported in Equation (1), ESR can be ideally expressed by means of a single reaction leading to the production of hydrogen and carbon dioxide.
The chemical equilibrium composition was calculated through the minimization of Gibbs free energy for a stoichiometric reaction mixture. Figure 1a presents the equilibrium gas composition at 1.5 bar for a temperature range between 200 and 800 °C, while Figure 1b presents the hydrogen yield as a function of temperature for different pressure values, calculated according to the following equation:
η H 2 = n ˙ H 2 6 n ˙ C 2 H 5 O H
where coefficient 6 represents the stoichiometric coefficient ratio between hydrogen and EtOH in Equation (1).
As expected, temperature increase favors hydrogen formation because the investigated reaction is endothermal. A plateau in the hydrogen yield is already reached at ≈400 °C, if the reaction of Equation (1) is carried out at 1.5 bar. As suggested by thermodynamics and as shown in Figure 1b, high pressure limits hydrogen yield since reaction (1) proceeds with an increase in the mole number.
In real operating conditions, the reaction pathway of ESR includes three reactions: ethanol dehydrogenation (EDH) (3), acetaldehyde steam reforming (ASR) (4) and water gas shift reaction (WGS) (5) [33].
C2H5OH ↔ CH3CHO + H2     ΔHr°298K = 68.9 kJ mol−1
CH3CHO + H2O ↔ 2CO + 3H2     ΔHr°298K = 168.8 kJ mol−1
CO + H2O ↔ CO2 + H2     ΔHr°298K = −41.4 kJ mol−1
At low temperatures, exothermic WGS is thermodynamically favored but CO for water gas shift should be produced from reaction (4), which is, in turn, endothermal and thus favored at high temperatures, as well as reaction (3). Increasing the reaction temperature should thus favor hydrogen production, but a maximum in H2 yield/selectivity may occur. Various undesired side reactions may take place in ESR reaction conditions, such as ethanol dehydration into ethylene (6), the decomposition of acetaldehyde (7) and acetone formation (8). Furthermore, aldol condensation and coke formation are also reported [33].
C2H5OH ↔ C2H4 + H2O     ΔHr°298K = 45.5 kJ mol−1
CH3CHO ↔ CH4 + CO     ΔHr°298K = −19.3 kJ mol−1
CH3CHO ↔ CH3COCH3 + CO+ H2     ΔHr°298K = 4.1 kJ mol−1
Particularly, acetaldehyde decomposition is targeted for its suppression because it produces inactive methane, decreasing the hydrogen selectivity as consequence. The promotion of adsorbed acetate species formation from acetaldehyde is effective for the preferential promotion of the ASR and the suppression of the acetaldehyde decomposition to CH4. The methanation of CO (9) and CO2 (10) are hydrogen-consuming undesired reactions that should be inhibited as much as possible during ethanol steam reforming [33]. Furthermore, reactions (9) and (10) are undesirable since they are very exothermal reactions (see reaction enthalpy values) and it is very difficult to control the reactor temperature when these reactions take place. From a thermodynamic and kinetic standpoint, a temperature rise is not challenging, but it may compromise the catalyst integrity.
CO + 3H2 ↔ CH4 + H2O     ΔHr°298K = −206.1 kJ mol−1
CO2 + 4H2 ↔ CH4 + 2H2O     ΔHr°298K = −164.8 kJ mol−1
These additional reactions introduce new compounds among the possible equilibrium products. Thus, the equilibrium calculation was carried out via global Gibbs energy minimization also assuming the presence of methane, acetaldehyde, acetone and carbon monoxide.
Figure 2 presents the simulation results revealing that ethylene, acetone and acetaldehyde are present only in traces when equilibrium is reached. Especially at low temperatures, methane formation is thermodynamically favored over the hydrogen one. This is mainly due to the equilibrium of the methanation reaction, leading to high conversion at temperatures below 400 °C. This behavior is amplified when pressure increases since methanation reactions (9) and (10) undergo through a mole reduction. On the other hand, H2 molar fraction grows with increasing temperature because its production is favored by the thermodynamics of the ESR reaction (1). At 1.5 bar, the hydrogen yield presents a peak at ≈700 °C, as hydrogen production is counterbalanced by the consumption via reverse water gas shift (RWGS), which is endothermal and starts to be predominant at high temperatures. This trend is confirmed by the simultaneous increase in CO and decrease in CO2 molar fractions when the temperature rises (especially above 600 °C).
Even though some reactions are favored by thermodynamics, the situation may thus change in the presence of a catalyst enabling only some reaction routes and affecting reaction kinetics. An ideal catalyst should inhibit methanation reaction, especially if the target consists of ESR at low temperatures. Furthermore, such catalyst should not enable (as preferential reaction routes) ethanol dehydration (EDHy) to ethylene, as well as all the reaction steps leading to carbon formation and deposition.
Several coke-forming reactions may also occur in ESR, via ethylene (11) or methane decomposition (12) and Boudouard reaction (13). Suppressing ethylene and methane formation and promoting WGS reaction can thus limit carbon deposition [33].
The addition of excess steam can be a solution for carbon formation as it is usually limited by increasing the steam excess. This aspect is offset by consequences on energy efficiency due to the required heat for generating steam, which should be considered during the whole process energy (and economic) analysis.
C2H4 ↔ C(s) + 2H2     ΔHr°298K = −52.3 kJ mol−1
CH4 ↔ C(s) + 2H2     ΔHr°298K = −74.9 kJ mol−1
2CO ↔ C(s) + CO2     ΔHr°298K = −172.6 kJ mol−1
Sharma et al. [31] reviewed the phenomenon of carbon formation during ESR, including the effect of involved active metal, support and operating conditions. Furthermore, the main strategies to limit the catalyst deactivation due to coking were summarized. The authors pointed out as the best approach to reduce catalyst deactivation due to carbon formation is either by preventing carbon from forming or, if formed, by converting it into the gaseous species for easy removal [31]. Some aspects related to the relationship between active metals and carbon formation are briefly mentioned in the following sections.

2.2. Reaction Mechanism

Figure 3 summarizes plausible reaction pathways involved in the ESR over the metal surface, based on both experimental and theoretical results. The EtOH activation may occur via several pathways. One possibility is the cleavage of O–H bond, followed by a dehydrogenation forming intermediates such as acetaldehyde (CH3CHO*), acetyl (CH3C*O), ketene (*CH2C*O) and ketenyl (*CHC*O). Another possible pathway is represented by the C–H bond activation and a subsequent dehydrogenation, preserving the O–H bond with the formation of generic intermediates of the type *CHyCHx*OH. Finally, the cleavage of bonds O–H and C–H may occur, forming the intermediate *CH2CH2O* (oxametallacycle). These pathways can take place simultaneously, depending on the nature of the metal. The cleavage of the C–C bond within the chemisorbed intermediates can be followed by the hydrogenation/dehydrogenation of CHx*, water activation and the oxidation of C* species [65,66].
The oxidation of C* is crucial in ESR in order to prevent catalyst deactivation via carbon formation. The development of catalysts for ESR is related to the tuning of different stages: ethanol activation, C* formation and the removal of the C*, as well as the ability to inhibit the hydrogenation rearrangements of CHx fragments leading to methane formation at low temperatures [65].
EtOH activation usually occurs through its O–H group. Nevertheless, recent studies have shown that ethanol can be activated through C–H bonds [67]. EtOH activation is an important step in determining the main reaction pathway. Dehydrogenated intermediates will bring interesting information to elucidate the C–C bond cleavage, usually occurring in an intermediate with a high dehydrogenation level.
For all pathways, CH4 selectivity is affected by the ability of the metal to hydrogenate/dehydrogenate the CHx* species [68]. Methane formation at low temperatures is undesirable because the reforming of CH4 (leading to hydrogen production) will only occur at greater temperatures (typically above 800 °C). The activation of H2O is of paramount importance in ESR due to the role of *OH species in oxidation steps. After the C–C bond cleavage and steam activation, the CO adsorbed on the metal is oxidized to generate CO2 via water gas shift (WGS) [65]. CHx species oxidation on the metal proceeds at about 320 °C. The most widely accepted mechanism for the oxidation of CHx species is CHx decomposition to C*, followed by oxidation to CO [65].
In summary, there are still some open questions related to the reaction mechanism of ESR on different metals. The right balance of the involved steps is crucial for the catalyst performance optimization, as well as for the sample stability against carbon deposition [65].

3. Metals for ESR

Several active metals have been recently investigated for ESR, including both noble (Rh, Ru, Pd, Ir and Pt) and non-noble metals (Ni and Co), generally active in the range 300–500 °C [69]. The activity decreases in the order of Rh > Co > Ni > Pd, consistent with their resistance to coke formation [34,70].
Although noble metal catalysts afford high ESR activity, their industrial application is mainly limited by the high cost [71,72]. Therefore, such metals will not be further discussed in this paper.
Co and Ni-based catalysts present lower activity than noble metals for the ESR reaction. Moreover, deactivation by coking, sintering and methane formation (e.g., via hydrogenation of CO and CO2) limit their potential [33,73]. However, even in the presence of the reported drawbacks, their low costs make them attractive for ESR. Moreover, redox supports and promoters, such as CeOx or MnOx, mitigate their deactivation by coking [74,75].
Among the non-noble metals, nickel shows the highest methane formation [33]. However, this metal is still widely used for ESR due to its significant bond-braking (e.g., C–C, O–H and C–H) [33,69] and H2 recombination capability. As previously shown, Ni dehydrogenates EtOH to hydrogen and acetaldehyde [65] at low temperatures. Besides, the C–C bond cleavage also occurs when increasing temperatures, leading to acetaldehyde decomposition into methane, carbon monoxide and hydrogen. Therefore, due to the C–C bond-breaking activity, the decomposition of ethanol (14) and acetaldehyde conversion to methane and CO easily proceed over Ni catalysts. Then, the Methane Steam Reforming (MSR) (15) and WGS reaction (5) lead to H2 and CO2. Notably, the endothermal MSR is not thermodynamically favored at low temperatures, leading to increased CH4 selectivity [33].
C2H5OH ⟶ CH4 + CO + H2     ΔHr°298K = 49.6 kJ mol−1
CH4 + 2H2O ⟶ CO + 3H2     ΔHr°298K = 206.1 kJ mol−1
For this reason, conventional nickel-based catalysts generally produce a greater number of byproducts (coke and CH4) than cobalt-based ones.
On the other side, highly dispersed Ni on hierarchical BEA zeolites present low deactivation by coking due to the low driving force for carbon diffusion [69]. Similarly, improved coking and sintering resistance were found on Ni/perovskites. Moreover, Sharma et al. [31] evidenced how highly dispersed metallic particles favor the oxidizing activity, while larger metallic particles promote carbon formation because of a reduced interaction with the support. It emerges that the preparation of highly dispersed Ni catalysts is advantageous from several points of view. Therefore, the sol–gel method [69,76] and the deposition of Ni on microporous/mesoporous supports are highly advantageous for preparing durable catalysts, promoting the presence of highly dispersed Ni particles [69] and increasing the catalyst activity as a consequence. Several studies focused on the dependence of the catalytic activity on the controlled size of metal particles [77]. Specifically, higher activities were observed using particles with average diameters lower than 5 nm due to their greater surface energy [78]. Notably, besides the presence of highly dispersed and stable metals, supports also play a key role in activity due to their acidity/basicity and interaction with the active phase (metal–support interaction) [31]. Specifically, basic supports could inhibit ethanol dehydration (EDHy) to ethylene, thus limiting one of the precursors for carbon formation [33]. Among the investigated supports were Al2O3, SiO2, CeO2, La2O3, ZrO2 [75,79,80,81,82,83], promoted M-Al2O3 (M = La, Zr, Ti), Ce-SiO2, [33], zeolites (BEA, ZSM-5, and Y) [33,69,84,85,86] and mesoporous materials (SBA-15, MCM-41) [54,87,88].
Cobalt, similarly to Ni, seems to favor the formation of acetaldehyde at low temperatures [65], enabling the formation of carbon oxides and decreasing their methane formation activity [33]. DFT studies [33] have shown that metallic Co favors C–C cleavage, while Co2+ species enable water activation and acetate species formation. Notably, while metallic Co is also active for coke formation, Co2+ can oxidize the deposited coke, enabling the formation of activated OHx species. Therefore, the control of Co0/Co2+ ratio is crucial to prevent catalyst deactivation [49]. Conversely, other species, such as Co3O4 spinel formed in the case of Co/CeO2 catalysts, together with metallic Co and CoO, are inactive [33,64,89].
Ni-Co bimetallic catalysts have shown a higher activity than single Ni or Co catalysts [54,87,90,91]. The idea is to exploit the positive characteristics of both metals in a synergistic way: nickel showed great activity, especially in the C–C cleavage, while cobalt usually limits coke formation. Furthermore, cobalt addition can promote WGS reaction [33].
The typical performances of the recently reported Ni, Co, and Ni–CO catalysts are reported in Table 1. A wider results overview can be found in previous reviews [33,64]. Notably, the space velocities reported in Table 1 were calculated in different ways: by referring to the whole inlet flow, the EtOH mass flow, the reaction volume or the catalyst mass. In some works [53,54,55,87,92], space velocity was calculated as the ratio between total inlet flow (mL h−1) and catalyst load (in grams). When total inlet flow (i.e., reacting mixture plus inert gas) is considered, reacting mixture mole concentration (or, in other terms, dilution grade) also plays a role, as the reaction rate is proportional to the reactants’ partial pressure.
A comparison between different catalysts should take into account all the parameters affecting the operating conditions: temperature, space velocity, inert presence and steam-to-carbon ratio (EtOH conversion will be higher if a steam-excess inlet mixture is employed). Furthermore, the catalyst should be usually activated/reduced by flowing hydrogen (pure or diluted) at high temperatures before activity testing in order to reduce the metal oxide presence.
By rationalizing main findings, nickel and cobalt can be employed as non-noble active metals for ESR. Bimetallic Ni-Co-based catalysts are extremely promising, as they can synergistically combine the advantages of both the main non-noble metals employed for ESR. Furthermore, in previous research works, it has been pointed out that catalysts with reduced metal particle sizes are more active and stable [33,97,98]. Besides the metal particles’ size, metal oxidation state control can help (especially when cobalt is employed) in limiting carbon formation.
Catalyst performance (activity and stability) is strongly affected by the support in terms of the metal–support interaction and acidity/basicity: acid supports can drive the formation of ethylene, which is considered a precursor for coke formation.

4. Zeolitic Support for ESR

Besides the selection of metal phases, the support plays a major role in catalytic reactions. A high surface area and the heat resistance of the support can improve the catalytic activity and stability [31].
Large surface area supports, such in the case of γ-Al2O3 [99,100,101], mesoporous silica materials [88,102,103] and mixed oxides [104,105], usually allow for the formation of strongly dispersed surface precursors and retard sintering processes. Zeolites possess remarkable properties, such as well-defined crystalline structures, good thermal stability, high internal surface area, intracrystalline mesopores [106] and low potential for support–metal interactions; such properties make this class of materials suitable for use as catalyst support [107].
Their channel structure enables the flow of EtOH molecules into the channel (generically, the diameter of the microporous structure of the zeolite utilized as support is slightly bigger than that of EtOH), while extending the time of reaction and promoting the EtOH reaction on active sites. Hierarchical zeolites have highly ordered transgranular and intragranular mesopore channels, large surface area, pore volume and thick pore walls [69].
Zeolite and mesoporous materials are very interesting materials not only for acid-catalyzed reactions, but also promoting red–ox reactions since it is very easy to incorporate some metals in the zeolitic framework or to add metals by the ionic exchange procedure [108,109].
As discussed above, the EtOH molecules in the stream reforming reaction are dehydrogenated, leading to the formation of ethoxy groups. A mechanism that favors this reaction is the absorption of EtOH on the surface of the catalysts. Such species can be subsequently transformed into various intermediates by successive EDH and EDHy reactions, C–C bond scission and oxidative surface reactions with the contribution of hydroxyl groups or oxygen species in the catalysts [85].
However, the acidity of the zeolites could also enhance the cracking and isomerization of the formed hydrocarbon (Figure 4) [110]. For this reason, the high molar concentration of the acid sites present on the surface of the zeolite support can catalyze the reaction toward the formation of undesired by-products [85].
The building blocks of the zeolite are SiO4 tetrahedra (with trivalent Al3+ cation substituting Si4+) [111]. The angle formed by atoms varies in a broad interval (130–180°) [112], giving rise to the elevated number of known zeolite structures [110]. The introduction of trivalent aluminum atoms in the tetrahedral framework brings negative charges, balanced by extra-framework cations [113]. Protons acting as compensating cations near aluminum centers behave as Brønsted acid sites [23,27,114,115,116,117,118], while tri-coordinated silicon atoms located on small structure defects act as electron-acceptor sites, i.e., as Lewis acid centers [113].
It has been mentioned that the acidity of the support is not a determining parameter during the ESR reaction at high temperatures, while a minimum concentration of the Brønsted sites is sufficient to promote the EDHy into ethylene and consequently the coke formation even at low temperatures [59,119,120]. Moreover, different zeolites have been investigated as catalytic support in ESR.

4.1. ZSM-5 Zeolites

ZSM-5 is a high-silica zeolite widely used as catalyst and its applicability in ESR and was investigated [56,121]. The framework type of the high-silica zeolite ZSM-5 can be described in terms of units, but it is easier to use pentasil units. These units are linked to form pentasil chains, and mirror images of these chains are connected via oxygen bridges to form corrugated sheets with 10-member ring holes. Each sheet is linked by oxygen bridges to the next to form the 3-dimensional structure. Adjacent sheets are related to one another by an inversion center. This produces straight 10-member ring channels parallel to the corrugations (along y) and sinusoidal 10-member ring channels perpendicular to the sheets (along x). Both channel types are interconnected to one another, forming a 3D 10-member ring channel system. Because the pores are formed of 10-member rings rather than 12-member rings, the shape selectivity for sorption and catalysis is different from that of FAU- or EMT-type zeolites, and this fact determined different catalytic applications as ZSM-5 is used in different refinery and petrochemical processes [122].
The physical characteristics of this zeolite make it good from a dimensional point of view if applied as a catalytic support for the ESR reaction [59]. Furthermore, the structure of the interconnected channel of the ZSM-5 zeolite can limit the deactivation of the catalyst caused by coke when ESR is carried out at high temperatures [57].
Kumar et al. [50] have tested ZSM-5 zeolite (with Si/Al ratio equal to 50) coupled with non-noble metals (Ni and Co). Prepared via wet impregnation, samples were subsequently tested at a space velocity of 35.4 ggas h−1 gcat−1. The performance of ZSM-5 has shown a more than 90% EtOH conversion at 450 °C. However, the product distribution varied with the temperature and the active metals deposited. The 10 wt.-% Ni/ZSM-5 catalyst has shown higher C2H4 and lower H2 selectivity than the 10 wt.-% Co/ZSM-5 samples, indicating that Ni is more active for the scission of the C–C bond at low temperatures with ZSM-5 support as compared to Co.
Lang et al. [56] tested metal/zeolite catalytic structures prepared with Rh and ZSM-5 treated with an alkaline solution. The use of the alkaline solution has the dual objective of exchanging cations of alkali metals (to reduce the intrinsic acidity of the zeolite) and to induce a mesoporosity via zeolite desilication. They found that the introduction of secondary porosity allows the better diffusion of reactants delaying the micropores blocking. The different diffusive flow rate lowered the deposition of carbon in the micropores, significantly improving the catalyst stability since an induced secondary mesoporosity reduces the diffusion limitation. The best results were obtained by treating a ZSM-5 with a solution of K2CO3 [56]. The authors reported that modified support improves the Rh dispersion preventing sintering as the metal was deposited via the wet impregnation method using a solution of RhCl3, allowing a concentration of 1% (w/w) of the metal in the final structure of the catalyst. From the characterization data, it can be deduced that the treatment led both to the insertion of K as counter-ion (about 3 wt.-%) and to suitable modifications of the support structure. Changes occurred mainly in terms of Si/Al ratio reduction (from 80 to 48.1) and textural properties (see Table 2). The ion exchange led to a significant decrease in acid sites from 450 mmol g−1 (for zeolite in H-form) to 93.5 mmol g−1 for modified ZSM-5. The net drop in acidity allows the catalytic system to work even at low temperatures; EtOH conversion (Χ) at 300 °C and at 400 °C was reported to be equal to 79.27% and 99.07%, respectively. H2, CO and C2H4 selectivity (S) are reported in Table 2. The authors justify this as a synergic effect between the active Rh component and the structure of the zeolite, which reduces the selectivity of CO produced via RWGS down to less than 3.5% at 400 °C. At low temperatures, the main reactions are the EDH to acetaldehyde (3) and its subsequent decarboxylation to CH4 and CO. The ASR and MSR reactions are not easy to be conducted at low temperatures, resulting in high CH4 selectivity for all catalysts with promising hydrogen selectivity [56].
Similar investigations were performed on MOR-type zeolites [123]. Additionally, in this case, the zeolite was desilicated (with a 0.2 M NaOH solution) to generate mesoporosity and to reduce the acid sites. The catalyst, including Ni based on desilicated MOR zeolite, showed a high activity, selectivity and resistance to coke deposition. Catalytic tests were carried out with a large steam excess in the inlet mixture (H2O/EtOH = 13). The results suggest that the mesoporosity generated in this MOR-type zeolite makes the Ni–MOR catalytic system even more selective towards hydrogen at relatively low temperatures (400 °C).

4.2. BETA Zeolites

The structure of zeolite beta is found to be a high defective intergrowth of two distinct new zeolite frameworks, named polymorphous A and B, respectively [124]. Zeolite beta compositions, thus, represent a narrow window of a broader family of zeolites, of which the polymorph A and B structures with unfaulted staking sequences represent the two end members. All members of this family of zeolites have two sets of perpendicular channels, which intersect to form a three-dimensional array of cages that have three 12-member ring apertures [125]. Zeolite beta is disordered in the c-direction. That is, well-defined layers are stacked in a random way. The units are joined to one another via four rings to form layers with saddle-shaped 12-member rings. Adjacent layers are related to one another by a rotation of 90° [126].
The disorder arises because this rotation can be in either a clockwise or counter-clockwise sense. If the counter-clockwise or clockwise rotation was maintained throughout the crystal, the structure would be ordered and chiral. Interestingly, whatever the stacking sequence, a 3-dimensional 12-member ring channel system results; so, for catalytic applications, the stacking sequence is not important [122]. The average diameter of the microporous structure of the BEA zeolite is intrinsically larger (6.7 Å) than that of the ZSM-5-type (from 5.1–5.6 Å). This characteristic allows the better diffusion of ethanol [29,59,69]. As in the case of ZSM-5, if the catalytic system works at low temperatures (300–500 °C), the Bronsted acid sites of the zeolite catalyze the EDHy reaction with a consequent high yield of ethylene. For example, a Ni-BEA catalyst (Ni: 10 wt.-%) tested at 500 °C tends to rapidly deactivate (decreasing values of conversion, of S to H2 and CO2 selectivity, and increasing values of ethylene selectivity), despite having excellent values of EtOH conversion and H2 selectivity in the first hours of the reaction as reported from Gac et al. [85]. To minimize the production of ethylene, the authors reported on the testing of tested dealuminated BEA zeolite as support. The removal of aluminum atoms from the zeolite structure not only led to a drastic acidity reduction (Bronsted acid sites decreased from 268 to 3 μmol g−1, while Lewis acid sites decreased from 130 to 3 μmol g−1 after dealumination), but also led to a slight decrease in crystallinity, which promoted the stabilization of the metal nanoparticles. BEA dealumination favored the adsorption of Ni precursors, hindering the agglomeration of the ionic species of the metal [59,127]. This catalytic material showed a greater active surface area than the non-dealuminated one. The sample was tested at 500 °C for 20 h and showed the complete conversion of the EtOH [85]. In the first hour of the reaction, the catalytic system of the non-dealuminated beta led to a higher hydrogen yield than that of the dealuminated zeolite (85% against 76%). However, the hydrogen selectivity for the unmodified zeolite drastically decreased (about 60%) in the first 5 h of time on stream (TOS) for the benefit of the ethylene yield (about 60% after three hours of reaction). This means that this kind of catalyst was more effective in olefin formation. Dealuminated zeolite, instead, showed a good constant H2 selectivity during all the 20 h of the reaction and also showed ethylene yield (Y) values close to 0 for about 17 h of the reaction. Despite the excellent results, the problems that arose for these catalysts is that the produced CO during ESR from the dealuminated sample was higher than that generated in the non-modified sample.
In several applications, H2 must be extremely pure, and the admitted percentage of impurities must be extremely low. For example, if hydrogen is used for PEM fuel cell applications, the acceptable content of CO is 10 ppm [128,129,130]. For this reason, the catalytic performance of the Ni-BEA catalyst was also studied by introducing other metals to maximize the yield of H2 and CO2, by converting the undesired by-products. In the work of Tian et al. [131], the acid force of the zeolite support is limited by partially replacing the Al content with trivalent Fe cations in the structure, thus limiting the concentration of proton sites within the catalyst. This hetero-atom substitution is very easy, as demonstrated in different articles [132,133]. In addition, the introduction of metals with variable valence inside the zeolite structure can improve the chemical–physical properties by satisfying the needs of specific catalytic reactions [131]. The Fe introduced in the structure has a different bond length (the Fe–O bond is 0.197 nm, while the Al–O bond is 0.175 nm), leading to a partial distortion of the crystal lattice. This Fe–BEA-type zeolitic support reduced the order of the mesoporous phase and led to a decrease in the number of both moderate and strong acid sites. Such a decrease in acid sites inhibited the EDHy, while the presence of Fe improved the yield of the WGS reaction. In particular, the 10% Ni/0.15% Fe–BEA catalyst showed an H2 selectivity of up to 72.15% and an EtOH conversion rate of 99.6% at 500 °C, while the amount of coke deposition was 4.3% after a 12 h reaction [131].
Using a trimetallic system with Cu, Fe, and Ni (loaded with 1.5, 1.5, and 10 wt.-%, respectively), a greater catalytic activity was found compared to monometallic catalysts, especially at low temperatures (300 ° C). Zheng et al. [134] showed how the synergistic effect among the Ni, Cu, and Fe phases influences the different reaction pathways. Figure 5 shows the reaction scheme as a function of the catalytic activity of the specific metals.
The first reaction step is the formation of acetaldehyde through the EDH catalyzed by the Cu sites. At 300 °C, Ni is active in the transformation of acetaldehyde into CH4 and CO, since, at this temperature, there is almost exclusively this reaction as the C–C bond strength of EtOH is much higher than that of acetaldehyde [72]. Cu has not completed its catalytic activity; in fact, at temperatures of 300 °C, it is strongly active in the transfection of CO (formed by EDHy) into H2 and CO2 through the WGS reaction. At temperatures above 350 °C, Fe catalyzes the ASR reaction, while at temperatures above 450 °C, it catalyzes the MSR reaction. From these experimental results, it is evident that the Fe and Cu phases play an important role in the ESR reaction mechanism, especially at low temperatures as promoters. The synergy of the Ni, Cu and Fe phases maintained the activity of the catalyst almost constant for 28 h [72].
Table 3 presents the results of experimental investigations carried out over BEA-supported metal catalysts.

4.3. Y Zeolite

Zeolite Y is another type of zeolite suitable for industrial applications also because of its affordable cost [135], especially for the catalyst used in fluid catalytic cracking processes. Zeolites Y have different forms: NaY zeolite synthesized from sodium aluminosilicate gels, NH4+ and HY zeolites prepared from zeolite Y by ion exchange with NH4Cl or polyvalent metal cations followed by the thermal treatment [136,137], and Ultra Stable Y (USY) zeolite characterized from a bulk Si/Al ratio higher than the Y form, resulting from hydrothermal treatments [138]. The cubic unit cell of all these aluminosilicates contains 192 (Si,Al)O4 tetrahedrons [126].
Inokawa et al. [139] tested nickel supported over zeolite Y in ESR. The first results showed that the incorporation of transition metal cations in the Y-type zeolite structure using an ion exchange process has a significant influence on ESR (Table 4). In fact, this type of catalytic support led to the production of C2H4 mainly during reforming with EtOH steam, indicating that the presence of cations favored the EDHy. A second study by the same authors on the type of catalysts [140] showed how basicity controlled by the exchange of alkaline cations in a zeolite (exchange of Na with K and Cs) led to a change in the characteristics of both the support and metal. Improved reducibility and catalytic activity led to a higher H2 yield and selectivity. Hydrogen selectivity has been related with the size of the cation (Ni/Na_Y < Ni/K_Y < Ni/Cs_Y). As previously shown, the dehydrogenation reaction of EtOH was accelerated by the basicity of the zeolite, while the EDHy was inhibited. The addition of K in catalysts using Y-type support was also investigated on Rh/NaY samples. Potassium addition favored the conversion of EtOH, which increased from 62 to 97% [84]. The authors of this study concluded that the addition of K increased the yield of hydrogen, but the flow of reactants has a greater influence. In fact, the results show that, at the same temperature (300 °C) and H2O/EtOH molar ratio (5:1) and with the same catalyst (K-Rh/NaY), there is a marked increase in the selectivity of hydrogen, which passes from about 45% for a reagent flow equal to 2.04 g min −1 to about 68% for a reagent flow equal to 2.77 g min −1. It has also been [141] shown that catalysts of the Mg/Ni/Ga_Y tend to reduce the formation of coke on the catalyst compared to the bimetallic catalyst Ni/Mg_Y. This type of catalyst showed a very high selectivity of H2 and CH4 only at a temperature ranging from 550 °C to 750 °C. Additionally, in this case, the results show how the multi-metal component plays a different role to that of the ESR reaction by increasing the EtOH conversion and H2 yield to 100% and 87%, respectively (see Table 4). This high performance was maintained for up to 59 h at a temperature of 700 °C (H2O/EtOH = 3 and GHSV = 6740 h−1).

4.4. ITQ Zeolites

Pure silica ITQ-2 [142] and ITQ-18 [143] are two zeolites with a high external surface area that have shown good results as catalytic supports for Ni and Co metals for ESR. These catalytic systems have shown excellent performances in terms of catalytic activity, H2 selectivity and low coke deposition. The excellent catalytic properties can be related to the low molar concentrations of acid sites in the zeolite. In fact, the absence of acid sites in pure silica ITQ zeolites limits the EDHy reaction that leads to the formation of ethylene and consequently of coke. Low coke deposition can be correlated with both the high external surface and the structure with a series of external pockets distributed along of sheets, which favor the stabilization of Ni and Co metal, improve their dispersion and prevent their agglomeration.
When the ITQ-2-supported catalysts are tested above 500 °C, H2, CO2, CO and CH4 are the only products. Both the catalysts Co_ITQ-2 and Ni _ITQ-2 exhibit lower CO selectivity and high hydrogen yields (see Table 5). The total absence of acid sites strongly limits the production of ethylene and, consequently also, the catalyst deactivation; both samples remain active even after 72 h at a temperature of 400 °C [142].
In the case of Co_ ITQ-18 catalysts [143], after 24 h of activity, no deactivation was detected, even though there was a significant concentration of deposited coke correlated to the slightly acidic properties of the ITQ-18 zeolite (Si/Al = 100). The catalytic system Co_ITQ-18 (H2O/EtOH = 13, GHSV = 4700 h−1, atmospheric pressure and temperature equal to 500 °C), showed high values of both H2 selectivity (71% after 24 h), but it also produced very low concentrations of CO and CH4 (of 3.1% and 5.7%, respectively).

4.5. Core Shell Zeolites

As seen before, the use of Ni deposited on zeolite support tends to greatly improve the catalytic activity and the H2 yield, especially at low temperatures. However, the competitive catalytic effects of the various metals lead to the excessive production of by-products, such as CO, CH4, and acetaldehyde. Therefore, a further challenge for the ESR reaction is the control of by-product formation and, thus, the selective production of hydrogen.
Core-shell type nanoparticles can be defined as constituted by a core (inner material) and a shell (outer layer material) [144,145]. These types of materials combine the advantages of the passivation induced by the shell with the high surface area and accessible pore channels of the core, offering new opportunities in several applications [146,147]. Furthermore, the successful synthesis of several zeolite composites, such as MFI-MEL, MFI-MFI, BEA-MFI and MOR-MFI, with core–shell structures for different catalytic areas has been reported [148,149]. In the specific case of the ESR reaction, the core–shell metal zeolitic structures showed both a great catalytic activity and a high selectivity of H2 even at low temperatures, as well as properties in the reduction of by-products [148,150,151,152].
Very promising results have been obtained using a BEA-type zeolite structure as a shell surrounding a metal core [148,152] (Table 6). Various research works propose the synthesis of a metallic nucleus for the reaction of ESR dispersed over different supports. A variety of products, such as H2, CO2, CH4, CO and intermediate acetaldehyde, are obtained within the core structure [151].
The products then pass through the shell again. A schematic representation of the diffusion in core–shell zeolites is represented in Figure 6.
The addition of some metals within the catalytic system can favor the transformation of by-products into CO2 and H2. For example, as seen previously, the addition of Cu inside the catalytic system can considerably improve the ability to break the O–H bond even at low temperatures, thus favoring both a complete conversion of EtOH and the MSR reaction, while the addition of Fe tends to catalyze the reaction of ASR and WGS [153]. The mixture of gaseous by-products, such as CO and CH4, therefore tends to pass through core–shell systems positioned in series. The catalytic activity of the metals favoring MSR and WGS enables the syngas purification from the smaller molecular size by-products within these systems. Zheng et al. [152] studied BEA-type core–shell catalysts consisting of a core supporting Cu and Fe and a Ni-based shell. The results showed significantly improved catalytic activity and a better H2 selectivity if compared to the BEA-type zeolite as support. Furthermore, the presence of Cu and Fe contributed to obtain high purity syngas.

5. Comparison and Future Trend

The previous paragraphs show that the ESR reaction is a reliable alternative to obtain “green” hydrogen, especially when using biomass-derived ethanol. However, a rationally designed catalysts and optimized conditions are necessary to maximize H2 yields and minimize its environmental impact (e.g., decreasing the amounts of GHG by-products such as CO, CO2 and CH4) (Table 2–6). Al2O3 was considered as suitable supports for this reaction due to its favorable textural properties and thermal stability. Since then, several non-noble metals/supports have been reported.
A comparison of the most promising catalysts, although unfair (due to the very different conditions reported in the literature), is reported in Table 7. Among these, metal-loaded zeolite (e.g., Ni/dealuminated BEA) and core–shell zeolite catalysts (with different distributions of one or more metals within the core and the shell zeolites of the composite particles) led to promising hydrogen yields and decreased formation of by-products and coke, even at low temperatures (Table 7). In particular, the catalyst developed by Wang et al. [153] demonstrated a high H2 selectivity, low coke formation and high stability (>100 h TOS).

6. Conclusions

The present work reviewed the thermodynamics and most suitable catalysts for hydrogen production via a low-temperature ethanol steam reforming reaction, focusing on the hybrid catalyst using zeolites as supports.
As compared with noble metals, nickel and cobalt ensure a good trade-off between costs and performance in the perspective of an economically feasible process. However, catalyst deactivation by coking is still an issue. Moreover, as suggested by the thermodynamic analysis, methane formation should be minimized to maximize the hydrogen yield. Therefore, active metals with decreased methane formation ability are ideally required. The addition of synergistic active metals to the main active metal, e.g., Ni added with Fe and Cu, promotes the water gas shift, acetaldehyde steam reforming and methane steam reforming, maximizing the yield of hydrogen and CO2. Compared to other supports, zeolites, due to their microporous features, allow the preparation of highly dispersed metallic phases. However, the presence of Bronsted acid sites can also lead to olefin formation and coking as a consequence. On the contrary, using alkali metal-doped zeolites as support coking can be effectively avoided also maximizing the H2 yields. At the same time, among the reviewed zeolites, the 12-member-ring pore system of the BEA zeolite is ideal for the ethanol steam reforming reaction due to its peculiar structure and texture. Further improvements can be achieved using dealuminated BEA (drastically reducing the amount of Bronsted acid sites) and core–shell particles with different distributions of active metals in the core and shell zeolites.

Author Contributions

Conceptualization, F.D. and A.A.; methodology, F.D. and E.G.; software, E.G. and A.M.; formal analysis, F.D., E.G. and G.G. (Gianfranco Giorgianni); resources, F.D. and G.G. (Girolamo Giordano); data curation, F.D. and E.G.; writing—original draft preparation, F.D., E.G. and G.G. (Girolamo Giordano); writing—review and editing; F.D., G.G. (Gianfranco Giorgianni) and M.M; visualization, E.G. and A.M.; supervision A.A., M.M. and G.G. (Girolamo Giordano). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

List of Abbreviations

ASRacetaldehyde steam reforming
EDHethanol dehydrogenation
EDHyethanol dehydration
ESRethanol steam reforming
EtOHethanol
GHSV gas hourly space velocity
MSRmethane steam reforming
n ˙ mole flow (mol s−1)
PEM polymer electrolyte membrane
Sselectivity of a reaction product
SBETspecific surface area (m2 g−1)
Ttemperature (°C)
TOS time on stream (h)
Vmicro volume of micropores (cm3 g−1)
WGSwater gas shift reaction
WHSV weight hourly space velocity
Xconversion of the limiting reactant
Yproduct yield
ΔHr°298Kreaction enthalpy at standard condition (kJ mol−1)
η product yield

References

  1. Sayed, E.T.; Wilberforce, T.; Elsaid, K.; Rabaia, M.K.H.; Abdelkareem, M.A.; Chae, K.; Olabi, A.G. A Critical Review on Environmental Impacts of Renewable Energy Systems and Mitigation Strategies: Wind, Hydro, Biomass and Geothermal. Sci. Total Environ. 2021, 766, 144505. [Google Scholar] [CrossRef]
  2. Wasif, M.; Shahbaz, M.; Sinha, A.; Sengupta, T.; Qin, Q. How Renewable Energy Consumption Contribute to Environmental Quality? The Role of Education in OECD Countries. J. Clean. Prod. 2020, 268, 122149. [Google Scholar] [CrossRef]
  3. Dalena, F.; Senatore, A.; Iulianelli, A.; Di Paola, L.; Basile, M.; Basile, A. Ethanol from Biomass: Future and Perspectives. In Ethanol: Science and Engineering; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 25–59. [Google Scholar]
  4. Rajeswari, S.; Baskaran, D.; Saravanan, P.; Rajasimman, M.; Rajamohan, N.; Vasseghian, Y. Production of Ethanol from Biomass–Recent Research, Scientometric Review and Future Perspectives. Fuel 2022, 317, 123448. [Google Scholar] [CrossRef]
  5. Molino, A.; Migliori, M.; Macrì, D.; Valerio, V.; Villone, A.; Nanna, F.; Iovane, P.; Marino, T. Glucose Gasification in Super-Critical Water Conditions for Both Syngas Production and Green Chemicals with a Continuous Process. Renew. Energy 2016, 91, 451–455. [Google Scholar] [CrossRef]
  6. Dalena, F.; Senatore, A.; Basile, M.; Marino, D.; Basile, A. From Sugars to Ethanol—From Agricultural Wastes to Algal Sources: An Overview. In Second and Third Generation of Feedstocks: The Evolution of Biofuels; Elsevier Inc.: Amsterdam, The Netherlands, 2019; pp. 3–34. ISBN 9780128151624. [Google Scholar]
  7. Yu, Y.; Wu, J.; Ren, X.; Lau, A.; Rezaei, H.; Takada, M.; Bi, X.; Sokhansanj, S. Steam Explosion of Lignocellulosic Biomass for Multiple Advanced Bioenergy Processes: A Review. Renew. Sustain. Energy Rev. 2022, 154, 111871. [Google Scholar] [CrossRef]
  8. Kumar, D.; Singh, B.; Korstad, J. Utilization of Lignocellulosic Biomass by Oleaginous Yeast and Bacteria for Production of Biodiesel and Renewable Diesel. Renew. Sustain. Energy Rev. 2017, 73, 654–671. [Google Scholar] [CrossRef]
  9. Papanikolaou, G.; Lanzafame, P.; Giorgianni, G.; Abate, S.; Perathoner, S.; Centi, G. Highly Selective Bifunctional Ni Zeo-Type Catalysts for Hydroprocessing of Methyl Palmitate to Green Diesel. Catal. Today 2020, 345, 14–21. [Google Scholar] [CrossRef]
  10. Hoang, A.T.; Sirohi, R.; Pandey, A.; Nižetić, S.; Lam, S.S.; Chen, W.-H.; Luque, R.; Thomas, S.; Arıcı, M.; Pham, V.V. Biofuel Production from Microalgae: Challenges and Chances. Phytochem. Rev. 2022, 0123456789. [Google Scholar] [CrossRef]
  11. Dalena, F.; Senatore, A.; Iulianelli, A.; Di Paola, L.; Basile, M.; Basile, A. Ethanol From Biomass. In Ethanol; Elsevier: Amsterdam, The Netherlands, 2019; pp. 25–59. ISBN 9780128114582. [Google Scholar]
  12. Sanchez, N.; Ruiz, R.; Hacker, V.; Cobo, M. Impact of Bioethanol Impurities on Steam Reforming for Hydrogen Production: A Review. Int. J. Hydrogen Energy 2020, 45, 11923–11942. [Google Scholar] [CrossRef]
  13. Giorgianni, G.; Abate, S.; Centi, G.; Perathoner, S.; van Beuzekom, S.; Soo-Tang, S.-H.; Van der Waal, J.C. Effect of the Solvent in Enhancing the Selectivity to Furan Derivatives in the Catalytic Hydrogenation of Furfural. ACS Sustain. Chem. Eng. 2018, 6, 16235–16247. [Google Scholar] [CrossRef]
  14. Wang, Y.; Zhao, D.; Rodríguez-Padrón, D.; Len, C. Recent Advances in Catalytic Hydrogenation of Furfural. Catalysts 2019, 9, 796. [Google Scholar] [CrossRef] [Green Version]
  15. Pizzi, R.; van Putten, R.-J.; Brust, H.; Perathoner, S.; Centi, G.; van der Waal, J. High-Throughput Screening of Heterogeneous Catalysts for the Conversion of Furfural to Bio-Based Fuel Components. Catalysts 2015, 5, 2244–2257. [Google Scholar] [CrossRef] [Green Version]
  16. Rodríguez-Machín, L.; Piloto-Rodríguez, R.; Rubio-González, A.; Iturria-Quintero, P.J.; Ronsse, F. Pretreatment of Sugarcane Residues for Combustion in Biomass Power Stations: A Review. Sugar Tech 2022, 24, 732–745. [Google Scholar] [CrossRef]
  17. Borgogna, A.; Centi, G.; Iaquaniello, G.; Perathoner, S.; Papanikolaou, G.; Salladini, A. Assessment of Hydrogen Production from Municipal Solid Wastes as Competitive Route to Produce Low-Carbon H2. Sci. Total Environ. 2022, 827, 154393. [Google Scholar] [CrossRef]
  18. Catizzone, E.; Migliori, M.; Purita, A.; Giordano, G. Ferrierite vs. Γ-Al2O3: The Superiority of Zeolites in Terms of Water-Resistance in Vapour-Phase Dehydration of Methanol to Dimethyl Ether. J. Energy Chem. 2019, 30, 162–169. [Google Scholar] [CrossRef] [Green Version]
  19. Migliori, M.; Aloise, A.; Catizzone, E.; Giordano, G. Kinetic Analysis of Methanol to Dimethyl Ether Reaction over H-MFI Catalyst. Ind. Eng. Chem. Res. 2014, 53, 14885–14891. [Google Scholar] [CrossRef]
  20. Semelsberger, T.A.; Borup, R.L.; Greene, H.L. Dimethyl Ether (DME) as an Alternative Fuel. J. Power Source 2006, 156, 497–511. [Google Scholar] [CrossRef]
  21. Catizzone, E.; Aloise, A.; Migliori, M.; Giordano, G. Dimethyl Ether Synthesis via Methanol Dehydration: Effect of Zeolite Structure. Appl. Catal. A Gen. 2015, 502, 215–220. [Google Scholar] [CrossRef]
  22. Migliori, M.; Catizzone, E.; Aloise, A.; Bonura, G.; Gómez-Hortigüela, L.; Frusteri, L.; Cannilla, C.; Frusteri, F.; Giordano, G. New Insights about Coke Deposition in Methanol-to-DME Reaction over MOR-, MFI- and FER-Type Zeolites. J. Ind. Eng. Chem. 2018, 68, 196–208. [Google Scholar] [CrossRef]
  23. Migliori, M.; Condello, A.; Dalena, F.; Catizzone, E.; Giordano, G. CuZnZr-Zeolite Hybrid Grains for DME Synthesis: New Evidence on the Role of Metal-Acidic Features on the Methanol Conversion Step. Catalysts 2020, 10, 671. [Google Scholar] [CrossRef]
  24. Catizzone, E.; Giglio, E.; Migliori, M.; Cozzucoli, P.C.; Giordano, G. The Effect of Zeolite Features on the Dehydration Reaction of Methanol to Dimethyl Ether: Catalytic Behaviour and Kinetics. Materials 2020, 13, 5577. [Google Scholar] [CrossRef]
  25. Jhang, S.; Lin, Y.-C.; Chen, K.; Lin, S.; Batterman, S. Evaluation of Fuel Consumption, Pollutant Emissions and Well-to-Wheel GHGs Assessment from a Vehicle Operation Fueled with Bioethanol, Gasoline and Hydrogen. Energy 2020, 209, 118436. [Google Scholar] [CrossRef]
  26. Zabed, H.; Sahu, J.N.; Suely, A.; Boyce, A.N.; Faruq, G. Bioethanol Production from Renewable Sources: Current Perspectives and Technological Progress. Renew. Sustain. Energy Rev. 2017, 71, 475–501. [Google Scholar] [CrossRef]
  27. Kadam, S.A.; Shamzhy, M.V. IR Operando Study of Ethanol Dehydration over MFI Zeolite. Catal. Today 2018, 304, 51–57. [Google Scholar] [CrossRef]
  28. Tarach, K.A.; Tekla, J.; Filek, U.; Szymocha, A.; Tarach, I.; Góra-Marek, K. Alkaline-Acid Treated Zeolite L as Catalyst in Ethanol Dehydration Process. Microporous Mesoporous Mater. 2017, 241, 132–144. [Google Scholar] [CrossRef]
  29. Phung, T.K.; Proietti Hernández, L.; Lagazzo, A.; Busca, G. Dehydration of Ethanol over Zeolites, Silica Alumina and Alumina: Lewis Acidity, Brønsted Acidity and Confinement Effects. Appl. Catal. A Gen. 2015, 493, 77–89. [Google Scholar] [CrossRef]
  30. Zoppi, G.; Pipitone, G.; Pirone, R.; Bensaid, S. Aqueous Phase Reforming Process for the Valorization of Wastewater Streams: Application to Different Industrial Scenarios. Catal. Today 2022, 387, 224–236. [Google Scholar] [CrossRef]
  31. Sharma, Y.C.; Kumar, A.; Prasad, R.; Upadhyay, S.N. Ethanol Steam Reforming for Hydrogen Production: Latest and Effective Catalyst Modification Strategies to Minimize Carbonaceous Deactivation. Renew. Sustain. Energy Rev. 2017, 74, 89–103. [Google Scholar] [CrossRef]
  32. Ni, M.; Leung, D.Y.C.; Leung, M.K.H. A Review on Reforming Bio-Ethanol for Hydrogen Production. Int. J. Hydrogen Energy 2007, 32, 3238–3247. [Google Scholar] [CrossRef]
  33. Ogo, S.; Sekine, Y. Recent Progress in Ethanol Steam Reforming Using Non-Noble Transition Metal Catalysts: A Review. Fuel Process. Technol. 2020, 199, 106238. [Google Scholar] [CrossRef]
  34. Frusteri, F.; Freni, S.; Spadaro, L.; Chiodo, V.; Bonura, G.; Donato, S.; Cavallaro, S. H2 Production for MC Fuel Cell by Steam Reforming of Ethanol over MgO Supported Pd, Rh, Ni and Co Catalysts. Catal. Commun. 2004, 5, 611–615. [Google Scholar] [CrossRef]
  35. Wang, Z.; Zhang, X.; Rezazadeh, A. Hydrogen Fuel and Electricity Generation from a New Hybrid Energy System Based on Wind and Solar Energies and Alkaline Fuel Cell. Energy Rep. 2021, 7, 2594–2604. [Google Scholar] [CrossRef]
  36. Khan, U.; Yamamoto, T.; Sato, H. An Insight into Potential Early Adopters of Hydrogen Fuel-Cell Vehicles in Japan. Int. J. Hydrogen Energy 2021, 46, 10589–10607. [Google Scholar] [CrossRef]
  37. Eba, H.; Masuzoe, Y.; Sugihara, T.; Yagi, H.; Liu, T. Ammonia Production Using Iron Nitride and Water as Hydrogen Source under Mild Temperature and Pressure. Int. J. Hydrogen Energy 2021, 46, 10642–10652. [Google Scholar] [CrossRef]
  38. Giglio, E.; Vitale, G.; Lanzini, A.; Santarelli, M. Integration between Biomass Gasification and High-Temperature Electrolysis for Synthetic Methane Production. Biomass Bioenergy 2021, 148, 106017. [Google Scholar] [CrossRef]
  39. Giglio, E.; Pirone, R.; Bensaid, S. Dynamic Modelling of Methanation Reactors during Start-up and Regulation in Intermittent Power-to-Gas Applications. Renew. Energy 2021, 170, 1040–1051. [Google Scholar] [CrossRef]
  40. Guzmán, H.; Salomone, F.; Bensaid, S.; Castellino, M.; Russo, N.; Hernández, S. CO2 Conversion to Alcohols over Cu/ZnO Catalysts: Prospective Synergies between Electrocatalytic and Thermocatalytic Routes. ACS Appl. Mater. Interfaces 2022, 14, 517–530. [Google Scholar] [CrossRef] [PubMed]
  41. Marchese, M.; Heikkinen, N.; Giglio, E.; Lanzini, A.; Lehtonen, J.; Reinikainen, M. Kinetic Study Based on the Carbide Mechanism of a Co-Pt/γ-Al2O3 Fischer–Tropsch Catalyst Tested in a Laboratory-Scale Tubular Reactor. Catalysts 2019, 9, 717. [Google Scholar] [CrossRef] [Green Version]
  42. Pein, M.; Neumann, N.C.; Venstrom, L.J.; Vieten, J.; Roeb, M.; Sattler, C. Two-Step Thermochemical Electrolysis: An Approach for Green Hydrogen Production. Int. J. Hydrogen Energy 2021, 46, 24909–24918. [Google Scholar] [CrossRef]
  43. Freire Ordóñez, D.; Shah, N.; Guillén-Gosálbez, G. Economic and Full Environmental Assessment of Electrofuels via Electrolysis and Co-Electrolysis Considering Externalities. Appl. Energy 2021, 286, 116488. [Google Scholar] [CrossRef]
  44. Martinelli, M.; Castro, J.D.; Alhraki, N.; Matamoros, M.E.; Kropf, A.J.; Cronauer, D.C.; Jacobs, G. Effect of Sodium Loading on Pt/ZrO2 during Ethanol Steam Reforming. Appl. Catal. A Gen. 2021, 610, 117947. [Google Scholar] [CrossRef]
  45. Chen, W.-H.; Li, S.; Lim, S.; Chen, Z.; Juan, J.C. Reaction and Hydrogen Production Phenomena of Ethanol Steam Reforming in a Catalytic Membrane Reactor. Energy 2021, 220, 119737. [Google Scholar] [CrossRef]
  46. Zhurka, M.D.; Lemonidou, A.A.; Anderson, J.A.; Kechagiopoulos, P.N. Kinetic Analysis of the Steam Reforming of Ethanol over Ni/SiO2 for the Elucidation of Metal-Dominated Reaction Pathways. React. Chem. Eng. 2018, 3, 883–897. [Google Scholar] [CrossRef] [Green Version]
  47. Grzybek, G.; Góra-Marek, K.; Patulski, P.; Greluk, M.; Rotko, M.; Słowik, G.; Kotarba, A. Optimization of the Potassium Promotion of the Co/α-Al2O3 Catalyst for the Effective Hydrogen Production via Ethanol Steam Reforming. Appl. Catal. A Gen. 2021, 614, 118051. [Google Scholar] [CrossRef]
  48. Ghasemzadeh, K.; Jalilnejad, E.; Sadati Tilebon, S.M. Hydrogen Production Technologies from Ethanol. In Ethanol; Elsevier: Amsterdam, The Netherlands, 2019; pp. 307–340. ISBN 9780128114582. [Google Scholar]
  49. Li, M.-R.; Wang, G.-C. The Mechanism of Ethanol Steam Reforming on the Co0 and Co2+ Sites: A DFT Study. J. Catal. 2018, 365, 391–404. [Google Scholar] [CrossRef]
  50. Kumar, A.; Prasad, R.; Sharma, Y.C. Ethanol Steam Reforming Study over ZSM-5 Supported Cobalt versus Nickel Catalyst for Renewable Hydrogen Generation. Chinese J. Chem. Eng. 2019, 27, 677–684. [Google Scholar] [CrossRef]
  51. Roychowdhury, S.; Mukthar Ali, M.; Dhua, S.; Sundararajan, T.; Ranga Rao, G. Thermochemical Hydrogen Production Using Rh/CeO2/γ-Al2O3 Catalyst by Steam Reforming of Ethanol and Water Splitting in a Packed Bed Reactor. Int. J. Hydrogen Energy 2021, 46, 19254–19269. [Google Scholar] [CrossRef]
  52. Martínez, A.H.; Lopez, E.; Cadús, L.E.; Agüero, F.N. Elucidation of the Role of Support in Rh/Perovskite Catalysts Used in Ethanol Steam Reforming Reaction. Catal. Today 2021, 372, 59–69. [Google Scholar] [CrossRef]
  53. Greluk, M.; Rotko, M.; Słowik, G.; Turczyniak-Surdacka, S. Hydrogen Production by Steam Reforming of Ethanol over Co/CeO2 Catalysts: Effect of Cobalt Content. J. Energy Inst. 2019, 92, 222–238. [Google Scholar] [CrossRef]
  54. Nejat, T.; Jalalinezhad, P.; Hormozi, F.; Bahrami, Z. Hydrogen Production from Steam Reforming of Ethanol over Ni-Co Bimetallic Catalysts and MCM-41 as Support. J. Taiwan Inst. Chem. Eng. 2019, 97, 216–226. [Google Scholar] [CrossRef]
  55. Ferencz, Z.; Varga, E.; Puskás, R.; Kónya, Z.; Baán, K.; Oszkó, A.; Erdőhelyi, A. Reforming of Ethanol on Co/Al2O3 Catalysts Reduced at Different Temperatures. J. Catal. 2018, 358, 118–130. [Google Scholar] [CrossRef]
  56. Lang, L.; Zhao, S.; Yin, X.; Yang, W.; Wu, C. Catalytic Activities of K-Modified Zeolite ZSM-5 Supported Rhodium Catalysts in Low-Temperature Steam Reforming of Bioethanol. Int. J. Hydrogen Energy 2015, 40, 9924–9934. [Google Scholar] [CrossRef]
  57. Vizcaino, A.; Carrero, A.; Calles, J. Hydrogen Production by Ethanol Steam Reforming over Cu–Ni Supported Catalysts. Int. J. Hydrogen Energy 2007, 32, 1450–1461. [Google Scholar] [CrossRef]
  58. Phung, T.K.; Busca, G. Diethyl Ether Cracking and Ethanol Dehydration: Acid Catalysis and Reaction Paths. Chem. Eng. J. 2015, 272, 92–101. [Google Scholar] [CrossRef]
  59. Rossetti, I.; Compagnoni, M.; Finocchio, E.; Ramis, G.; Di Michele, A.; Zucchini, A.; Dzwigaj, S. Syngas Production via Steam Reforming of Bioethanol over Ni–BEA Catalysts: A BTL Strategy. Int. J. Hydrogen Energy 2016, 41, 16878–16889. [Google Scholar] [CrossRef]
  60. Anggoro, D.D.; Oktavianty, H.; Sasongko, S.B.; Buchori, L. Effect of Dealumination on the Acidity of Zeolite Y and the Yield of Glycerol Mono Stearate (GMS). Chemosphere 2020, 257, 127012. [Google Scholar] [CrossRef] [PubMed]
  61. Dias, S.C.L.; Dias, J.A. Effects of the Dealumination Methodology on the FER Zeolite Acidity: A Study with Fractional Factorial Design. Mol. Catal. 2018, 458, 139–144. [Google Scholar] [CrossRef]
  62. Peron, D.V.; Zholobenko, V.L.; de Melo, J.H.S.; Capron, M.; Nuns, N.; de Souza, M.O.; Feris, L.A.; Marcilio, N.R.; Ordomsky, V.V.; Khodakov, A.Y. External Surface Phenomena in Dealumination and Desilication of Large Single Crystals of ZSM-5 Zeolite Synthesized from a Sustainable Source. Microporous Mesoporous Mater. 2019, 286, 57–64. [Google Scholar] [CrossRef]
  63. Li, D.; Li, X.; Gong, J. Catalytic Reforming of Oxygenates: State of the Art and Future Prospects. Chem. Rev. 2016, 116, 11529–11653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Contreras, J.L.; Salmones, J.; Colín-Luna, J.A.; Nuño, L.; Quintana, B.; Córdova, I.; Zeifert, B.; Tapia, C.; Fuentes, G.A. Catalysts for H2 Production Using the Ethanol Steam Reforming (a Review). Int. J. Hydrogen Energy 2014, 39, 18835–18853. [Google Scholar] [CrossRef]
  65. Zanchet, D.; Santos, J.B.O.; Damyanova, S.; Gallo, J.M.R.; Bueno, J.M.C. Toward Understanding Metal-Catalyzed Ethanol Reforming. ACS Catal. 2015, 5, 3841–3863. [Google Scholar] [CrossRef]
  66. Choi, Y.; Liu, P. Understanding of Ethanol Decomposition on Rh(1 1 1) from Density Functional Theory and Kinetic Monte Carlo Simulations. Catal. Today 2011, 165, 64–70. [Google Scholar] [CrossRef]
  67. Sutton, J.E.; Vlachos, D.G. Ethanol Activation on Closed-Packed Surfaces. Ind. Eng. Chem. Res. 2015, 54, 4213–4225. [Google Scholar] [CrossRef]
  68. Jones, G.; Jakobsen, J.G.; Shim, S.S.; Kleis, J.; Andersson, M.P.; Rossmeisl, J.; Abild-Pedersen, F.; Bligaard, T.; Helveg, S.; Hinnemann, B.; et al. First Principles Calculations and Experimental Insight into Methane Steam Reforming over Transition Metal Catalysts. J. Catal. 2008, 259, 147–160. [Google Scholar] [CrossRef]
  69. Wang, S.; He, B.; Tian, R.; Sun, C.; Dai, R.; Li, X.; Wu, X.; An, X.; Xie, X. Ni-Hierarchical Beta Zeolite Catalysts Were Applied to Ethanol Steam Reforming: Effect of Sol Gel Method on Loading Ni and the Role of Hierarchical Structure. Mol. Catal. 2018, 453, 64–73. [Google Scholar] [CrossRef]
  70. Iulianelli, A.; Dalena, F.; Basile, A. Steam Reforming, Preferential Oxidation, and Autothermal Reforming of Ethanol for Hydrogen Production in Membrane Reactors. In Ethanol; Elsevier: Amsterdam, The Netherlands, 2019; pp. 193–213. ISBN 9780128114582. [Google Scholar]
  71. Phung, T.K.; Pham, T.L.M.; Nguyen, A.-N.T.; Vu, K.B.; Giang, H.N.; Nguyen, T.; Huynh, T.C.; Pham, H.D. Effect of Supports and Promoters on the Performance of Ni-Based Catalysts in Ethanol Steam Reforming. Chem. Eng. Technol. 2020, 43, 672–688. [Google Scholar] [CrossRef]
  72. Zheng, Z.; Sun, C.; Dai, R.; Wang, S.; Wu, X.; An, X.; Wu, Z.; Xie, X. Ethanol Steam Reforming on Ni-Based Catalysts: Effect of Cu and Fe Addition on the Catalytic Activity and Resistance to Deactivation. Energy Fuels 2017, 31, 3091–3100. [Google Scholar] [CrossRef]
  73. Vicente, J.; Ereña, J.; Montero, C.; Azkoiti, M.J.; Bilbao, J.; Gayubo, A.G. Reaction Pathway for Ethanol Steam Reforming on a Ni/SiO2 Catalyst Including Coke Formation. Int. J. Hydrogen Energy 2014, 39, 18820–18834. [Google Scholar] [CrossRef]
  74. Kwak, B.S.; Lee, G.; Park, S.M.; Kang, M. Effect of MnOx in the Catalytic Stabilization of Co2MnO4 Spinel during the Ethanol Steam Reforming Reaction. Appl. Catal. A Gen. 2015, 503, 165–175. [Google Scholar] [CrossRef]
  75. Greluk, M.; Gac, W.; Rotko, M.; Słowik, G.; Turczyniak-Surdacka, S. Co/CeO2 and Ni/CeO2 Catalysts for Ethanol Steam Reforming: Effect of the Cobalt/Nickel Dispersion on Catalysts Properties. J. Catal. 2021, 393, 159–178. [Google Scholar] [CrossRef]
  76. Zhang, H.; Fan, Y.F.; Huan, Y.H.; Yue, M.B. Dry-Gel Synthesis of Shaped Transition-Metal-Doped M-MFI (M = Ti, Fe, Cr, Ni) Zeolites by Using Metal-Occluded Zeolite Seed Sol as a Directing Agent. Microporous Mesoporous Mater. 2016, 231, 178–185. [Google Scholar] [CrossRef] [Green Version]
  77. Inokawa, H.; Maeda, M.; Nishimoto, S. Synthesis of Nickel Nanoparticles with Excellent Thermal Stability in Micropores of Zeolite. Int. J. Hydrogen Energy 2013, 38, 13579–13586. [Google Scholar] [CrossRef]
  78. Bozbag, S.E.; Zhang, L.C.; Aindow, M.; Erkey, C. Carbon Aerogel Supported Nickel Nanoparticles and Nanorods Using Supercritical Deposition. J. Supercrit. Fluids 2012, 66, 265–273. [Google Scholar] [CrossRef]
  79. Isarapakdeetham, S.; Kim-Lohsoontorn, P.; Wongsakulphasatch, S.; Kiatkittipong, W.; Laosiripojana, N.; Gong, J.; Assabumrungrat, S. Hydrogen Production via Chemical Looping Steam Reforming of Ethanol by Ni-Based Oxygen Carriers Supported on CeO2 and La2O3 Promoted Al2O3. Int. J. Hydrogen Energy 2020, 45, 1477–1491. [Google Scholar] [CrossRef]
  80. Zhang, L.; Liu, J.; Li, W. Ethanol Steam Reforming over Ni-Cu/Al2O3-MyOz (M = Si, La, Mg, and Zn) Catalysts. J. Nat. Gas Chem. 2009, 18, 55–65. [Google Scholar] [CrossRef]
  81. Zhurka, M.D.; Lemonidou, A.A.; Kechagiopoulos, P.N. Elucidation of Metal and Support Effects during Ethanol Steam Reforming over Ni and Rh Based Catalysts Supported on (CeO2)-ZrO2-La2O3. Catal. Today 2021, 368, 161–172. [Google Scholar] [CrossRef]
  82. Garbarino, G.; Wang, C.; Valsamakis, I.; Chitsazan, S.; Riani, P.; Finocchio, E.; Flytzani-Stephanopoulos, M.; Busca, G. A Study of Ni/Al2O3 and Ni–La/Al2O3 Catalysts for the Steam Reforming of Ethanol and Phenol. Appl. Catal. B Environ. 2015, 174–175, 21–34. [Google Scholar] [CrossRef] [Green Version]
  83. Song, J.H.; Yoo, S.; Yoo, J.; Park, S.; Gim, M.Y.; Kim, T.H.; Song, I.K. Hydrogen Production by Steam Reforming of Ethanol over Ni/Al2O3-La2O3 Xerogel Catalysts. Mol. Catal. 2017, 434, 123–133. [Google Scholar] [CrossRef]
  84. Campos-Skrobot, F.C.; Rizzo-Domingues, R.C.P.; Fernandes-Machado, N.R.C.; Cantão, M.P. Novel Zeolite-Supported Rhodium Catalysts for Ethanol Steam Reforming. J. Power Source 2008, 183, 713–716. [Google Scholar] [CrossRef]
  85. Gac, W.; Greluk, M.; Słowik, G.; Millot, Y.; Valentin, L.; Dzwigaj, S. Effects of Dealumination on the Performance of Ni-Containing BEA Catalysts in Bioethanol Steam Reforming. Appl. Catal. B Environ. 2018, 237, 94–109. [Google Scholar] [CrossRef]
  86. Bizkarra, K.; Barrio, V.L.; Gartzia-Rivero, L.; Bañuelos, J.; López-Arbeloa, I.; Cambra, J.F. Hydrogen Production from a Model Bio-Oil/Bio-Glycerol Mixture through Steam Reforming Using Zeolite L Supported Catalysts. Int. J. Hydrogen Energy 2019, 44, 1492–1504. [Google Scholar] [CrossRef]
  87. Rodriguez-Gomez, A.; Caballero, A. Bimetallic Ni-Co/SBA-15 Catalysts for Reforming of Ethanol: How Cobalt Modifies the Nickel Metal Phase and Product Distribution. Mol. Catal. 2018, 449, 122–130. [Google Scholar] [CrossRef] [Green Version]
  88. Lindo, M.; Vizcaíno, A.J.; Calles, J.A.; Carrero, A. Ethanol Steam Reforming on Ni/Al-SBA-15 Catalysts: Effect of the Aluminium Content. Int. J. Hydrogen Energy 2010, 35, 5895–5901. [Google Scholar] [CrossRef]
  89. Bayram, B.; Soykal, I.I.; Von Deak, D.; Miller, J.T.; Ozkan, U.S. Ethanol Steam Reforming over Co-Based Catalysts: Investigation of Cobalt Coordination Environment under Reaction Conditions. J. Catal. 2011, 284, 77–89. [Google Scholar] [CrossRef]
  90. 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]
  91. Pinton, N.; Vidal, M.V.; Signoretto, M.; Martínez-Arias, A.; Cortés Corberán, V. Ethanol Steam Reforming on Nanostructured Catalysts of Ni, Co and CeO2: Influence of Synthesis Method on Activity, Deactivation and Regenerability. Catal. Today 2017, 296, 135–143. [Google Scholar] [CrossRef]
  92. Rodriguez-Gomez, A.; Holgado, J.P.; Caballero, A. Cobalt Carbide Identified as Catalytic Site for the Dehydrogenation of Ethanol to Acetaldehyde. ACS Catal. 2017, 7, 5243–5247. [Google Scholar] [CrossRef]
  93. Gonçalves, A.A.S.; Faustino, P.B.; Assaf, J.M.; Jaroniec, M. One-Pot Synthesis of Mesoporous Ni-Ti-Al Ternary Oxides: Highly Active and Selective Catalysts for Steam Reforming of Ethanol. ACS Appl. Mater. Interfaces 2017, 9, 6079–6092. [Google Scholar] [CrossRef]
  94. Słowik, G.; Greluk, M.; Rotko, M.; Machocki, A. Evolution of the Structure of Unpromoted and Potassium-Promoted Ceria-Supported Nickel Catalysts in the Steam Reforming of Ethanol. Appl. Catal. B Environ. 2018, 221, 490–509. [Google Scholar] [CrossRef]
  95. Riani, P.; Garbarino, G.; Canepa, F.; Busca, G. Cobalt Nanoparticles Mechanically Deposited on α-Al2O3: A Competitive Catalyst for the Production of Hydrogen through Ethanol Steam Reforming. J. Chem. Technol. Biotechnol. 2019, 94, 538–546. [Google Scholar] [CrossRef]
  96. Chen, M.; Wang, C.; Wang, Y.; Tang, Z.; Yang, Z.; Zhang, H.; Wang, J. Hydrogen Production from Ethanol Steam Reforming: Effect of Ce Content on Catalytic Performance of Co/Sepiolite Catalyst. Fuel 2019, 247, 344–355. [Google Scholar] [CrossRef]
  97. Parlett, C.M.A.; Aydin, A.; Durndell, L.J.; Frattini, L.; Isaacs, M.A.; Lee, A.F.; Liu, X.; Olivi, L.; Trofimovaite, R.; Wilson, K.; et al. Tailored Mesoporous Silica Supports for Ni Catalysed Hydrogen Production from Ethanol Steam Reforming. Catal. Commun. 2017, 91, 76–79. [Google Scholar] [CrossRef] [Green Version]
  98. da Silva, A.L.M.; den Breejen, J.P.; Mattos, L.V.; Bitter, J.H.; de Jong, K.P.; Noronha, F.B. Cobalt Particle Size Effects on Catalytic Performance for Ethanol Steam Reforming – Smaller Is Better. J. Catal. 2014, 318, 67–74. [Google Scholar] [CrossRef]
  99. Kourtelesis, M.; Verykios, X. Stability of Pt/γ-Al2O3 Catalyst for the Low Temperature Steam Reforming of Ethanol & Acetaldehyde. Effect of Carrier Modification with Ca. Mater. Today Proc. 2018, 5, 27406–27415. [Google Scholar] [CrossRef]
  100. Wu, Y.; Chung, W.; Chang, M. Modification of Ni/γ-Al2O3 Catalyst with Plasma for Steam Reforming of Ethanol to Generate Hydrogen. Int. J. Hydrogen Energy 2015, 40, 8071–8080. [Google Scholar] [CrossRef]
  101. Lucredio, A.F.; Bellido, J.D.A.; Zawadzki, A.; Assaf, E.M. Co Catalysts Supported on SiO2 and γ-Al2O3 Applied to Ethanol Steam Reforming: Effect of the Solvent Used in the Catalyst Preparation Method. Fuel 2011, 90, 1424–1430. [Google Scholar] [CrossRef]
  102. Kim, D.; Sub, B.; Min, B.; Kang, M. Applied Surface Science Characterization of Ni and W Co-Loaded SBA-15 Catalyst and Its Hydrogen Production Catalytic Ability on Ethanol Steam Reforming Reaction. Appl. Surf. Sci. 2015, 332, 736–746. [Google Scholar] [CrossRef]
  103. He, S.; He, S.; Zhang, L.; Li, X.; Wang, J.; He, D.; Lu, J.; Luo, Y. Hydrogen Production by Ethanol Steam Reforming over Ni/SBA-15 Mesoporous Catalysts: Effect of Au Addition. Catal. Today 2015, 258, 162–168. [Google Scholar] [CrossRef]
  104. Bussi, J.; Musso, M.; Veiga, S.; Bespalko, N.; Faccio, R.; Roger, A. Ethanol Steam Reforming over NiLaZr and NiCuLaZr Mixed Metal Oxide Catalysts. Catal. Today 2013, 213, 42–49. [Google Scholar] [CrossRef]
  105. Coleman, L.J.I.; Epling, W.; Hudgins, R.R.; Croiset, E. Ni/Mg–Al Mixed Oxide Catalyst for the Steam Reforming of Ethanol. Appl. Catal. A Gen. 2009, 363, 52–63. [Google Scholar] [CrossRef]
  106. Marino, A.; Aloise, A.; Hernando, H.; Fermoso, J.; Cozza, D.; Giglio, E.; Migliori, M.; Pizarro, P.; Giordano, G.; Serrano, D.P. ZSM-5 Zeolites Performance Assessment in Catalytic Pyrolysis of PVC-Containing Real WEEE Plastic Wastes. Catal. Today 2022, 390–391, 210–220. [Google Scholar] [CrossRef]
  107. Cychosz, K.A.; Guillet-Nicolas, R.; García-Martínez, J.; Thommes, M. Recent Advances in the Textural Characterization of Hierarchically Structured Nanoporous Materials. Chem. Soc. Rev. 2017, 46, 389–414. [Google Scholar] [CrossRef] [PubMed]
  108. Centi, G.; Perathoner, S.; Pino, F.; Arrigo, R.; Giordano, G.; Katovic, A.; Pedulà, V. Performances of Fe-[Al, B]MFI Catalysts in Benzene Hydroxylation with N2O: The Role of Zeolite Defects as Host Sites for Highly Active Iron Species. Catal. Today 2005, 110, 211–220. [Google Scholar] [CrossRef]
  109. Katovic, A.; Giordano, G.; Bonelli, B.; Onida, B.; Garrone, E.; Lentz, P.; Nagy, J.B. Preparation and Characterization of Mesoporous Molecular Sieves Containing Al, Fe or Zn. Microporous Mesoporous Mater. 2001, 44–45, 275–281. [Google Scholar] [CrossRef]
  110. Sandoval-Díaz, L.-E.; González-Amaya, J.-A.; Trujillo, C.-A. General Aspects of Zeolite Acidity Characterization. Microporous Mesoporous Mater. 2015, 215, 229–243. [Google Scholar] [CrossRef]
  111. Weitkamp, J. Zeolites and Catalysis. Solid State Ionics 2000, 131, 175–188. [Google Scholar] [CrossRef]
  112. Maesen, T. The Zeolite Scene—An Overview. In Introduction to Zeolite Science and Practice; Čejka, J., van Bekkum, H., Corma, A., Schüth, F., Eds.; Elsevier: Amsterdam, The Netherlands, 2007; Volume 168, pp. 1–12. ISBN 0444530630. [Google Scholar]
  113. Milliken, T.H.; Mills, G.A.; Oblad, A.G. The Chemical Characteristics and Structure of Cracking Catalysts. Discuss. Faraday Soc. 1950, 8, 279–290. [Google Scholar] [CrossRef]
  114. Busca, G. Acidity and Basicity of Zeolites: A Fundamental Approach. Microporous Mesoporous Mater. 2017, 254, 3–16. [Google Scholar] [CrossRef]
  115. Madeira, F.F.; Gnep, N.S.; Magnoux, P.; Maury, S.; Cadran, N. Ethanol Transformation over HFAU, HBEA and HMFI Zeolites Presenting Similar Brønsted Acidity. Appl. Catal. A Gen. 2009, 367, 39–46. [Google Scholar] [CrossRef]
  116. Santi, D.; Rabl, S.; Calemma, V.; Dyballa, M.; Hunger, M.; Weitkamp, J. Effect of Noble Metals on the Strength of Brønsted Acid Sites in Bifunctional Zeolites. ChemCatChem 2013, 5, 1524–1530. [Google Scholar] [CrossRef]
  117. Aloise, A.; Marino, A.; Dalena, F.; Giorgianni, G.; Migliori, M.; Frusteri, L.; Cannilla, C.; Bonura, G.; Frusteri, F.; Giordano, G. Desilicated ZSM-5 Zeolite: Catalytic Performances Assessment in Methanol to DME Dehydration. Microporous Mesoporous Mater. 2020, 302, 110198. [Google Scholar] [CrossRef]
  118. Dalena, F.; Giglio, E.; Giorgianni, G.; Cozza, D.; Marino, A.; Aloise, A. DME Production via Methanol Dehydration with H Form and Desilicated ZSM-5 Type Zeolitic Catalysts: Study on the Correlation between Acid Sites and Conversion. Chem. Eng. Trans. 2021, 84, 211–216. [Google Scholar] [CrossRef]
  119. Rossetti, I.; Lasso, J.; Nichele, V.; Signoretto, M.; Finocchio, E.; Ramis, G.; Di Michele, A. Silica and Zirconia Supported Catalysts for the Low-Temperature Ethanol Steam Reforming. Appl. Catal. B Environ. 2014, 150–151, 257–267. [Google Scholar] [CrossRef]
  120. Gołąbek, K.; Tarach, K.A.; Filek, U.; Góra-Marek, K. Ethylene Formation by Dehydration of Ethanol over Medium Pore Zeolites. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 192, 464–472. [Google Scholar] [CrossRef]
  121. Kuz’mina, R.I.; Frolov, M.P.; Liventsev, V.T.; Vetrova, T.K.; Kovnev, A. V Development of Zeolite-Containing Reforming Catalysts. Catal. Ind. 2010, 2, 329–333. [Google Scholar] [CrossRef]
  122. Čejka, J.; van Bekkum, H.; Corma, A.; Schüth, F. (Eds.) Introduction to Zeolite Science and Practice, 3rd ed.; Elsevier B.V.: Amsterdam, The Netherlands, 2007. [Google Scholar]
  123. Da Costa-Serra, J.F.; Navarro, M.T.; Rey, F.; Chica, A. Bioethanol Steam Reforming on Ni-Based Modified Mordenite. Effect of Mesoporosity, Acid Sites and Alkaline Metals. Int. J. Hydrogen Energy 2012, 37, 7101–7108. [Google Scholar] [CrossRef]
  124. Rokicińska, A.; Drozdek, M.; Dudek, B.; Gil, B.; Michorczyk, P.; Brouri, D.; Dzwigaj, S.; Kuśtrowski, P. Cobalt-Containing BEA Zeolite for Catalytic Combustion of Toluene. Appl. Catal. B Environ. 2017, 212, 59–67. [Google Scholar] [CrossRef] [Green Version]
  125. Newsam, J.M.; Treacy, M.M.J.; Koetsier, W.T.; De Gruyter, C.B. Structural Characterization of Zeolite Beta. Proc. R. Soc. Lond. A. Math. Phys. Sci. 1988, 420, 375–405. [Google Scholar] [CrossRef]
  126. International Zeolite Association (IZA)—Database of Zeolite Structures. Available online: http://www.iza-structure.org/databases/ (accessed on 28 April 2022).
  127. Pinheiro, A.N.; Valentini, A.; Sasaki, J.M.; Oliveira, A.C. Highly Stable Dealuminated Zeolite Support for the Production of Hydrogen by Dry Reforming of Methane. Appl. Catal. A Gen. 2009, 355, 156–168. [Google Scholar] [CrossRef]
  128. Che, Q.; Yang, M.; Wang, X.; Yang, Q.; Rose Williams, L.; Yang, H.; Zou, J.; Zeng, K.; Zhu, Y.; Chen, Y.; et al. Influence of Physicochemical Properties of Metal Modified ZSM-5 Catalyst on Benzene, Toluene and Xylene Production from Biomass Catalytic Pyrolysis. Bioresour. Technol. 2019, 278, 248–254. [Google Scholar] [CrossRef]
  129. Shi, W.; Yi, B.; Hou, M.; Shao, Z. The Effect of H2S and CO Mixtures on PEMFC Performance. Int. J. Hydrogen Energy 2007, 32, 4412–4417. [Google Scholar] [CrossRef]
  130. Baschuk, J.J.; Li, X. Carbon Monoxide Poisoning of Proton Exchange Membrane Fuel Cells. Int. J. Energy Res. 2001, 25, 695–713. [Google Scholar] [CrossRef]
  131. Tian, R.; Wang, S.; Lian, C.; Wu, X.; An, X.; Xie, X. Synthesis of the Hierarchical Fe-Substituted Porous HBeta Zeolite and the Exploration of Its Catalytic Performance. J. Fuel Chem. Technol. 2019, 47, 1476–1485. [Google Scholar] [CrossRef]
  132. Centi, G.; Perathoner, S.; Arrigo, R.; Giordano, G.; Katovic, A.; Pedulà, V. Characterization and Reactivity of Fe-[Al,B]MFI Catalysts for Benzene Hydroxylation with N2O. Appl. Catal. A Gen. 2006, 307, 30–41. [Google Scholar] [CrossRef]
  133. Centi, G.; Genovese, C.; Giordano, G.; Katovic, A.; Perathoner, S. Performance of Fe–BEA Catalysts for the Selective Hydroxylation of Benzene with N2O. Catal. Today 2004, 91–92, 17–26. [Google Scholar] [CrossRef]
  134. Zheng, Z.; Sun, C.; Dai, R.; Wang, S.; Wu, X.; An, X.; Xie, X. Organotemplate-Free Synthesis of Hollow Beta Zeolite Supported Pt-Based Catalysts for Low-Temperature Ethanol Steam Reforming. Catal. Sci. Technol. 2016, 6, 6472–6475. [Google Scholar] [CrossRef]
  135. Khaleque, A.; Alam, M.M.; Hoque, M.; Mondal, S.; Haider, J.B.; Xu, B.; Johir, M.A.H.; Karmakar, A.K.; Zhou, J.L.; Ahmed, M.B.; et al. Zeolite Synthesis from Low-Cost Materials and Environmental Applications: A Review. Environ. Adv. 2020, 2, 100019. [Google Scholar] [CrossRef]
  136. Lutz, W. Zeolite Y: Synthesis, Modification, and Properties—A Case Revisited. Adv. Mater. Sci. Eng. 2014, 2014, 1–20. [Google Scholar] [CrossRef] [Green Version]
  137. Sang, S.; Liu, Z.; Tian, P.; Liu, Z.; Qu, L.; Zhang, Y. Synthesis of Small Crystals Zeolite NaY. Mater. Lett. 2006, 60, 1131–1133. [Google Scholar] [CrossRef]
  138. Domoroshchina, E.N.; Chernyshev, V.V.; Kuz’micheva, G.M.; Dorokhov, A.V.; Pirutko, L.V.; Kravchenko, G.V.; Chumakov, R.B. Changing the Characteristics and Properties of Zeolite Y and Nano-Anatase in the Formation of a Nano-Anatase/Y Composite with Improved Photocatalytic and Adsorption Properties. Appl. Nanosci. 2018, 8, 19–31. [Google Scholar] [CrossRef] [Green Version]
  139. Inokawa, H.; Nishimoto, S.; Kameshima, Y.; Miyake, M. Difference in the Catalytic Activity of Transition Metals and Their Cations Loaded in Zeolite Y for Ethanol Steam Reforming. Int. J. Hydrogen Energy 2010, 35, 11719–11724. [Google Scholar] [CrossRef]
  140. Inokawa, H.; Nishimoto, S.; Kameshima, Y.; Miyake, M. Promotion of H2 Production from Ethanol Steam Reforming by Zeolite Basicity. Int. J. Hydrogen Energy 2011, 36, 15195–15202. [Google Scholar] [CrossRef]
  141. Sub, B.; Su, J.; Sung, J.; Choi, B.; Jung, M.; Kang, M. Hydrogen-Rich Gas Production from Ethanol Steam Reforming over Ni/Ga/Mg/Zeolite Y Catalysts at Mild Temperature. Appl. Energy 2011, 88, 4366–4375. [Google Scholar] [CrossRef]
  142. Chica, A.; Sayas, S. Effective and Stable Bioethanol Steam Reforming Catalyst Based on Ni and Co Supported on All-Silica Delaminated ITQ-2 Zeolite. Catal. Today 2009, 146, 37–43. [Google Scholar] [CrossRef]
  143. Da Costa-Serra, J.F.; Chica, A. Bioethanol Steam Reforming on Co/ITQ-18 Catalyst: Effect of the Crystalline Structure of the Delaminated Zeolite ITQ-18. Int. J. Hydrogen Energy 2011, 36, 3862–3869. [Google Scholar] [CrossRef]
  144. Giordano, G.; Migliori, M.; Ferrarelli, G.; Giorgianni, G.; Dalena, F.; Peng, P.; Debost, M.; Boullay, P.; Liu, Z.; Guo, H.; et al. Passivated Surface of High Aluminum Containing ZSM-5 by Silicalite-1: Synthesis and Application in Dehydration Reaction. ACS Sustain. Chem. Eng. 2022, 10, 4839–4848. [Google Scholar] [CrossRef]
  145. Ferrarelli, G.; Giordano, G.; Migliori, M. ZSM-5@Sil-1 Core Shell: Effect of Synthesis Method over Textural and Catalytic Properties. Catal. Today 2022, 390–391, 176–184. [Google Scholar] [CrossRef]
  146. Deng, Y.; Zhou, W.; Lv, H.; Zhang, Y.; Au, C.; Yin, S. Synthesis of HZSM-5@silicalite-1 Core–Shell Composite and Its Catalytic Application in the Generation of p-Xylene by Methylation of Toluene with Methyl Bromide. RSC Adv. 2014, 4, 37296–37301. [Google Scholar] [CrossRef]
  147. Li, N.; Zhang, Y.; Chen, L.; Au, C.; Yin, S. Synthesis and Application of HZSM-5@silicalite-1 Core–Shell Composites for the Generation of Light Olefins from CH3Br. Microporous Mesoporous Mater. 2016, 227, 76–80. [Google Scholar] [CrossRef]
  148. Li, X.; Zheng, Z.; Wang, S.; Sun, C.; Dai, R.; Wu, X.; An, X.; Xie, X. Preparation and Characterization of Core–Shell Composite Zeolite BEA@MFI and Their Catalytic Properties in ESR. Catal. Lett. 2019, 149, 766–777. [Google Scholar] [CrossRef]
  149. Bouizi, Y.; Diaz, I.; Rouleau, L.; Valtchev, V.P. Core-Shell Zeolite Microcomposites. Adv. Funct. Mater. 2005, 15, 1955–1960. [Google Scholar] [CrossRef]
  150. Kwak, B.S.; Kim, J.; Kang, M. Hydrogen Production from Ethanol Steam Reforming over Core–Shell Structured NixOy–, FexOy–, and CoxOy–Pd Catalysts. Int. J. Hydrogen Energy 2010, 35, 11829–11843. [Google Scholar] [CrossRef]
  151. Dai, R.; Zheng, Z.; Sun, C.; Li, X.; Wang, S.; Wu, X.; An, X.; Xie, X. Pt Nanoparticles Encapsulated in a Hollow Zeolite Microreactor as a Highly Active and Stable Catalyst for Low-Temperature Ethanol Steam Reforming. Fuel 2018, 214, 88–97. [Google Scholar] [CrossRef]
  152. Zheng, Z.; Yang, D.; Li, T.; Yin, X.; Wang, S.; Wu, X.; An, X.; Xie, X. A Novel BEA-Type Zeolite Core–Shell Multiple Catalyst for Hydrogen-Rich Gas Production from Ethanol Steam Reforming. Catal. Sci. Technol. 2016, 6, 5427–5439. [Google Scholar] [CrossRef]
  153. Wang, S.; He, B.; Tian, R.; Wu, X.; An, X.; Liu, Y.; Su, J.; Yu, Z.; Xie, X. Novel Core-Shell-like Ni-Supported Hierarchical Beta Zeolite Catalysts on Bioethanol Steam Reforming. Int. J. Hydrogen Energy 2020, 45, 16409–16420. [Google Scholar] [CrossRef]
  154. Sun, C.; Zheng, Z.; Wang, S.; Li, X.; Wu, X.; An, X.; Xie, X. Yolk-Shell Structured Pt-CeO2@Ni-SiO2 as an Efficient Catalyst for Enhanced Hydrogen Production from Ethanol Steam Reforming. Ceram. Int. 2018, 44, 1438–1442. [Google Scholar] [CrossRef]
  155. Pizzolitto, C.; Menegazzo, F.; Ghedini, E.; Innocenti, G.; Di Michele, A.; Cruciani, G.; Cavani, F.; Signoretto, M. Increase of Ceria Redox Ability by Lanthanum Addition on Ni Based Catalysts for Hydrogen Production. ACS Sustain. Chem. Eng. 2018, 6, 13867–13876. [Google Scholar] [CrossRef]
  156. Dai, R.; Zheng, Z.; Shi, K.; Wu, X.; An, X.; Xie, X. A Multifunctional Core–Shell Nanoreactor with Unique Features of Sintering Resistance for High-Performance Ethanol Steam Reforming Reaction. Fuel 2021, 287, 119514. [Google Scholar] [CrossRef]
Figure 1. ESR equilibrium composition, considering only the components in reaction (1): (a) effect of temperature on equilibrium mixture at 1.5 bar and (b) thermodynamic hydrogen fractional yield at different temperatures and pressures. The equilibrium was calculated via Gibbs free energy minimization.
Figure 1. ESR equilibrium composition, considering only the components in reaction (1): (a) effect of temperature on equilibrium mixture at 1.5 bar and (b) thermodynamic hydrogen fractional yield at different temperatures and pressures. The equilibrium was calculated via Gibbs free energy minimization.
Catalysts 12 00617 g001
Figure 2. ESR equilibrium with stoichiometric inlet mixture: (a) effect of temperature on outlet composition at 1.5 bar and (b) H2 outlet molar fraction vs. temperature at different pressures. The equilibrium was calculated via Gibbs free energy minimization.
Figure 2. ESR equilibrium with stoichiometric inlet mixture: (a) effect of temperature on outlet composition at 1.5 bar and (b) H2 outlet molar fraction vs. temperature at different pressures. The equilibrium was calculated via Gibbs free energy minimization.
Catalysts 12 00617 g002
Figure 3. ESR reaction pathways as a function of temperature for metals such as nickel and cobalt (adapted from [65]).
Figure 3. ESR reaction pathways as a function of temperature for metals such as nickel and cobalt (adapted from [65]).
Catalysts 12 00617 g003
Figure 4. Reactions involved in the ethanol steam reforming and influence of the acid sites (adapted from [64]).
Figure 4. Reactions involved in the ethanol steam reforming and influence of the acid sites (adapted from [64]).
Catalysts 12 00617 g004
Figure 5. Schematic reaction pathway of the ESR reaction over the Cu/Fe/Ni_BEA system (adapted from [72]).
Figure 5. Schematic reaction pathway of the ESR reaction over the Cu/Fe/Ni_BEA system (adapted from [72]).
Catalysts 12 00617 g005
Figure 6. Schematic representation of the ESR reaction over the core–shell system (adapted from [151]).
Figure 6. Schematic representation of the ESR reaction over the core–shell system (adapted from [151]).
Catalysts 12 00617 g006
Table 1. Recent ESR studies on nickel, cobalt and bimetallic Ni-Co catalysts: operating condition, EtOH conversion and hydrogen yield.
Table 1. Recent ESR studies on nickel, cobalt and bimetallic Ni-Co catalysts: operating condition, EtOH conversion and hydrogen yield.
CatalystSteam-to-Carbon Ratio (S/C)T (°C)Space VelocityEthanol ConversionH2 YieldRef.
Ni/SiO23500W/FEtOH = 91.88 gcat s g EtOH−1≈35%≈20%[46]
Ni/Al2O3-La2O3345023,140 mL h−1 gcat−1100%62%[83]
10 Ni/TiO2-Al2O31.5500WHSV = 2773 h−193%75%[93]
10Ni/CeO2642060,000 mL gcat−1 h−1100%68%[94]
10Ni/SBA-151.8550060,000 mL gcat−1 h−169%41%[87]
17Co/α-Al2O33500GHSV= 51700 h−186%64%[95]
29Co/CeO2650060,000 mL gcat−1 h−1100%94%[53]
29Co/CeO2350060,000 mL gcat−1 h−1≈85%≈80%[53]
10Co-0.3Ce/SEP (Sepiolite)3600WHSV = 21.5 h−191%69%[96]
10Co/SEP (Sepiolite)3600WHSV = 21.5 h−154%34%[96]
10Co/SBA-151.8550060,000 mL gcat−1 h−189%49%[92]
10Co/Al2O31.555072,000 mL gcat−1 h−199%86%[55]
9Ni-1Co/MCM-412.54909000 mL gcat−1 h−190%80%[54]
5Ni-5Co/MCM-412.54909000 mL gcat−1 h−1≈80%≈65%[54]
1Ni-9Co/MCM-412.54909000 mL gcat−1 h−1≈75%≈60%[54]
(Ni, Co) NPs (Nanoparticles)
(Ni/Co = 0.26)
3500GHSV = 324,000 h−1100%87%[90]
8Ni-2Co/SBA-151.8550060,000 mL gcat−1 h−186%53%[87]
5Ni-5Co/SBA-151.8550060,000 mL gcat−1 h−168%43%[87]
2Ni-8Co/SBA-151.8550060,000 mL gcat−1 h−159%42%[87]
20Ni-20Co/CeO23500W/F => 0.12 gcat h molEtOH−1≈85%≈55%[91]
Table 2. ZSM-5 and MOR support for various types of metal catalysts for the ESR reaction.
Table 2. ZSM-5 and MOR support for various types of metal catalysts for the ESR reaction.
MetalSupportExperimental ConditionΧEtOH (%)Selectivity (%)Ref.
TypeSi/AlSBET (m2·g−1)Vmicro (cm3·g−1)T (°C)Space VelocityH2COCO2CH4C2H4
Co (10 wt.-%)ZSM-5504400.281500WHSV = 35.4 ggas h−1 gcat−1~97~45~1~1~8~90[50]
600~100~60~40~20~6~25
Ni (10 wt.-%)ZSM-550345.50.190500~97~47~5~16~13~43
600~100~72~20~24~10~0
Rh (1 wt.-%)ZSM-548.13050.083300n.a.79282918482[56]
400~10033347461
Ni (19.9 wt.-%)MOR103600.18400GHSV = 4700 h−184.324.11.65.82.864.2[123]
Ni (19.4 wt.-%)MOR
(treated)
103400.06400GHSV = 4700 h−197.569.08.24.414.40.6
Conversion and selectivity data of Ref. [50] were extrapolated from the graph.
Table 3. BEA support for various types of metal catalysts for the ESR reaction.
Table 3. BEA support for various types of metal catalysts for the ESR reaction.
MetalSupportExperimental ConditionΧEtOH (%)Selectivity (%)Ref.
TypeSi/AlSBET (m2 g−1)Vmicro (cm3 g−1)T (°C)Space VelocityH2COCO2CH4C2H4
Ni (10 wt.-%)BEA17481.70.48500WHSV ≈ 9.5 gEtOH h−1 gcat−110060330170[85]
Ni (10 wt.-%)BEA 496.80.42500100751055350
Ni (10 wt.-%)BEA1005700.19300WHSV ≈ 7.35 gEtOH h−1 gcat−1≈87≈35≈14≈9≈16n.a.[72]
400≈97≈57≈7≈17≈11n.a.
500100≈68≈5≈20≈7n.a.
Fe (1.5 wt.-%)BEA1005080.21300WHSV ≈ 7.35 gEtOH h−1 gcat−1≈90≈52≈7≈15≈14n.a.
Cu (1.5 wt.-%)400≈99≈68≈3≈21≈8n.a.
Ni (10 wt.-%)500100≈72≈3≈21≈4n.a.
Conversion and selectivity data of Ref. [72] were extrapolated from the graph.
Table 4. Y-type zeolite-supported catalysts for ESR.
Table 4. Y-type zeolite-supported catalysts for ESR.
MetalSupportExperimental ConditionProduct/Ethanol FeedRef.
TypeSi/AlT (°C)Space VelocityH2COCO2CH4C2H4
Ni (9 wt.-%)Na_Y2.7530096.83.80.40.10.20.06[140]
K_Y2.7530096.84.70.50.20.40.01
Cs_Y2.7530096.85.70.70.20.50.01
Ni (4.5 wt.-%)Y (Ni loaded via wet impregnation)2.7530087.00.90.50.50.2[139]
Ni (2.4 wt.-%)Y (Ni loaded via ionic exchange)2.7530080.50.00.00.029.5
Table 5. ITQ-18 and ITQ-2 support for Co and Ni catalysts for the ESR reaction.
Table 5. ITQ-18 and ITQ-2 support for Co and Ni catalysts for the ESR reaction.
CatalystSupportExperimental ConditionΧEtOH (%)Selectivity (%)Ref.
TypeSBET (m2·g−1)Pore Volume (BJH) (cm3·g−1)T (°C)Space VelocityH2COCO2CH4C2H4
Co
(19.9 wt.-%)
ITQ-182930.17500 °CGHSV ≈ 4700 h−197.971.23.119.55.70.1[143]
Co
(20.1 wt.-%)
ITQ-25070.54300 °CGHSV ≈ 4700 h−1≈70≈55≈12≈8~15≈0[142]
400 °C≈86≈62≈4≈15~16≈0
500 °C≈97≈69≈1≈21~9≈0
Ni
(19.5 wt.-%)
ITQ-25170.53300 °CGHSV ≈ 4700 h−1≈77≈50≈18≈2~19≈0
400 °C≈95≈59≈12≈10~18≈0
500 °C≈100≈66≈2≈21~10≈0
Conversion and selectivity data of Ref. [142] were extrapolated from graphs.
Table 6. Core–shell support for the various types of metal catalysts for the ESR reaction.
Table 6. Core–shell support for the various types of metal catalysts for the ESR reaction.
CoreShellSupportExperimental ConditionXEtOH (%)Selectivity (%)Ref.
MetalTypeMetalTypeSBET (m2·g−1)VTOT (cm3·g−1)T (°C)WHSV (gEtOH h−1 gcat−1)H2COCO2CH4
Pt
(1 wt.-%)
(core + shell)
BEAPt
(1 wt.-%)
(core + shell)
SiO25570.653008.5≈98≈64≈4≈19≈10[151]
350≈100≈68≈2≈21≈7
400≈100≈72≈1≈22≈2
Cu
(2.5 wt.-%)
Fe
(2.5 wt.-%)
BEANi
(10 wt.-%)
Al-BEA4280.243007.3≈97≈57≈4≈21≈18[152]
350≈100≈69≈0≈24≈7
400≈100≈67≈2≈28≈2
Ni
(22 wt.-%)
(core + shell)
BEANi
(22 wt.-%)
(core + shell)
BEA295.51.1935029.4≈85≈76≈7≈16≈1[153]
400≈89≈73≈6≈18≈3
450≈92≈72≈8≈17≈2
500≈95≈71≈8≈17≈2
Conversion and selectivity values were extrapolated from graphs.
Table 7. Comparison of ESR catalytic performance among various catalytic systems.
Table 7. Comparison of ESR catalytic performance among various catalytic systems.
Sample NameOperating ConditionsΧEtOH (%)SH2 (%)Ref.
Temperature (°C)Space VelocityEtOH/H2OTOS (h)
Pt-CeO2@Ni-SiO2400WHSV = 8.9 h−11:628100~67[154]
Co-Ni/_La-Ce550WHSV = 2.26 h−11:66090 (100% for the first 20 h)~69[155]
Co/CeO2_N-CA (Citric Acid)420GHSV = 60,000 mL g−1 h−11:1221~60~76[75]
Pt-Cu@Ni-SiO2450WHSV = 7.2 h−11:650100~71[156]
Ni(10)/Ga(30)/Mg(30)_Zeolite Y600WHSV = 6.7 h−11:359100~69[141]
Ni10SiBEA500WHSV = 9.5 gEtOH h−1gcat−11:1222100~65 (~75 for t = 18 h)[85]
10.0 wt% CoxOy@Pd_
Zeolite Y
600GHSV = 16,800 h−11:34510075–100[150]
2.5Fe2.5CuSB@NB
(Si-Beta core and Ni-Beta
shell)
500WHSV = 7.3 h−11:6810071[152]
NiNPs/OH-MBEA400WHSV = 29.4 h−11:5100~93~77[153]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dalena, F.; Giglio, E.; Marino, A.; Aloise, A.; Giorgianni, G.; Migliori, M.; Giordano, G. Steam Reforming of Bioethanol Using Metallic Catalysts on Zeolitic Supports: An Overview. Catalysts 2022, 12, 617. https://doi.org/10.3390/catal12060617

AMA Style

Dalena F, Giglio E, Marino A, Aloise A, Giorgianni G, Migliori M, Giordano G. Steam Reforming of Bioethanol Using Metallic Catalysts on Zeolitic Supports: An Overview. Catalysts. 2022; 12(6):617. https://doi.org/10.3390/catal12060617

Chicago/Turabian Style

Dalena, Francesco, Emanuele Giglio, Alessia Marino, Alfredo Aloise, Gianfranco Giorgianni, Massimo Migliori, and Girolamo Giordano. 2022. "Steam Reforming of Bioethanol Using Metallic Catalysts on Zeolitic Supports: An Overview" Catalysts 12, no. 6: 617. https://doi.org/10.3390/catal12060617

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