Next Article in Journal
Dynamic Ultra-Fast Sorption/Desorption of Indigo Carmine onto/from Versatile Core-Shell Composite Microparticles
Previous Article in Journal
Enzymatic Analysis of Chitin Deacetylases on Crystalline Chitin with Varied Molecular Weights: Insights from Active Pocket Characteristic Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Catalyst Development for Dry Reforming of Methane and Ethanol into Syngas: Recent Advances and Perspectives

by
Manshuk Mambetova
1,2,*,
Moldir Anissova
1,
Laura Myltykbayeva
1,2,
Nursaya Makayeva
1,2,
Kusman Dossumov
1 and
Gaukhar Yergaziyeva
1,2,*
1
Center of Physical Chemical Methods of Research and Analysis, Al-Farabi Kazakh National University, Almaty 050040, Kazakhstan
2
Institute of Combustion Problems, Almaty 050012, Kazakhstan
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2025, 15(19), 10722; https://doi.org/10.3390/app151910722 (registering DOI)
Submission received: 4 September 2025 / Revised: 29 September 2025 / Accepted: 1 October 2025 / Published: 5 October 2025

Abstract

Dry reforming of methane and ethanol is a promising catalytic process for the conversion of carbon dioxide and hydrocarbon feedstocks into synthesis gas (H2/CO), which serves as a key platform for the production of fuels and chemicals. Over the past decade, substantial progress has been achieved in the design of catalysts with enhanced activity and stability under the demanding conditions of these strongly endothermic reactions. This review summarizes the latest developments in catalyst systems for DRM and EDR, including Ni-based catalysts, perovskite-type oxides, MOF-derived materials, and high-entropy alloys. Particular attention is given to strategies for suppressing carbon deposition and preventing metal sintering, such as oxygen vacancy engineering in oxide supports, rare earth and transition metal doping, strong metal–support interactions, and morphological control via core–shell and mesoporous architectures. These approaches have been shown to improve coke resistance, maintain metal dispersion, and extend catalyst lifetimes. The review also highlights emerging concepts such as multifunctional hybrid systems and innovative synthesis methods. By consolidating recent findings, this work provides a comprehensive overview of current progress and future perspectives in catalyst development for DRM and EDR, offering valuable guidelines for the rational design of advanced catalytic materials.

1. Introduction

Global challenges related to climate change, growing energy demand, and the depletion of fossil resources are driving the search for sustainable and environmentally friendly energy production methods. Fossil fuels oil, coal, and natural gas still dominate the global energy mix, accounting for over 80% of total consumption [1]. However, their intensive use is associated with substantial greenhouse gas emissions, particularly carbon dioxide (CO2), which exacerbates anthropogenic impacts on the Earth’s climate system [2,3]. In response to these pressing issues, the Paris Agreement was adopted in 2015 with the aim of limiting the rise in global average temperature to well below 2 °C, while pursuing efforts to restrict it to 1.5 °C by 2100 [4,5]. Achieving these goals requires the development and implementation of technologies capable of simultaneously reducing carbon emissions and ensuring a stable energy supply. In this context, carbon-neutral processes that combine environmental compatibility with economic viability have gained particular importance.
One such strategic direction involves the production of synthesis gas (syngas), a versatile intermediate consisting of carbon monoxide and hydrogen. Syngas serves as a key feedstock for the synthesis of ammonia, methanol, liquid hydrocarbons (including synthetic diesel and kerosene), and is also widely utilized in the production of olefins and aromatic compounds [6,7,8]. The current distribution of syngas utilization worldwide is presented in Figure 1.
According to literature sources, the global annual consumption of syngas is approximately 6000 PJ, representing about 2% of total primary energy demand [10]. Its primary use is concentrated in the chemical industry, with approximately 40% allocated to ammonia synthesis, 25% to methanol production, and about 15% to Fischer-Tropsch synthesis for generating liquid fuels [11]. The remaining portion is utilized in energy generation (10%) and in the production of organic compounds (~5%) such as plastics, solvents, and alcohols.
Traditionally, syngas is produced via steam methane reforming (SMR), partial oxidation (POX), or autothermal reforming (ATR). SMR is the most commonly used method, offering a high hydrogen yield (H2/CO ≈ 3), but it suffers from high energy demand due to the need for superheated steam (>800 °C) and generates significant CO2 emissions [12]. Moreover, SMR requires extensive feedstock pretreatment to remove sulfur compounds and the use of highly durable materials to withstand harsh reaction conditions [13]. These limitations have driven interest in alternative methods, particularly dry reforming of methane (DRM) and dry reforming of ethanol (EDR), where carbon dioxide is used as the oxidant.
The DRM process, first reported by Fischer and Tropsch in 1928, involves the reaction of methane with CO2 to form syngas with a near-stoichiometric H2/CO ratio of approximately 1:1 (CH4 + CO2 → 2CO + 2H2, ΔH° = +247 kJ/mol) [14]. Although the H2/CO ratio from DRM (1:1) is lower than the optimal value for Fischer-Tropsch synthesis (2:1), the resulting syngas can still be adjusted for downstream applications. In addition to its thermodynamic viability, DRM presents environmental advantages through the simultaneous utilization of two potent greenhouse gases CH4 and CO2, making it highly attractive within the framework of carbon circularity and green chemistry [15].
However, practical implementation of DRM requires elevated temperatures (>700 °C) due to the high bond dissociation energies of C–H in methane and C=O in carbon dioxide, which introduces several challenges such as carbon deposition (coking), sintering of active sites, and catalyst deactivation [16,17].
Of additional interest is the EDR, in which ethanol reacts with CO2 to produce syngas according to the reaction: C2H5OH + CO2 → 3CO + 3H2. This process offers multiple advantages. First, ethanol is a renewable feedstock derived from biomass, which contributes to reduced dependence on fossil fuels [18]. Second, the use of CO2 as the oxidant allows for integration into carbon capture and utilization (CCU) schemes. Third, the resulting H2/CO ratio of 1 is ideal for producing high-energy-density fuels and value-added chemicals [19]. Additionally, ethanol exhibits a lower C–C bond activation energy compared to methane, enabling operation at milder temperatures (300–600 °C). Despite these advantages, both DRM and EDR processes face catalyst-related challenges. In the case of DRM, catalyst stability against coking and thermal degradation remains a critical concern, especially under prolonged operation [20,21,22]. For EDR, the focus lies on the development of highly active and stable catalysts capable of withstanding side reactions such as ethanol dehydration, cracking, and polycondensation [22,23].
This review aims to systematize recent strategies in the design of catalytic systems for DRM and EDR processes. Special emphasis is placed on Ni-based and bimetallic catalysts, as well as promising support materials such as perovskites, zeolites, and metal–organic frameworks (MOFs). The article provides a comparative analysis of approaches to enhancing catalytic activity and stability, strategies for mitigating carbon deposition, the role of promoters, and key thermodynamic and kinetic aspects of the reactions. A unified examination of DRM and EDR allows for the evaluation of their respective advantages and limitations and offers guidance for future research and industrial deployment within the context of sustainable energy and chemical production.

2. Dry Reforming of Methane

DRM is considered a promising technology for simultaneously converting two major greenhouse gases, CH4 and CO2, into synthesis gas (a mixture of H2 and CO) with a near-stoichiometric H2/CO ratio of 1:1. Compared to conventional reforming methods such as SMR and POM, DRM offers distinctive environmental and thermodynamic advantages. While SMR requires high steam input and produces excessive hydrogen (H2/CO ≈ 3), and POM demands pure oxygen and suffers from heat management issues, DRM avoids water consumption, utilizes CO2 directly as an oxidant, and aligns well with Fischer-Tropsch synthesis requirements [2]. The term “dry” in DRM reflects the replacement of water with CO2 as the oxidizing agent. With renewable energy sources, DRM can become a carbon-neutral process, contributing to the circular carbon economy [24,25].
Thermodynamically, DRM is a strongly endothermic reaction (CO2 + CH4 → 2CO + 2H2, ΔH° = +247 kJ/mol), requiring temperatures above 700 °C to achieve significant conversions [26]. However, these high temperatures also promote side reactions such as methane decomposition, Boudouard reaction, and carbon monoxide disproportionation, leading to coke formation and catalyst deactivation [27]. Table 1 summarizes the main and side reactions in DRM and their enthalpy changes [28,29].
Although DRM suffers from coke formation due to side reactions such as methane decomposition and the Boudouard reaction, several carbon-removal mechanisms can mitigate this challenge. Notably, deposited carbon can be partially gasified through reactions such as C + H2O → CO + H2 and oxidized via C + O2 → CO2 or CO, which help regenerate active sites and prolong catalyst life [17]. Despite these challenges, DRM continues to attract attention due to its dual benefit of greenhouse gas mitigation and syngas production. As shown in Figure 2, the number of publications related to DRM has been steadily increasing over the past decade, reflecting the growing scientific and industrial interest.

2.1. Nickel-Based Catalysts: Functionalization of Catalyst Supports

Despite significant progress in catalyst design, industrial implementation of nickel-based systems remains hindered by their rapid deactivation, primarily due to coke formation and sintering of the active metal phase. Recent studies (2023–2025) have demonstrated that deliberate functionalization of catalyst supports is a key strategy to mitigate these challenges, providing improved activity and long-term stability under harsh reaction conditions [24,30].
Zeolites can effectively stabilize nickel nanoparticles. SSZ-13 (CHA) features 8-membered-ring windows (~0.38 nm) and internal cages (~0.67–1.0 nm). However, 2–3 nm Ni nanoparticles reside predominantly on the external surface of the crystals and in mesopores formed during synthesis/post-treatment, as the micropores are too small to host them. Lanthanum doping of SSZ-13 introduces additional Brønsted acid sites that act as anchoring points for Ni precursors and strengthen metal–support interactions, promoting highly dispersed surface Ni. The zeolitic framework ensures efficient transport of reactants and products to active sites, while a hierarchical pore network provides additional stabilization. This synergy combining high Ni dispersion with improved site accessibility reduces carbon deposition by up to ~60% compared with conventional oxide supports [31,32].
Recent studies confirm that Ni-based zeolite catalysts remain a viable route for DRM: Ce-promoted Ni/SSZ-13 (CHA) exhibits enhanced coking resistance and stability, while interzeolite transformation enables CHA-type nickel silicate zeolites (Ni-CHA) with high activity under DRM conditions [33,34].
Among oxide-based supports, particular attention has been devoted to ZrO2 systems promoted with cerium oxide [35]. The introduction of 5 wt.% CeO2 into Ni/ZrO2 significantly enhances resistance to coke formation. This improvement is primarily associated with a twofold increase in oxygen vacancy concentration, as confirmed by Raman spectroscopy. Oxygen vacancies are critical for CO2 activation and for oxidizing carbonaceous intermediates. The promotional effect of CeO2 stems from its ability to undergo redox cycling between Ce3+ and Ce4+, facilitating continuous regeneration of the catalyst surface. These vacancies enable CO2 dissociation to form reactive O* species, which oxidize CHx fragments into CO, thereby preventing the accumulation of stable graphitic coke. As a result of this synergy, the Ni/CeO2-ZrO2 catalyst maintains high activity (CH4 conversion >85%, CO2 conversion >90%) over 500 h of continuous operation at 800 °C under a gas hourly space velocity (GHSV) of approximately 39,000 mL·gcat−1·h−1 without significant deactivation.
The authors [36] propose enhancing the stability of Ni-based catalysts through the functionalization of supports with nanomaterials possessing controlled oxygen mobility. The incorporation of cobalt into Ni/ZrO2 leads to the formation of Ni-Co-O clusters, which intensify redox cycling by enabling rapid oxygen transfer. In situ XRD and EXAFS analyses confirmed dynamic oxygen redistribution between the support and active phase, effectively suppressing the sintering of Ni particles (with sizes maintained below 5 nm after 100 h). It was shown that Co-activated oxygen vacancies in ZrO2 accelerate carbon gasification via CO2 according to the reaction: C + O* → CO. The high mobility of O2− ions across the Ni/ZrO2 interface prevents the accumulation of stable graphitic carbon. As a result, the coke content was reduced by 55% compared to unmodified Ni/ZrO2, while CH4 and CO2 conversions remained above 85% over 300 h of continuous operation at 800 °C under a WHSV of 18,000 mL·g−1·h−1.
An alternative approach involves the use of complex oxide systems, such as CaFe2O4 [37]. Investigations of this material as a support for nickel catalysts (15 wt.% Ni) have demonstrated its high efficiency in the DRM process. Under optimized reaction conditions (832.5 °C, CO2/CH4 ratio = 0.96, and gas hourly space velocity of 35,000 h−1), conversions of CH4 and CO2 reached 85% and 88%, respectively, with corresponding H2 and CO yields of 77.8% and 75.7%. Notably, despite the formation of carbon deposits confirmed by TGA and TEM analyses, the 15% Ni/CaFe2O4 catalyst exhibited no signs of deactivation after 20 h of continuous operation. This behavior is attributed to a dynamic equilibrium between carbon deposition and its gasification by surface oxygen species provided by the Ni/CaFe2O4 support. Such a “self-cleaning” mechanism represents a key advantage of this catalytic system.
Natural materials have attracted increasing attention as cost-effective and environmentally benign supports for nickel-based catalysts [38]. Diatomites subjected to plasma etching exhibit a unique hierarchical porous architecture that promotes effective dispersion of the nickel phase and enhances mass transport. Treatment with low-temperature Ar/H2 plasma generates mesopores ranging from 5 to 15 nm and increases the specific surface area up to 320 m2/g. Studies have demonstrated that a NiO-Co3O4/diatomite catalyst with 5–15 nm mesopores achieves a CH4 conversion of 89% at 850 °C at a volume rate of 1000 h−1, outperforming systems based on synthetic γ-Al2O3 supports. The key factor underlying the enhanced activity is the synergy between iron-containing impurities in the diatomite and cobalt, which leads to the formation of catalytically active Fe-Ni-Co alloys. These alloys exhibit improved resistance to coking due to their modified surface electronic properties.
Another natural material with promising characteristics is acid-treated tripolite [39]. Treatment with sulfuric acid (1 M H2SO4) selectively removes aluminosilicate impurities, generating a mesoporous structure with pronounced Brønsted acidity. These acid sites enhance the interaction between nickel particles and the support (SMSI—Strong Metal-Support Interaction), thereby reducing the catalyst deactivation rate by approximately 20%. The modified tripolite not only provides structural stabilization of Ni particles but also participates in CO2 activation, promoting in situ oxidation of carbonaceous species. This effect is particularly important for ensuring the long-term operational stability of the catalyst under dry reforming conditions.
Overall, the functionalization of catalyst supports offers broad opportunities for developing highly stable Ni-based catalysts for dry reforming of methane. Current strategies encompass the use of microporous zeolites with tunable acidity, oxide systems promoted with rare earth elements, and modified natural materials with hierarchical porosity. The key mechanisms responsible for improved stability include spatial confinement of nickel particles, enhanced metal-support interaction, increased oxygen mobility, and in situ oxidation of carbon deposits. Future research should focus on elucidating the fundamental interactions between nickel particles and functionalized supports at the atomic level, as well as on developing scalable synthesis methods suitable for industrial implementation. Particular attention should be given to designing composite supports that simultaneously provide high nickel dispersion, efficient oxygen transport, and optimized acid–base surface properties.

2.2. Bimetallic Ni-Based Catalysts

In pursuit of more robust DRM catalysts, extensive research has focused on bimetallic Ni-based catalysts that combine Ni with a second metal. The addition of an appropriate secondary metal can significantly enhance carbon tolerance and catalytic stability, thanks to synergistic effects that modify Ni’s catalytic properties. A variety of bimetallic combinations have been explored, including Ni alloyed with other transition metals (e.g., Co, Fe, Mo, W) as well as with noble metals (e.g., Pt, Ag) or rare-earth promoters (e.g., La, Ce), each offering distinct mechanisms to mitigate deactivation. These bimetallic systems aim to suppress coke formation and prevent Ni sintering simultaneously, thereby prolonging catalyst lifetime under demanding DRM conditions.
Ni–Co catalysts exemplify the benefits of alloying Ni with another transition metal. In bimetallic Ni–Co/γ-Al2O3 systems, the formation of a surface Ni–Co alloy plays a crucial role in suppressing carbon deposition and increasing the concentration of active metallic sites, enabling methane conversions approaching thermodynamic equilibrium at 600 °C [40,41,42,43].
Recent studies confirmed that layered-double-hydroxide-derived Ni–Co with an optimized Ni:Co ≈ 5:1 maintained CH4 and CO2 conversions above 85% for 200 h, the stability being linked to Ni–Co alloying and the generation of surface bidentate carbonates that accelerate the turnover of carbon intermediates [44].
Incorporating iron into Ni catalysts can engender unique active phases that actively consume deposited carbon. Under DRM conditions, Ni–Fe catalysts may form intermetallic carbides such as Fe5C2 that promote CO2-assisted gasification of carbon. This approach has enabled exceptionally stable performance; for instance, a Ni–Fe composite maintained ~95% CH4 and CO2 conversion over 350 h with negligible coke accumulation. Ab initio calculations confirm that the Ni–Fe5C2 phase lowers the activation barrier and continuously removes carbon species, thereby preventing deactivation [45].
The integration of molybdenum oxide into nickel-based catalysts [46,47] leads to the formation of several critical phases that significantly influence the catalytic behavior. Molybdenum carbides (Mo2C, MoCx) are generated via carburization during catalyst reduction in CH4/H2 [48]. The redox cycling between MoO3 and MoCx enables efficient carbon removal, reducing graphitic carbon formation by ~30% [49]. The formation of Ni-Mo alloys is commonly observed in Ni-Mo/YSZ systems [50], where Mo stabilizes metallic Ni and improves dispersion. For example, in Ni–Mo/Al2O3, the average Ni particle size decreased from 25 nm to 12 nm (Figure 3). Expanding on these findings, a hierarchical Ni–Mo/MgO catalyst sustained >80% conversion for 100 h through in situ Ni–Mo alloying and oxygen vacancies supplied by MgO, which together inhibited sintering and carbon accumulation [46].
Similar benefits arise from other refractory elements. Ni–W catalysts prepared by solid-state grinding exhibited CH4 and CO2 conversions of ~58% and 66% over 24 h, with ultrasmall Ni particles, strong Ni–support bonding, and WC phases suppressing coke accumulation [51]. These carbide-related mechanisms align with strategies to develop self-regenerating systems, such as Ni–Mo2C/CeO2, which exploit cyclic carburization–oxidation to maintain activity above 95% after 200 h on stream. In situ XRD confirmed that MoCx phases catalyze CO2-assisted gasification of carbon, providing autonomous coke removal.
Beyond transition metals, noble-metal additions have proven effective in stabilizing Ni. Even trace amounts of platinum can profoundly influence Ni’s behavior. Pt-decorated Ni/Al2O3 prepared by atomic-layer deposition preserved full activity at 650 °C for 48 h, whereas the monometallic analog deactivated rapidly. Mechanistically, Pt facilitated CO2 activation and diverted carbon into defective filamentous deposits that did not encapsulate Ni [52]. In fact, even inexpensive noble metals like silver can significantly improve Ni catalyst performance. Ni–Ag/MgAlO with Ag:Ni ≈ 1:19 retained ~95% of its initial activity after 26 h at 800 °C by suppressing filamentous coke and mitigating Ni aggregation [53].
Rare-earth-containing bimetallic designs exploit oxygen storage and vacancy chemistry to couple CH4 activation at Ni with rapid CO2-derived oxidation of surface carbon. In the study [54], the acyl-ammonia evaporation method was employed to synthesize a series of Ni–M–SiO2@SiO2 (M = Ce, Zr, La, Co, Mg), revealing that rare-earth oxides outperform transition metals in enhancing catalytic activity. CeO2-promoted catalysts achieved the highest CH4 conversion (57.2% at 600 °C and GHSV 24,000 mL (g_cat·h)−1) at 5 wt.% promoter loading, compared to 47.5% for the unpromoted analog. The 10Ni–1Ce–SiO2@SiO2 catalyst with a core–shell architecture outperformed its impregnated counterpart by delivering 18% higher CH4 conversion and 40% lower carbon deposition after 12 h, attributed to steric barriers restricting Ni growth (<5 nm) and the formation of nickel phyllosilicate structures that stabilize nanoparticles. Increasing calcination from 700 to 900 °C further enhanced strong metal–support interaction (SMSI), resulting in highly dispersed Ni particles (2–3 nm). The fraction of weakly bound NiO, which can easily reduce to larger Ni0 particles, was minimized. This stabilization suppresses the formation of agglomerated metallic Ni and thereby lowers the risk of sintering at elevated temperatures. Oxygen vacancies in CeO2 promoted CO2 activation and CHx oxidation, while hierarchical porosity improved mass transfer. Along similar lines, La-modified Ni catalysts generate La2O2CO3 species and abundant oxygen vacancies, which accelerate CO2 uptake and oxidation of surface carbon [55]. Ce-doped Ni/SSZ-13 further boosted activity by raising the Ce3+/Ce4+ ratio, stabilizing Ni dispersion, and suppressing coking under DRM [56].
Finally, unconventional alloying such as Ni–Zn offers a distinct anti-coking route. A Ni–Zn/SiO2 catalyst obtained via selective Zn sublimation achieved 88% CH4 and 95% CO2 conversion at 750 °C over 50 h, while reducing carbon deposition by an order of magnitude compared to Ni/SiO2. This was attributed to Ni–Zn interfacial stabilization and the in situ formation of hydroxyl species that continuously gasified surface carbon [57].
CO2-assisted regeneration of Ni–Zr catalysts. After four DRM cycles, a Ni–Zr catalyst accumulated filamentous (1D) and layered (2D) graphitic carbon that blocked Ni active sites, accompanied by sintering-induced particle growth from 5–10 nm to >50 nm. Regeneration by CO2 treatment (C + CO2 → 2 CO) reactivated the surface by gasifying carbon deposits, raising CH4 conversion from ~40% to ~70% and approaching the initial ~87%. The surface carbon content decreased from 55.4 to 26.4 at.% (XPS), while SEM showed shrinkage of Ni aggregates and effective coke removal. The regenerated catalyst exhibited improved stability in subsequent DRM cycles; notably, residual filamentous carbon did not block the active sites, as evidenced by sustained CH4 conversion (~45%) even in a partially deactivated state. This CO2-regeneration approach combines environmental benefit (CO2 utilization) with operational practicality, offering a viable path toward self-healing, continuous DRM operation [58].
Mechanistic note. Quantum-chemical DFT+U modeling shows that introducing ~5% Y into Ni decreases the barrier for CH4 dissociation from ≈1.8 to ≈0.9 eV through d-orbital polarization, thereby facilitating C–H bond cleavage. Experiments corroborate a multi-fold increase in CH4 conversion and reduced progression of CHx intermediates to graphitic carbon, while EXAFS detects shortened Ni–Y distances (~2.35 Å) and a positive shift in Ni partial charge, both consistent with electronically tuned, more reactive Ni sites [31].
Taken together, these findings delineate a coherent design space for bimetallic Ni catalysts in DRM: (i) alloying with Co or Zn to reshape Ni carbon chemistry and stabilize dispersion; (ii) pairing with carbide-forming Fe, Mo, or W to couple CH4 activation with continuous CO2-assisted carbon removal; and (iii) integrating oxygen-storing rare-earth elements (La, Ce) or trace noble metals (Pt, Ag) to unify high activity with coking resistance. By aligning metal–metal synergy with vacancy engineering and support effects, bimetallic Ni systems achieve long-term stability and activity approaching industrial requirements.

2.3. High-Entropy Alloy-Based Catalysts

A particularly promising direction is the application of high-entropy alloys. Catalysts of the Ni-Co-Fe-Cu-Mn/Al2O3 type have demonstrated superior coking resistance due to the suppression of carbon diffusion through the crystalline lattice. Testing at 800 °C confirmed long-term stability over 1000 h with less than 3% deactivation. A comparative overview of bimetallic nickel-based catalysts employed in dry reforming of methane is presented in Table 2.
A critical factor determining the coking resistance of Ni-based catalysts in DRM is the nickel particle size. Recent studies have shown that reducing the Ni particle size to the sub-10 nm range significantly decreases the catalyst’s tendency toward graphitization and carbon accumulation. Finely dispersed Ni particles promote methane activation primarily via full dissociation, forming reactive CHx intermediates that are less prone to polymerization and subsequent growth of graphitic layers. Moreover, the formation of active sites for the nucleation of filamentous and layered graphitic carbon is significantly hindered on small Ni particles.
In particular, Akri et al. [59] demonstrated that atomically dispersed Ni sites almost completely suppress carbon formation, even under prolonged reaction conditions. The authors of [60] showed that decreasing the Ni particle size on La-doped CeO2 significantly enhances the catalyst’s selectivity and coking resistance.
Han et al. [61] further emphasized the importance of Ni species strongly interacting with the support, showing that so-called “bounded Ni” sites on γ-Al2O3, associated with smaller particle sizes, provide enhanced resistance to coking and improved catalytic performance compared to more weakly bound, free-state Ni.
Chen, Phan, and co-workers [62,63] systematically investigated the effect of various preparation methods and established a clear correlation between smaller Ni particles and lower coke deposition. A recent study [64] further confirmed that an optimal fraction of ultra-fine Ni nanoparticles ensures maximum catalytic activity with minimal coking by increasing the energy barrier for graphitization. Achieving high Ni dispersion should be considered a strategically important goal in the development of efficient DRM catalysts, which must be addressed through careful selection of support materials, metal incorporation techniques, and post-synthetic treatments. Thus, bimetallic and promoter-modified Ni-based catalysts have demonstrated substantial progress in addressing the critical challenges of dry reforming of methane namely, coking and sintering. Ni-Co/γ-Al2O3 systems achieve methane conversions approaching thermodynamic equilibrium at 600 °C due to the formation of surface Ni-Co alloys that effectively suppress carbon deposition. The integration of MoO3 results in the generation of carbide phases (Mo2C) and Ni-Mo alloys, which not only reduce Ni particle size to approximately 12 nm but also facilitate a dynamic redox cycle for carbon gasification. The incorporation of MgO enhances the basicity of the support, thereby increasing CO2 adsorption and reducing coke formation by up to 70%.
Rare-earth elements such as Ce and Pr profoundly alter the electronic environment of Ni. CeO2 introduces oxygen vacancies that enable the oxidation of CHx intermediates, while Y doping lowers the activation energy barrier for CH4 dissociation from 1.8 eV to 0.9 eV. Innovative strategies such as BN coatings and high-entropy alloys confer long-term catalytic stability by suppressing carbon crystallization and limiting its diffusion through the active phase. A particularly promising direction involves the combination of multiple modifiers (e.g., Ni-Co-Ce/MgO-Al2O3), which may yield enhanced synergistic effects. Key objectives for future research include the optimization of synthesis techniques such as acyl ammonia evaporation for core–shell architectures and the in-depth investigation of in situ active site dynamics using EXAFS and DFT modeling approaches.
Table 2. Comparative table of bimetallic nickel catalysts for dry reforming of methane.
Table 2. Comparative table of bimetallic nickel catalysts for dry reforming of methane.
CatalystMethod of SynthesisReaction Conditions (T, P)Conversion (CH4/CO2)Rate of Coke
Formation
Stability (Running Time)Key FeaturesRef
20 wt.% Ni-Ru/CeO2 (NR)Wet impregnation450 °C, 1 atm92/70Low (TGA: 1.25 mg/g·h)1 hNi dispersion (2.19 nm), oxygen vacancies CeO2, Ru weakens the Ni-C bond[65]
7 wt.%Ni-3 wt.%Co/SiO2Sol–gel750 °C, 1 atm87/85Moderate (TGA: 3.5 mg/g·h)50 hNi-Co synergy, mesoporous SiO2 structure, sintering suppression[66]
15 wt.% Ni-5 wt.%Fe-Al (SCS)Solution combustion900 °C, 1 atm93/94High (TGA: >10%)20 hNi3Fe alloy formation, NiAl2O4 spinel, deactivation resistance[67]
2.5 wt.% Ni-2.5 wt.% Co/MgO-Al2O3 Impregnation700 °C, 1 atm73/76-30 hNi–Co synergy, basic MgO–Al2O3 (CO2 activation), NiO–MgO solid solution, MgAl2O4 matrix (anti-sintering)[68]
15 wt.% Ni-0.5 wt.%Re/MgAl2O4Wet impregnation750 °C, 1 atm79/67Moderate (TGA: 4.8 mg/g·h)50 hRe increases Ni dispersion, reduces carrier acidity[69]
3.75 wt.% Ni-1.25 wt.%Co/ScCeZrSolvothermal700 °C, 1 atm46.8/60Low (TGA: 2.5 mg/g·h)5.5 hMixed oxides Sc-Ce-Zr, strong metal-support interaction, isolated O2−[70]
1.2 wt.% Ni-0.3 wt.%Co/SiO2Sol–gel550 °C (light), 1 atm75/80Minimum (TGA: 0.8 mg/g·h)30 hPhotoactivation by hot carriers, selective oxidation *CHO→ CO[71]
Notes: *CHO → CO indicates the selective oxidation of the formyl group (–CHO) into carbon monoxide (CO) during photoactivation by hot carriers.

2.4. Perovskite-Based Catalysts

Perovskite-based catalysts offer unique advantages in the DRM, effectively addressing key limitations of conventional systems, such as carbon deposition, thermal instability, and high energy demand. Their structural and functional versatility including tunable composition, oxygen mobility, and metal synergy—opens new pathways for sustainable conversion of greenhouse gases into synthesis gas [72,73,74].
Lanthanum-based perovskites, such as LaNiO3 and La2NiO4, exhibit exceptional resistance to carbon accumulation. Their intrinsic surface properties neutralize acidic sites responsible for methane cracking, while high lattice oxygen mobility facilitates the oxidation of carbon deposits [75,76,77]. For instance, LaNiO3 synthesized via nanocasting using an SBA-15 mesoporous template was shown to deliver about 90% CH4 and CO2 conversion at 800 °C in temperature-programmed tests, while at 700 °C it maintained stable conversion for 48 h without detectable coke, consistent with the decomposition of LaNiO3 into Ni-based phases under reaction conditions [77]. The introduction of cerium into LaNi0.75Ce0.05 Zr0.20O3/8MgO-SiO2 enhances metal-support interactions and further reduces carbon deposition [78,79,80].
Elemental substitution in A/B sites (e.g., La, Ca, Co, Ti) enables the optimization of physicochemical properties. Replacing La with Ca in La1−xCaxNiO3 increases basicity and CO2 adsorption capacity, while incorporating elements such as Co or Ti into La–Ni perovskites enhances oxygen mobility and stabilizes the lattice, thereby improving redox activity [81,82].
The co-presence of Ni and Cu at the B-site generates bimetallic active centers, wherein Ni promotes CH4 dissociation and Cu facilitates CO2 activation. This synergy results in CH4 conversion exceeding 83% and CO selectivity above 97% at 700 °C [83,84].
Ni-Fe diamond-shaped nanoparticles produced via atomic layer deposition exhibit enhanced activity attributed to a favorable spontaneous alloying energy of −0.43 eV. Perovskite-based catalysts also display superior thermal stability compared to conventional systems. Nanostructured perovskites such as LaNiO3 and SrZrRuO3 retain structural integrity at temperatures up to 900 °C. LaNiO3 synthesized via auto-combustion shows a specific surface area of 150 m2/g, which prevents sintering of Ni nanoparticles (15–20 nm) even during extended operation [85]. SrZrRuO3 prepared via solid-state synthesis exhibits a space-time yield of 500 L·h−1·g·cat−1 at 900 °C, owing to the formation of a thermally stable perovskite matrix [86]. A promising direction involves the integration of DRM with solar energy. LaCu0.5Ni0.5O3 remains catalytically active at 450–750 °C, which is 200–300 °C lower than conventional Ni/Al2O3 catalysts. This reduction in reaction temperature decreases energy consumption to 1.61 MJ/kg of syngas, compared to 11.5 MJ/kg, leading to an 86% reduction in carbon footprint.
The reversible formation of oxygen vacancies (V0) in perovskites contributes to the regeneration of active sites. For instance, LaCu0.5Ni0.5O3 was shown to achieve 98% CO2 conversion per cycle in a two-step process involving partial methane oxidation followed by CO2 reduction. An overview of DRM-active perovskite catalysts is shown in Table 3.
Metal synergy in bimetallic (Ni-Cu, Ni-Co) and ternary (Co-Ni-Fe) systems enhances the activation of CH4 and CO2, resulting in high conversion and selectivity. The expression of Ni nanoparticles in structured formulations such as PrBaMnCoNi-15-Fe effectively prevents sintering of active sites, maintaining catalytic activity even at temperatures as high as 900 °C. Modification of oxygen vacancies through the incorporation of Ce, Sr, or Ca increases the oxidative capacity of the catalysts and effectively suppresses carbon accumulation. Hybrid architectures with supports (e.g., SiO2, Al2O3) contribute to the stabilization of dispersed metal nanoparticles, preventing their aggregation. Additionally, the low-temperature activity of Cu-containing perovskites (450–750 °C) reduces process energy requirements, facilitating integration with renewable energy sources.
A key factor influencing the activity of perovskite-based catalysts is the controlled exsolution of metal nanoparticles from the perovskite lattice under reducing conditions [88,89,90]. During the exsolution process, B-site metals (such as Ni, Fe, or Co) diffuse toward the surface of the oxide matrix, forming uniformly distributed and strongly embedded nanoparticles. This approach offers excellent resistance to sintering, carbon deposition, and structural degradation, thereby significantly improving catalyst stability compared to conventional methods of active metal incorporation. For instance, it has been demonstrated that Ni nanoparticles exsolved from La0.9Ce0.1Fe0.95Ni0.05O3 significantly enhance methane conversion, reduce carbon accumulation, and maintain long-term stability for over 120 h of operation [91,92,93].
Studies [94,95,96] emphasize the role of oxygen vacancies and redox cycling in regenerating exsolved active sites, further confirming the potential of such catalysts in DRM. Recent research indicates that the rational design of perovskite materials capable of in situ exsolution enables the development of catalysts with a high density of active sites, strong anchoring of metal nanoparticles, and resistance to aggregation under high-temperature DRM conditions.
The synergistic effect of exsolved metal species and lattice oxygen facilitates the activation of CH4 and CO2, while simultaneously promoting continuous oxidative removal of surface carbon deposits. Thus, the exsolution of metal nanoparticles from the perovskite lattice opens new avenues for the development of advanced catalysts for DRM. Recent studies have demonstrated that such materials effectively address the key challenges of the DRM process: they suppress carbon formation, exhibit strong resistance to thermal degradation, and improve the overall energy efficiency of the reaction. Due to their high catalytic activity, resistance to deactivation even after more than 100 reaction cycles, and ability to operate at intermediate temperatures, perovskite-based catalysts are emerging as a foundation for scalable and economically viable DRM technologies. Future research directions include the optimization of perovskite compositions, the development of cost-effective and environmentally friendly synthesis methods, and in-depth investigation of their long-term stability under industrial conditions. Thus, perovskite-based catalysts have the potential to play a pivotal role in the development of sustainable greenhouse gas conversion technologies, contributing to emissions reduction and the transition toward a closed carbon cycle in the energy and chemical sectors. These innovations directly address the main challenges associated with DRM: coke formation, thermal degradation, and high-energy consumption. The improved resistance of catalysts to deactivation (over 100 cycles) and their ability to operate in the intermediate temperature range offer a promising route toward scalable deployment of the technology. Future research should focus on compositional optimization, long-term stability evaluation, and the development of cost-effective synthesis methods suitable for industrial application. Consequently, perovskite-based catalysts are emerging as a key component in the development of sustainable systems for the conversion of greenhouse gases into synthesis gas.

2.5. MOF-Derived Catalysts

Metal–organic frameworks (MOFs) serve as versatile precursors for catalyst design, enabling the development of systems with controlled morphology, high dispersion of active metals, and strong metal-support interactions. Their unique architecture provides structural tunability and spatial confinement, which are critical for mitigating catalyst deactivation via coking and sintering [97,98,99].
The stability of MOF-derived Ni–MgO@mSiO2 is attributed to the mesoporous SiO2 framework and to interfacial M–O–Si linkages that anchor node-derived oxide/metal nanoparticles (Ni–O–Si, phyllosilicate-like) to the support [100]. Physicochemical analysis confirmed strong bonding between metallic centers and the SiO2 scaffold, effectively preventing Ni particle agglomeration (mean size: 7.2 nm) even at 700–800 °C. The coke formation rate (1.25 mg·g−1·h−1) is reduced sixfold compared to conventional analogs due to the presence of surface OH groups that react with CHx intermediates and the mesoporous structure that limits Ni migration. After 60 h of operation, the average Ni particle size increased only to 18.3 nm, while CH4 conversion remained at 75% at 700 °C.
Mechanistic aspects of MOF-derived catalyst formation at elevated temperatures. Under DRM-relevant temperatures, MOFs do not persist as crystalline frameworks: organic linkers undergo pyrolysis/decarboxylation (≈350–600 °C), while node species reorganize into oxide/metal nanoparticles at positions pre-defined by the original lattice. When the MOF is silicified, subsequent calcination promotes interfacial condensation and the formation of M–O–Si bridges that chemically tether these node-derived nanodomains within the mesoporous SiO2 network (Ni–O–Si, phyllosilicate-like), thereby limiting particle migration/coalescence and suppressing filamentous carbon. The atmosphere and thermal program govern crystallite size and defect chemistry (e.g., Ce3+/oxygen vacancies), which in turn affects coking and stability. This MOF-to-catalyst route differs from the pyrolysis of generic metal–organic precursors by offering a pre-organized distribution of metal centers (high nucleation density), inherited porosity/confinement, and upon silicification, chemical M–O–Si anchoring, yielding smaller, better-tethered nanoparticles and improved DRM stability [64,65,101].
In a related study [65], the Ni-Ce-BTC MOF was synthesized via solvothermal treatment, and after 120 h of operation, the average Ni particle size was still limited to 18.3 nm. Subsequent pyrolysis in different atmospheres (N2, H2, CO2) yielded Ni/CeO2 catalysts with controlled morphology. The N2-pyrolyzed sample (N2-pyr) possessed the smallest Ni (2.19 nm) and CeO2 (2.78 nm) crystallite sizes, high specific surface area (245 m2/g), and a mesoporous structure. CO2 pyrolysis resulted in larger Ni particles (13.6 nm) due to oxidation-induced agglomeration. N2-pyr exhibited minimal coke accumulation (2.88 mg·g−1·h−1), attributed to a high Ce3+ fraction (37%) that promoted oxygen vacancy formation for CHx oxidation, along with the presence of amorphous carbon forming a protective layer around Ni particles. In contrast, CO2-treated catalysts accumulated up to 65.4 mg·g−1·h−1 of coke due to weak metal-support interactions.
The Ni/CeO2-M catalyst, derived from a Ce-MOF precursor, demonstrated superior low-temperature performance (400–600 °C) [101]. The CeO2-M support, synthesized from Ce-MOF, features a nanorod morphology (50 nm diameter) and mesoporosity, leading to a high surface area (22.4 m2/g) and uniform Ni dispersion. The average Ni particle size in Ni/CeO2-M was 17.2 nm (versus 38.4 nm in Ni/CeO2-C), as confirmed by TEM and XRD. A higher Ce3+ content (28% vs. 24%) and elevated ID/IF2 g ratios in Raman spectra (0.18 vs. 0.16) indicated a greater concentration of oxygen vacancies, which facilitated CO2 activation. The coke formation rate on Ni/CeO2-M (1.25 mg·g−1·h−1) was significantly lower than that of Ni/CeO2-C (7.47 mg·g−1·h−1). Oxygen vacancies enhanced CO2 dissociation to CO and reactive oxygen species, which oxidized CHx intermediates and prevented their polymerization.
Strong Ni-CeO2 interaction and the mesoporous structure of CeO2-M inhibited Ni migration [102]. After 10 h of reaction, the Ni particle size increased only to 17.9 nm (vs. 39.5 nm for Ni/CeO2-C). At 550 °C, CH4 and CO2 conversions for Ni/CeO2-M reached 30.8% and 40.1%, respectively (compared to 15.5% and 27.3% for Ni/CeO2-C). After 10 h, Ni/CeO2-M showed a 39.6% activity loss, whereas Ni/CeO2-C experienced a 57.7% decline.
A solvothermal synthesis route (120 °C, 12 h) of MOG-Al-La gel followed by calcination at 750 °C yielded mesoporous LaAlO3 with high specific surface area (33 m2/g) and small crystallite size (28 nm) [79]. Ni nanoparticles (5.03 nm) were uniformly distributed on the perovskite surface, as confirmed by TEM and XRD. Strong Ni-support interactions suppressed agglomeration even at 700–800 °C. The coke formation rate was 1.25 mg·g−1·h−1, compared to 7.47 mg·g−1·h−1 for the sol–gel-derived counterpart. Oxygen vacancies in LaAlO3 promoted the oxidation of CHx intermediates, inhibiting their polymerization. The mesoporous architecture and MSI further limited Ni mobility. After 60 h of reaction, the average Ni particle size increased only to 18.3 nm. At 700 °C, CH4 and CO2 conversions reached 75% and 80%, respectively, after 20 h. The H2/CO selectivity ratio was 0.88, closely matching thermodynamic equilibrium. The catalyst outperformed conventional Ni/Al2O3 and Ni/LaAlO3 systems in both activity and stability [103,104].
The use of Zr-based metal–organic frameworks (UiO-66) enabled the synthesis of catalysts with controlled morphology [105]. Among the series, Ni/ZrO2-B (prepared by impregnating Zr-MOF with nickel) exhibited the smallest NiO particle size (20.47 nm) and the highest dispersion, as confirmed by XRD and TEM. In contrast, Ni/ZrO2-A (in situ Ni incorporation) and Ni/ZrO2-C (post-ZrO2 formation Ni addition) produced larger NiO particles (23.42 nm and 32.02 nm, respectively). All catalysts exhibited mesopores of ~25 nm, which improved mass transport, and BET surface areas up to 12.93 m2·g−1. Ni/ZrO2-B exhibited low coke accumulation (3.8 wt% after reaction) owing to abundant oxygen vacancies in the ZrO2 support that promote CO2 activation and subsequent oxidation of CHₓ intermediates, as well as strong metal–support interaction (MSI) that limits Ni agglomeration. For comparison, Ni/ZrO2-A and Ni/ZrO2-C accumulated 6.9 wt% and 0.7 wt% coke, respectively. After 6 h of reaction, the Ni particle size in Ni/ZrO2-B reached 16.60 nm, substantially smaller than in Ni/ZrO2-A (27.64 nm). Table 4 summarizes the key physicochemical characteristics of these MOF-derived catalysts for dry reforming of methane.
Thus, MOF-derived precursors enable the fabrication of DRM catalysts with high stability and activity. Recent studies on MOF-based catalysts for dry reforming of methane highlight several interconnected approaches aimed at overcoming key process limitations. Morphology control plays a central role: mesoporous structures based on SiO2, CeO2, or ZrO2 not only stabilize active metal nanoparticles by preventing agglomeration but also facilitate mass transport of reactants and products. For instance, the SiO2 mesopores in Ni–MgO@mSiO2 create diffusion channels that enhance reactant mobility while simultaneously restricting Ni migration. An important complement to morphological tuning is the generation of oxygen vacancies via the introduction of Ce3+/Ce4+ or Zr3+/Zr4+ species. These defects enhance the oxidative potential of the catalysts by activating CO2 for the oxidation of CHx intermediates, reducing the coke formation rate by ~60–80%. In hybrid systems such as LaAlO3/Ni, oxygen vacancies within the perovskite matrix further stabilize Ni nanoparticles, ensuring long-term catalyst stability even at 700–800 °C. Hybrid materials combining MOF-derived routes with perovskites (e.g., LaAlO3) or oxides (e.g., MgO) exhibit synergistic properties: perovskites provide thermal robustness, whereas the MOF-derived texture ensures high dispersion of active sites. For example, the Ni/CeO2-M composite derived from a Ce-MOF integrates the nanorod morphology of CeO2 with a mesoporous network, achieving 30.8% CH4 conversion at 550 °C. The low-temperature activity of such systems opens promising avenues for integrating DRM with renewable energy sources, reducing energy consumption by ~40–50% compared with conventional processes operating at 8,001,000 °C. These advances demonstrate that MOF-derived catalysts address coking and sintering while improving the economic and environmental viability of DRM.

3. Ethanol Dry Reforming

3.1. General Aspects of the EDR Process

Ethanol dry reforming has emerged as a promising pathway for the sustainable conversion of renewable ethanol and CO2 into syngas, with a theoretical H2/CO ratio of 1:1. Ethanol is an attractive feedstock due to its renewability, low toxicity, ease of handling, and potential for large-scale production from lignocellulosic biomass, municipal solid waste, and other sources [106,107]. Although steam reforming of ethanol has been extensively studied, recent years have witnessed increasing interest in EDR due to its advantages in CO2 utilization and compatibility with carbon-neutral strategies. Compared to methane, ethanol exhibits easier activation at lower temperatures due to the presence of -OH and -CH2- functional groups. Thermodynamically, the EDR reaction becomes favorable above 318 °C and is highly endothermic, requiring optimized conditions for effective syngas production [108,109]. Figure 4 illustrates the growing number of publications on EDR between 2015 and May 2025, reflecting the expanding research interest.
The primary EDR reaction is a highly endothermic process. The resulting syngas, with a theoretical H2/CO molar ratio of approximately 1:1, serves as a versatile precursor for methanol synthesis, higher alcohols, Fischer-Tropsch hydrocarbons, and hydrogen fuel cell applications [110,111,112]. Nevertheless, the EDR process is accompanied by side reactions that reduce selectivity and promote carbon deposition. Common competing reactions include ethanol dehydrogenation to acetaldehyde, subsequent decomposition to methane and CO, methanation, and coke formation (Table 5). Typical by-products include CH4, C2H4, H2O, and solid carbon. The overall process stability and yield of target products depend strongly on the catalyst nature and reaction parameters such as temperature, pressure, and reactant ratios.
To enhance the efficiency of EDR, both thermodynamic and kinetic studies have been conducted. For example, [113] used Aspen Plus modeling to demonstrate that the addition of O2 and H2O can regulate the H2/CO ratio and effectively suppress carbon formation. Feeding an O2/CO2/ethanol mixture at a molar ratio of 0.2/1/1 at 800 °C yields an H2/CO ratio of approximately 1, while the introduction of 2 moles of H2O at 650 °C increases this ratio to around 2. The integration of heat exchangers reduced overall energy consumption by 33–35%, highlighting the advantages of hybrid operating modes.
The kinetic aspects of EDR have also been explored by [114] using a LaCuO3 catalyst synthesized via the citrate sol–gel method. The authors found that the reaction follows a power-law kinetic model, with reaction orders of 0.64 for C2H5OH and 0.26 for CO2. The activation energy was calculated to be 44.2 kJ/mol. The catalyst exhibited high stability and minimal carbon deposition, as confirmed by SEM and FTIR analyses, making it a promising candidate for scale-up applications.
In summary, despite certain thermodynamic and technological limitations, EDR remains one of the most promising pathways for the sustainable transformation of ethanol and CO2. Modern strategies involving catalyst modification, optimization of reaction parameters, and the adoption of hybrid approaches (e.g., photothermal activation) pave the way toward stable and energy-efficient syngas production.

3.2. Catalytic Systems for the EDR Reaction

Noble metal-based catalysts (such as Pt, Rh, and Pd) are traditionally considered the most effective systems for reforming processes, including EDR [109,115,116]. These metals exhibit strong capabilities in activating C–C and C–H bonds in reactant molecules while also demonstrating high resistance to carbon deposition. However, their high cost and limited availability significantly hinder their widespread industrial application. This has prompted extensive efforts to identify more affordable and abundant alternatives.
Among the most promising substitutes are transition metals such as Ni, Co, and Cu, which also exhibit notable catalytic activity in reforming reactions [110,117,118]. Nickel, in particular, has gained widespread attention for EDR due to its strong ability to activate ethanol and CO2 molecules, combined with its relatively low cost and wide availability. Ni-based catalysts are known for their efficient bond cleavage and high yields of desired products [119].
Nevertheless, the application of nickel is accompanied by two key challenges: the tendency of Ni particles to sinter at high temperatures, and the formation of carbon deposits on the catalyst surface, which block active reaction sites [120]. Accelerated coke formation leads to rapid catalyst deactivation, posing one of the main limitations in EDR. Nickel is conventionally supported on γ-Al2O3 due to its high surface area and excellent thermal stability, which enable good metal dispersion and moderate metal-support interaction. However, under EDR conditions, the Lewis acid sites on γ-Al2O3 promote side ethanol dehydration reactions, forming ethylene, which significantly intensifies coking. The accumulation of dense carbon species on the nickel surface further accelerates catalyst deactivation. Cobalt has also been explored as an alternative active phase due to its effective C–C bond cleavage ability in organic molecules, although its application in EDR remains less investigated. Copper typically shows lower activity than Ni and Co in ethanol activation and is thus used primarily as a promoter or as part of complex mixed-oxide systems.
In summary, while nickel remains an active and cost-effective metal for EDR, its practical application requires overcoming issues of thermal stability and coke formation. This can be addressed through careful optimization of the catalyst support, control over structural features, and the incorporation of promoters to enhance the resistance of Ni-based systems to deactivation.

3.2.1. Influence of the Support on Catalyst Activity and Stability

The physicochemical properties and composition of the support play a crucial role in determining the activity, selectivity, and stability of EDR catalysts. The nature of the support governs the dispersion and state of the deposited nickel particles, as well as the prevalence of side reactions.
High surface area and well-developed porosity are also critical parameters. The use of mesostructured materials such as SBA-15 or KIT-6 allows for uniform distribution of the active phase and reduces the average Ni particle size, thereby minimizing sintering [119]. Furthermore, the catalyst synthesis method strongly influences the support’s defect structure. Techniques such as co-precipitation and sol–gel synthesis can produce nanostructured supports with high surface area and defect density, which improves metal dispersion and may indirectly enhance metal–support interactions rather than strengthening them solely by surface area increase. As a result, the concentration of easily reducible sites and oxygen vacancies responsible for coke oxidation can be significantly increased. In the case of EDR, excessively strong metal–support interaction is not required; a balanced interaction that stabilizes Ni particles while maintaining accessible active sites is more beneficial. For example, in [120], the effect of varying the Al2O3 content in the Al2O3-(Zr-Yb)O2 support was demonstrated. The Ni/65AZ catalyst exhibited higher ethanol and CO2 conversions and greater H2 and CO yields compared to Ni/35AZ under identical conditions, attributed to a higher concentration of oxygen vacancies and a more stable post-reaction texture.
It has been established that oxide supports with high oxygen mobility significantly suppress coke formation due to the involvement of oxygen vacancies in the oxidation of carbonaceous intermediates. Specifically, the incorporation of CeO2 or the use of mixed oxides such as CeO2-ZrO2 and Ce(La)Ox enhances the oxygen storage and release capacity, thereby mitigating carbon deposition on the catalyst surface [121]. The in situ activated oxygen oxidizes intermediate carbide species (*C → CO) and hydrogen species (*H → OH), thus preventing the growth of carbon deposits. In contrast, acidic supports such as γ-Al2O3 promote parallel reactions (e.g., ethanol dehydration), leading to increased coke formation. Consequently, replacing or modifying γ-Al2O3 with more inert or basic supports is essential for ensuring the long-term stability of Ni-based catalysts in EDR.
Shi et al. [122] compared synthesis routes for Ni-Ce-Zr catalysts and found that co-precipitation produced smaller Ni particles (<5 nm) and a higher proportion of Ni0 (58% vs. 27%, per XPS). This nanodispersed catalyst achieved 65.3% ethanol conversion at 700 °C and demonstrated improved resistance to carbon deposition. Similarly, [123] showed that a hard-templated CoZnO-HT catalyst achieved complete ethanol conversion at 550 °C and stable performance for 40 h, whereas the impregnated CoZnO-C counterpart accumulated acetone as a byproduct and exhibited lower activity. The enhanced performance of CoZnO-HT was linked to its smaller Co particle size and a higher oxygen vacancy concentration (39%). Raman spectroscopy confirmed that coke deposits on CoZnO-HT were mostly amorphous (D/G = 1.52), while CoZnO-C contained more graphitic carbon (D/G = 1.39), accelerating deactivation.
Another effective strategy involves encapsulating the active metal within an inert shell. For instance, the Co@SiO2 catalyst with a 20 nm silica coating exhibited complete ethanol conversion at 550 °C and no activity loss over 40 h [124]. The silica shell prevented direct contact between metal particles and coke precursors, while the high interfacial oxygen vacancy density (1.03) enhanced in situ carbon gasification. A promising alternative support is the Cu-Ce-Zr mixed oxide with a flower-like morphology (CuCeZr-M) [125]. This catalyst, synthesized via a one-step microwave-assisted method, delivered full ethanol conversion at 750 °C and stable operation for 50 h. Its high mobile oxygen content (28.7% Ce3+ per XPS) and low carbon accumulation (<3 wt%, TGA) account for its exceptional stability. The structural and coke-resistance properties of the Ni/xAl₂O₃–(Zr+Ce)O₂ catalyst after the ethanol dry reforming reaction are shown in Figure 5.
Support composition not only affects the rate but also the nature of coke formation. Fionov et al. [126] demonstrated that increasing Al2O3 content to 50 mol.% in Al2O3-(Zr+Ce)O2 led to smaller Ni particles (15 nm), stronger metal-support interactions, and nearly complete suppression of coking (mass loss <1% by TGA after 7 h of testing). At lower Al2O3 content (5–20%), ordered carbon structures such as nanospheres and thick nanotubes formed, while higher Al2O3 content (50–75%) resulted in amorphous carbon and hollow nanotubes. These latter forms did not block active Ni sites, explaining the minimal deactivation rate. Raman spectroscopy confirmed the shift toward more amorphous carbon (increasing ID/IG ratio) with increasing Al2O3, whereas samples with low Al2O3 content showed prominent D and G bands typical of graphitic carbon. TGA further confirmed the sharp reduction in coke formation: samples with 50–75% Al2O3 lost <1% of their mass during deposit burnout (500–700 °C), compared to 3.9–6.4% for supports with 5–20% Al2O3 (Figure 5a). These findings clearly demonstrate that optimizing support composition and structure-enhancing basicity, introducing oxygen vacancies, and increasing surface area are highly effective strategies for improving EDR catalyst stability and performance.

3.2.2. Modified and Multifunctional Catalytic Systems

In addition to support optimization, considerable research has focused on the development of modified and multifunctional catalytic systems that combine the advantages of multiple components to improve activity and resistance to coking. These include bimetallic catalysts, perovskite-type oxides, high-entropy materials, core–shell structures, and metal–organic framework (MOF)-derived catalysts. Key advancements in each of these approaches are discussed below.
Bimetallic and Promoted Catalysts. The incorporation of a second metal or promoter frequently leads to reduced coke formation without compromising catalytic activity. For instance, the addition of cobalt to a Ni-based catalyst supported on CeO2-ZrO2 significantly reduces carbon accumulation during DRM and EDR processes, while Ni remains the primary active component. Fedorova et al. [127] demonstrated that a Ni-Co/Ce0.75Zr0.25O2 bimetallic catalyst exhibited superior H2 and CO yields at 700 °C. The presence of Co decreased the rate of coking without negatively influencing the activity of the main reforming reactions. TPR analysis showed that Co weakens the strong interaction between Ni and the oxide support, inhibiting the formation of hard-to-reduce NiO phases and thus enhancing the long-term stability of the catalytic system.
Similarly, doping Ni catalysts with copper can improve metal dispersion and reduce the tendency for carbon deposition via the formation of Cu-Ni alloys. In [128], the addition of 1 wt.% Cu to a Ni/Al-Zr-Ce-O (50ACZ) system led to the formation of fine bimetallic particles (10–30 nm), improving metal–support interaction and inhibiting Ni agglomeration. The resulting 1Cu-9Ni/50ACZ catalyst achieved high hydrogen (78%) and CO (70%) yields at 750 °C (H2/CO ≈ 1.1). Copper also promoted the water-gas shift (WGS) reaction, increasing the relative H2 yield, although excessive Cu loading induced side reactions from ethanol decomposition, lowering the overall activity.
Beyond transition metal promoters, the introduction of alkali elements has proven effective in increasing surface basicity. For example, potassium-modified Ni/Al2O3 catalysts showed significantly improved coke resistance. The Ni/K2O-Al2O3 catalyst exhibited a much slower deactivation rate (activity slope −0.96 vs. −2.77 for unmodified Ni/Al2O3). TEM analysis revealed dense carbon nanotube bundles on Ni/Al2O3, whereas significantly fewer filamentous carbon structures were observed on Ni/K2O-Al2O3. TGA measurements indicated that the potassium-modified sample experienced <7% mass loss from carbon deposits (500–700 °C), compared to 20% for unmodified Ni/Al2O3. Raman spectra showed a reduced D/G intensity ratio (ID/IG ≈ 1.0 vs. 1.5), suggesting a lower fraction of amorphous, reactive carbon. CO2-TPD analysis confirmed an increase in weak basic sites upon K addition, correlating with the observed suppression of coke formation.
In summary, both bimetallic and alkali-promoted catalytic systems exhibit extended operational lifetimes by effectively limiting carbon deposition, thereby preserving Ni activity over prolonged reaction periods. The stability of Ni/Al₂O₃ and Ni/M₂O–Al₂O₃ (M = Li, Na, K) catalysts during the EDR process was evaluated using TEM, TGA, and Raman analyses (Figure 6).
Core–Shell structures and calcination optimization. Specifically engineered multiphase structures, such as core–shell architectures, provide a spatial separation between the reaction sites and regions prone to coke formation.
A recent study [130] demonstrated that, for a SiO2@Co@CeO2 core–shell catalyst, the calcination temperature plays a crucial role in determining catalytic performance. Calcination at 550 °C preserved the integrity of the shell structure and facilitated the formation of an active interfacial layer between Co and CeO2. As a result, the catalyst achieved complete ethanol conversion at just 500 °C and exhibited excellent stability. At this optimal treatment temperature, the amount of carbon deposited was less than 1 wt.% (according to TGA data), with the carbon deposits primarily in an amorphous form, as indicated by a Raman ID/IG ratio of approximately 0.77. Figure 7 presents TEM images of the spent SiO2@Co@CeO2 catalysts calcined at 550 °C, 700 °C, and 850 °C after undergoing the EDR reaction, illustrating the morphological changes and structural degradation with increasing calcination temperature. Increasing the calcination temperature to 700–850 °C led to a partial loss of the core–shell configuration, as evidenced by the agglomeration of Co3O4 and CeO2 nanoparticles and the formation of filamentous carbon deposits on the spent catalyst surface. The ID/IG ratio in Raman spectra decreased to the range of 0.4–0.6, indicating a higher proportion of graphitized carbon. TEM images of the spent catalysts further support these findings: in samples calcined at 550 °C and 700 °C (Figure 7a,b), the shell structure was preserved with minimal carbon deposition, while the sample calcined at 850 °C (Figure 7c) exhibited significant filamentous carbon formation along with aggregation of Co3O4 and CeO2 nanoparticles. These results highlight the critical importance of optimizing the calcination temperature to enhance the structural integrity and operational stability of core–shell catalysts under EDR conditions.
Such structural degradation is associated with accelerated catalyst deactivation. Therefore, precise control of thermal treatment parameters is essential for improving catalyst longevity by ensuring the formation of favorable structural and carbon deposition characteristics.

3.2.3. Advanced Catalytic Materials

Perovskite-based oxide catalysts. Complex oxides with perovskite-like structures have attracted considerable interest due to their compositional flexibility, high redox capacity, and robustness under harsh reaction environments [131]. In particular, copper-based perovskites such as LaCuO3 and CeCuO3 have demonstrated promising activity in EDR. LaCuO3 synthesized via the citrate gel method enabled complete CO2 conversion at approximately 706 °C, whereas CeCuO3 required a significantly higher temperature (840 °C) to achieve comparable performance.
As the reaction temperature increased from 725 °C to 800 °C, the H2/CO ratio for LaCuO3 decreased from 1.7 to 1.0 due to the suppression of side reactions, bringing the syngas composition closer to the stoichiometric ratio. The maximum CO yield reached 30% for LaCuO3 at the optimal feed ratio. Post-reaction XPS analysis revealed the formation of metallic Cu0 and an enrichment of surface-active oxygen species, with LaCuO3 exhibiting a higher surface oxygen content than CeCuO3. This finding is supported by TEM data: LaCuO3 retained its porous morphology with negligible carbon deposition, whereas CeCuO3 showed extensive formation of filamentous and amorphous carbon structures on its surface. Recent progress has also been made in the development of high-entropy perovskite oxides (Figure 8), which represent an advanced class of catalysts for EDR due to their enhanced thermal stability and oxygen mobility.
Eremeev et al. [132] synthesized a series of five-component perovskites, La(Cr0.2Mn0.2Fe0.2Co0.2Ni0.2)O3 and La(Ti0.2Mn0.2Fe0.2Co0.2Ni0.2)O3, which were subsequently impregnated with 5 wt.% Ni. Upon reduction, these high-entropy oxide-based catalysts exhibited outstanding catalytic activity in EDR: ethanol conversion reached 95–100%, while CO2 conversion was up to 65%, yielding a H2/CO ratio of 1.1–1.2 at 700 °C. The best performance was observed for the La(Cr0.2Mn0.2Fe0.2Co0.2Ni0.2)O3-based catalyst, which is attributed to its enhanced oxygen mobility and the presence of thermally stable defect states within the lattice. Importantly, after 10 h of continuous operation, no significant growth of Ni nanoparticles or formation of graphitized carbon was observed, as confirmed by TEM and Raman spectroscopy. Oxygen isotope exchange measurements revealed a high level of anionic conductivity (D × 10−12 cm2/s at 700 °C), which facilitates continuous removal of carbon species from the catalyst surface.
Therefore, perovskite and high-entropy oxide catalysts act as highly effective matrices for dispersing active metals and promoting the in situ oxidation of intermediate carbonaceous species, thereby substantially improving the long-term stability and performance of EDR systems.
MOF-derived and coordination polymer-based catalysts. Metal–organic frameworks (MOFs) have recently gained significant attention as precursors for the development of highly dispersed and porous catalytic systems. Pyrolysis of MOFs at moderate temperatures enables the formation of nanostructured metal oxide composites with uniformly distributed active sites. Tian et al. [133] synthesized a series of Co-MOx catalysts (M = Ce, Zr, Zn) via thermal decomposition of corresponding MOF precursors at 600 °C. The highest catalytic activity was observed for Co-CeO2, which achieved complete ethanol conversion at 550 °C. At 600 °C, this catalyst yielded approximately 52.3% H2 and 44.3% CO (molar H2/CO ratio = 1.18). Notably, the catalyst maintained stability over 40 h of operation. Its excellent performance is attributed to the formation of ultradispersed Co3O4 nanoparticles (5.8 nm) embedded within a CeO2 matrix. Strong Co3O4-CeO2 interaction was evidenced by a shift in the H2-TPR reduction temperature (271 °C), indicating facile oxygen transfer from the support to the active metal sites. In contrast, analogous MOF-derived Co-ZrO2 and Co-ZnO catalysts exhibited lower activity and higher coke formation, attributed to weaker interfacial interaction and lower oxygen mobility.
In another study [134], a bimetallic Co(Ni)/ZnO catalyst was synthesized from coordination polymers. The addition of Ni to the original Co/ZnO system significantly enhanced its performance: Co(Ni)/ZnO achieved complete ethanol conversion at 500 °C, whereas the monometallic Co/ZnO catalyst required 650 °C. Furthermore, the Co-Ni catalyst completely suppressed acetone formation as a side product (which reached up to 10 mol.% with Co/ZnO). Structural analyses (XRD, TEM, Raman, XPS) revealed that Ni incorporation reduced the average Co3O4 particle size (from ~20 to 17 nm) and increased the concentration of oxygen vacancies in the oxide matrix. This led to enhanced reducibility of oxide phases (as shown by H2-TPR) and consequently higher catalytic stability. Both examples highlight the strong potential of MOF-derived catalysts, which combine high activity and coke resistance due to their nanocomposite architecture. It is worth noting, however, that while these catalysts show delayed coke formation, further investigation is required to fully understand their deactivation mechanisms, as direct quantification of deposited carbon is often not reported.
Photothermal catalysis as a hybrid approach. One of the most recent advances in enhancing EDR performance involves the integration of thermal and photon-driven activation, known as photothermal catalysis. In this approach, catalytic materials are irradiated with concentrated light sources (e.g., infrared lasers or solar simulators), inducing localized heating and the excitation of plasmonic or polaronic states that facilitate the reaction. Photothermal activation allows high conversion rates at lower external temperatures by directly heating active sites and potentially leveraging non-thermal effects (e.g., generation of hot charge carriers) [135]. For instance, it has been shown [136] that infrared illumination of a Ni/ZrO2 catalyst increases ethanol conversion from 10% in the dark to 40% under irradiation, while simultaneously reducing the formation of byproducts such as acetone. This enhancement is attributed to photo-induced bond activation and accelerated oxidation of surface carbon intermediates, effectively raising the local temperature of catalytic sites without overheating the bulk reactor. Even more remarkable results have been achieved with specially engineered photoresponsive architectures. A dual-hollow microstructured NiZnOx-M catalyst, developed by Li et al. [137], demonstrated more than a twofold increase in ethanol conversion under light irradiation at 450 °C compared to thermal operation alone. Simultaneously, coke formation was significantly suppressed: TGA showed only 11% mass loss from deposits, and the coke accumulation rate was as low as 1.1 mg/g·h. This photothermally activated catalyst remained stable for over 100 h. Thus, harnessing photon energy enables more energy-efficient EDR by reducing thermal input and coke formation. Photothermal strategies are emerging as promising enhancements to traditional thermocatalysis, offering a pathway toward more economical and stable ethanol CO2 conversion processes. Integrated strategies in EDR catalyst design including bimetallic compositions, promoter additives, core–shell structures, perovskite-type materials, MOF-derived architectures, and photothermal systems demonstrate superior performance and durability compared to conventional Ni-based catalysts. These approaches provide improved metal dispersion, enhanced metal-support interactions, in situ generation of active oxygen species, and suppression of side reactions. Collectively, they significantly delay catalyst deactivation by mitigating coke accumulation and preserving active surface sites. Each strategy contributes to overcoming the inherent limitations of nickel systems and supports the advancement of sustainable and energy-efficient EDR technologies. Recent advances in EDR catalysts over the last five years are summarized in Table 6.

4. Comparative Analysis of DRM and EDR

Although both dry reforming of methane (DRM) and dry reforming of ethanol (EDR) share the overarching goal of transforming CO2 into valuable syngas, the two processes differ in fundamental ways that strongly influence their thermodynamics, reaction mechanisms, and catalyst design strategies. DRM is governed by the activation of methane and carbon dioxide, two of the most inert molecules in chemistry. As a consequence, high operating temperatures (typically above 700 °C) are required to overcome the formidable energy barriers associated with C–H and C=O bond cleavage. These harsh conditions inevitably exacerbate catalyst deactivation, particularly through carbon deposition via methane cracking and the Boudouard reaction, as well as through thermal sintering of metallic particles. In contrast, EDR benefits from the presence of ethanol, whose C–C and C–O bonds are significantly weaker, enabling the process to proceed at milder temperatures (300–600 °C). Yet this apparent advantage comes with its own challenges: ethanol undergoes a complex network of side reactions, including dehydration to ethylene and dehydrogenation to acetaldehyde, both of which are prone to further polymerization and condensation. Such pathways lead to the formation of amorphous or polymeric carbon, which can rapidly accumulate and poison active sites even under relatively moderate conditions.
Beyond thermodynamics, the feedstock basis of DRM and EDR also underscores their complementary roles in carbon management. DRM primarily relies on natural gas or biogas as a methane source, enabling the simultaneous utilization of CH4 and CO2 two of the most potent greenhouse gases. This dual mitigation strategy makes DRM attractive in regions where methane-rich resources are abundant. On the other hand, EDR employs bio-ethanol, a renewable and widely available liquid fuel produced from biomass. This direct coupling of renewable carbon with CO2 utilization situates EDR within the broader context of biorefineries and the circular carbon economy, highlighting its potential contribution to sustainable fuel and chemical production.
Catalyst development reveals important parallels between the two processes. Nickel remains the most extensively investigated active phase for both DRM and EDR, striking a practical balance between activity and cost. However, the intrinsic drawbacks of Ni, including its susceptibility to coking and sintering, necessitate careful catalyst engineering. Strategies such as the incorporation of CeO2, ZrO2, or other oxygen-storage materials to generate lattice oxygen, the promotion with rare-earth or alkaline elements to enhance CO2 activation, and the design of confinement architectures (core–shell structures, mesoporous frameworks) have been shown to suppress carbon accumulation and extend catalyst lifetime. In DRM, these measures primarily address graphitic or filamentous carbon formed at high temperatures, while in EDR they combat polymeric and oxygen-rich carbon species derived from ethanol intermediates. In both cases, oxygen-vacancy engineering and strong metal-support interactions emerge as unifying design principles that underpin catalyst stability. However, the optimum characteristics of the support are not identical. In DRM, carried out at ≥700 °C, thermally robust oxides with strong metal–support interaction (e.g., ZrO2, CeO2–ZrO2) are required to suppress Ni sintering and facilitate the removal of graphitic carbon. In EDR, which proceeds at 300–600 °C, excessively strong interaction is less advantageous; defect-rich but moderately interacting supports (e.g., CeO2, Ce–La oxides) are preferable, as they provide oxygen vacancies and limit ethanol dehydration while maintaining accessible Ni sites. This distinction highlights that MSI must be tuned to the reaction environment rather than applied as a universal rule. From a process-engineering standpoint, DRM remains thermally demanding, raising questions of heat management, reactor materials, and long-term catalyst integrity. EDR, while more moderate in terms of thermal input, requires stringent control over selectivity to avoid the build-up of undesired by-products. Both processes typically yield syngas with an H2/CO ratio close to unity well suited for Fischer-Tropsch synthesis and methanol production, yet in DRM this ratio often requires further adjustment, while in EDR it can be tuned more flexibly through co-feeding of oxygen or steam.
Modern synthesis routes for DRM and EDR catalysts. Recent progress in catalyst synthesis for dry reforming of methane (DRM) and ethanol dry reforming (EDR) centers on confinement architectures, defect/oxygen-mobility engineering, and thin-film or exsolution routes tailored to suppress coking and sintering while preserving metal dispersion.
For DRM, core–shell confinement has proved especially effective: Ni nanoparticles anchored as phyllosilicates and encapsulated by a porous SiO2 shell (∼20 nm) limit carbon growth and particle migration; a representative Ni/SiO2@SiO2 system prepared via ammonia-evaporation anchoring followed by sol–gel shelling remains stable at 600 °C [142]. In parallel, physical vapor deposition (PVD) enables ultrathin, adherent Ni layers on microchannel supports (Al2O3–ZrO2), combining short diffusion lengths with excellent coke resistance under DRM; magnetron-sputtered films demonstrate stable time-on-stream behavior [143]. MOF/mesotemplate routes continue to deliver high dispersion and fast mass transfer (e.g., CeO2-, ZrO2- or SiO2-based mesostructures), leveraging oxygen-vacancy formation to gasify CHx intermediates and suppress graphitic carbon. For EDR, high-entropy perovskite (HEO) precursors synthesized by modified Pechini/citrate methods and then Ni-impregnated yield defect-rich lattices with moderate high oxygen mobility and socketed Ni after reduction, offering high conversion with limited carbon deposition at 700 °C [132]. These HEOs align with the broader exsolution strategy whereby reducible perovskites nucleate strongly anchored nanoparticles under activation.
Nanocasting and MOF-derived approaches also feature prominently in EDR: Ni/KIT-6 and co-nanocast Ni/Ce0.8Zr0.2O2 frameworks enhance dispersion and stability while tuning acid–base sites to suppress ethanol dehydration [119,122]. Finally, infiltration of active phases into porous backbones (e.g., fuel-cell style electrodes) provides high triple-phase boundary density and robust metal anchoring for EDR operation [111].
In DRM, thin-film/confinement routes (core–shell or PVD) paired with vacancy-rich supports are now the most reliable path to coke tolerance above 700 °C. In EDR, HEO/perovskite-derived scaffolds with controlled exsolution, complemented by nanocasting to regulate acid–base chemistry, offer the best stability–selectivity balance at 300–600 °C. Prioritizing socketed metal states and oxygen-transport pathways should be the design default across both chemistries.
Taken together, DRM and EDR should not be viewed as competing but rather as complementary routes in the portfolio of CO2 utilization technologies. DRM is ideally positioned where methane resources and CO2 emissions intersect, while EDR aligns more naturally with renewable ethanol platforms and biorefinery schemes. Importantly, advances in catalyst design whether through defect engineering, perovskite or MOF-derived structures, or high-entropy alloys tend to be transferable across both systems, suggesting that progress in one domain accelerates development in the other. Ultimately, the integration of DRM and EDR within future energy and chemical production pathways will depend not only on catalyst innovation but also on their alignment with regional feedstock availability and decarbonization strategies. The advantages and disadvantages of DRM and EDR are compared in Table 7.

5. Conclusions and Future Perspectives

Catalytic dry reforming of methane and ethanol represents a promising technological pathway for simultaneously addressing environmental challenges and producing valuable chemical products. These processes play a critical role in mitigating climate change while fostering the transition toward more sustainable and environmentally responsible technologies in the energy and chemical sectors.
Rational catalyst design remains central to optimizing reforming reactions. Careful selection of active metal components, support materials, and promoters is essential for achieving high catalytic activity, long-term stability, and resistance to carbon deposition.
This review demonstrates that suppression of coking and improvement in catalyst stability can be achieved through a combination of specific strategies, including the creation of oxygen vacancies in supports (CeO2, ZrO2), doping with rare-earth (La, Ce, Pr) and transition metals, spatial confinement of active particles within mesoporous structures, and morphological control through core–shell architectures. Notably, similar stabilization strategies applied to Ni-, Cu-, and Co-based as well as perovskite-type systems enhance their resilience in both DRM and EDR, despite differences in feedstocks and coping mechanisms.
Future progress should focus on several key directions: optimization of catalyst synthesis for improved control of morphology, porosity, and metal dispersion; investigation of the in situ dynamics of active sites using advanced characterization methods (e.g., EXAFS, DFT modeling); development of hybrid catalysts that integrate the advantages of different systems; integration with renewable energy sources to reduce the carbon footprint; and scaling up laboratory technologies for industrial applications. The implementation of these approaches will enable the development of next-generation catalysts capable of meeting modern requirements for environmental compatibility, energy efficiency, and economic viability in the conversion of greenhouse gases into valuable chemical products.

Author Contributions

Conceptualization, M.M. and G.Y.; methodology, M.M.; K.D. and M.A.; validation, M.A., L.M. and M.M.; formal analysis, N.M. and L.M.; data curation, M.M. and G.Y.; writing—original draft preparation, M.M. and G.Y.; writing—review and editing, M.M. and G.Y.; supervision, project administration, and funding acquisition, M.A. All authors have read and agreed to the published version of the manuscript.

Funding

The work was carried out with financial support from the Ministry of Science and Higher Education of the Republic of Kazakhstan within the framework of scientific projects No. AP19678248, BR24992935.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gujar, J.P.; Verma, A.; Modhera, B. Optimizing Glycerol Conversion to Hydrogen: A Critical Review of Catalytic Reforming Processes and Catalyst Design Strategies. Int. J. Hydrogen Energy 2025, 109, 823–850. [Google Scholar] [CrossRef]
  2. Jang, W.-J.; Shim, J.-O.; Kim, H.-M.; Yoo, S.-Y.; Roh, H.-S. A Review on Dry Reforming of Methane in Aspect of Catalytic Properties. Catal. Today 2019, 324, 15–26. [Google Scholar] [CrossRef]
  3. Rajendran, S.; Al-Samydai, A.; Palani, G.; Trilaksana, H.; Sathish, T.; Giri, J.; Saravanan, R.; Lalvani, J.I.J.; Nasri, F. Replacement of Petroleum Based Products with Plant-Based Materials, Green and Sustainable Energy—A Review. Eng. Rep. 2025, 7, e70108. [Google Scholar] [CrossRef]
  4. Osazuwa, O.U.; Ng, K.H. An Overview on the Carbon Deposited during Dry Reforming of Methane (DRM): Its Formation, Deposition, Identification, and Quantification. Results Eng. 2025, 25, 104328. [Google Scholar] [CrossRef]
  5. Schleussner, C.-F.; Rogelj, J.; Schaeffer, M.; Lissner, T.; Licker, R.; Fischer, E.M.; Knutti, R.; Levermann, A.; Frieler, K.; Hare, W. Science and Policy Characteristics of the Paris Agreement Temperature Goal. Nat. Clim. Change 2016, 6, 827–835. [Google Scholar] [CrossRef]
  6. Torres Galvis, H.M.; De Jong, K.P. Catalysts for Production of Lower Olefins from Synthesis Gas: A Review. ACS Catal. 2013, 3, 2130–2149. [Google Scholar] [CrossRef]
  7. Vo, C.-M.; Cao, A.N.T.; Saleh Qazaq, A.; Pham, C.Q.; Nguyen, D.L.T.; Alsaiari, M.; Vu, T.V.; Sharma, A.; Phuong, P.T.T.; Tran Van, T.; et al. Toward Syngas Production from Simulated Biogas Dry Reforming: Promotional Effect of Calcium on Cobalt-Based Catalysts Performance. Fuel 2022, 326, 125106. [Google Scholar] [CrossRef]
  8. Wang, Y.; Li, R.; Zeng, C.; Sun, W.; Fan, H.; Ma, Q.; Zhao, T.-S. Recent Research Progress of Methane Dry Reforming to Syngas. Fuel 2025, 398, 135535. [Google Scholar] [CrossRef]
  9. Chavando, J.A.M.; Silva, V.; Caballero, C.G.T.; Cardoso, J.S.; Tarelho, L.A.C.; Eusébio, D. Future Prospects and Industrial Outlook of Syngas Applications. In Advances in Synthesis Gas: Methods, Technologies and Applications; Elsevier: Amsterdam, The Netherlands, 2023; pp. 427–463. ISBN 978-0-323-91878-7. [Google Scholar]
  10. Abdellatief, T.M.M.; Ershov, M.A.; Kapustin, V.M.; Chernysheva, E.A.; Savelenko, V.D.; Makhmudova, A.E.; Potanin, D.A.; Salameh, T.; Abdelkareem, M.A.; Olabi, A.G. Innovative Conceptional Approach to Quantify the Potential Benefits of Gasoline-Methanol Blends and Their Conceptualization on Fuzzy Modeling. Int. J. Hydrogen Energy 2022, 47, 35096–35111. [Google Scholar] [CrossRef]
  11. Okonye, L.U.; Yao, Y.; Ren, J.; Liu, X.; Hildebrandt, D. A Perspective on the Activation Energy Dependence of the Fischer-Tropsch Synthesis Reaction Mechanism. Chem. Eng. Sci. 2023, 265, 118259. [Google Scholar] [CrossRef]
  12. Jang, W.-J.; Jeong, D.-W.; Shim, J.-O.; Kim, H.-M.; Roh, H.-S.; Son, I.H.; Lee, S.J. Combined Steam and Carbon Dioxide Reforming of Methane and Side Reactions: Thermodynamic Equilibrium Analysis and Experimental Application. Appl. Energy 2016, 173, 80–91. [Google Scholar] [CrossRef]
  13. Djinović, P.; Osojnik Črnivec, I.G.; Erjavec, B.; Pintar, A. Influence of Active Metal Loading and Oxygen Mobility on Coke-Free Dry Reforming of Ni–Co Bimetallic Catalysts. Appl. Catal. B Environ. 2012, 125, 259–270. [Google Scholar] [CrossRef]
  14. Kumar Yadav, P.; Sharma, S. Ni, Co & Cu Substituted CeO2: A Catalyst That Improves H2/CO Ratio in the Dry Reforming of Methane. Fuel 2024, 358, 130163. [Google Scholar] [CrossRef]
  15. Agún, B.; Abánades, A. Comprehensive Review on Dry Reforming of Methane: Challenges and Potential for Greenhouse Gas Mitigation. Int. J. Hydrogen Energy 2025, 103, 395–414. [Google Scholar] [CrossRef]
  16. Rosha, P.; Singh, R.; Mohapatra, S.K.; Mahla, S.K.; Dhir, A. Optimization of Hydrogen-Enriched Biogas by Dry Oxidative Reforming with Nickel Nanopowder Using Response Surface Methodology. Energy Fuels 2018, 32, 6995–7001. [Google Scholar] [CrossRef]
  17. Zafarnak, S.; Rahimpour, M.R. Dry Reforming of Methane for CO2 Conversion Using Sintering-Resistant Halloysite-Supported Ni Catalysts Enhanced by Alkaline-Earth Metal Oxide Promoters. Fuel 2025, 395, 135168. [Google Scholar] [CrossRef]
  18. Benajes, J.; García, A.; Monsalve-Serrano, J.; Guzmán-Mendoza, M. A Review on Low Carbon Fuels for Road Vehicles: The Good, the Bad and the Energy Potential for the Transport Sector. Fuel 2024, 361, 130647. [Google Scholar] [CrossRef]
  19. Shafiqah, M.-N.N.; Siang, T.J.; Kumar, P.S.; Ahmad, Z.; Jalil, A.A.; Bahari, M.B.; Van Le, Q.; Xiao, L.; Mofijur, M.; Xia, C.; et al. Advanced Catalysts and Effect of Operating Parameters in Ethanol Dry Reforming for Hydrogen Generation. A Review. Environ. Chem. Lett. 2022, 20, 1695–1718. [Google Scholar] [CrossRef]
  20. Abbasi, S.; Abbasi, M.; Tabkhi, F.; Akhlaghi, B. Syngas Production plus Reducing Carbon Dioxide Emission Using Dry Reforming of Methane: Utilizing Low-Cost Ni-Based Catalysts. Oil Gas Sci. Technol. Rev. D’IFP Energ. Nouv. 2020, 75, 22. [Google Scholar] [CrossRef]
  21. Arora, S.; Prasad, R. An Overview on Dry Reforming of Methane: Strategies to Reduce Carbonaceous Deactivation of Catalysts. RSC Adv. 2016, 6, 108668–108688. [Google Scholar] [CrossRef]
  22. Hussien, A.G.S.; Polychronopoulou, K. A Review on the Different Aspects and Challenges of the Dry Reforming of Methane (DRM) Reaction. Nanomaterials 2022, 12, 3400. [Google Scholar] [CrossRef]
  23. Yousefi, S.; Tavakolian, M.; Rahimpour, M.R. Bio-Templated Ni/MgO-Al2O3 as an Efficient Catalyst toward Methane Dry Reforming. Inorg. Chem. Commun. 2025, 171, 113557. [Google Scholar] [CrossRef]
  24. Alhassan, M.; Jalil, A.A.; Nabgan, W.; Hamid, M.Y.S.; Bahari, M.B.; Ikram, M. Bibliometric Studies and Impediments to Valorization of Dry Reforming of Methane for Hydrogen Production. Fuel 2022, 328, 125240. [Google Scholar] [CrossRef]
  25. Wittich, K.; Krämer, M.; Bottke, N.; Schunk, S.A. Catalytic Dry Reforming of Methane: Insights from Model Systems. ChemCatChem 2020, 12, 2130–2147. [Google Scholar] [CrossRef]
  26. Yentekakis, I.V.; Panagiotopoulou, P.; Artemakis, G. A Review of Recent Efforts to Promote Dry Reforming of Methane (DRM) to Syngas Production via Bimetallic Catalyst Formulations. Appl. Catal. B Environ. 2021, 296, 120210. [Google Scholar] [CrossRef]
  27. Pham, T.T.P.; Ro, K.S.; Chen, L.; Mahajan, D.; Siang, T.J.; Ashik, U.P.M.; Hayashi, J.; Pham Minh, D.; Vo, D.-V.N. Microwave-Assisted Dry Reforming of Methane for Syngas Production: A Review. Environ. Chem. Lett. 2020, 18, 1987–2019. [Google Scholar] [CrossRef]
  28. Ordoñez Ramos, F.; Sapag, K.; Múnera, J.F.; Tarditi, A.M. One-Pot Ni-ZSM-5 Catalysts for Hydrogen Production from the Dry Reforming of Methane. Mol. Catal. 2025, 577, 114963. [Google Scholar] [CrossRef]
  29. Park, H.-R.; Kim, B.-J.; Ryu, S.-J.; Jeon, Y.; Lee, S.S.; Bae, J.-W.; Roh, H.-S. Super-Dry Reforming of Methane over Surface Oxygen Mobility Enhanced Ni/MgO-Ce/SBA-15 Catalysts. Catal. Today 2025, 453, 115257. [Google Scholar] [CrossRef]
  30. Aldosari, O.F.; Hussain, I.; Mohammed Aitani, A.; Alotaibi, S.; Abdul Jalil, A. The Critical Role of Intrinsic Catalytic Properties for Enhanced Dry Reforming of Methane (DRM): Recent Advances, Challenges and Techno-Feasibility Assessments. J. Ind. Eng. Chem. 2024, 133, 1–37. [Google Scholar] [CrossRef]
  31. Ahn, S.; Littlewood, P.; Liu, Y.; Marks, T.J.; Stair, P.C. Stabilizing Supported Ni Catalysts for Dry Reforming of Methane by Combined La Doping and Al Overcoating Using Atomic Layer Deposition. ACS Catal. 2022, 12, 10522–10530. [Google Scholar] [CrossRef]
  32. Zhu, H.; Chen, H.; Zhang, M.; Liang, C.; Duan, L. Recent Advances in Promoting Dry Reforming of Methane Using Nickel-Based Catalysts. Catal. Sci. Technol. 2024, 14, 1712–1729. [Google Scholar] [CrossRef]
  33. Lee, S.; Thuan Khiet Nguyen, T.; Kweon, S.; Park, M.B. Nickel Silicate CHA-Type Zeolite Prepared by Interzeolite Transformation and Its Catalytic Activity in Dry Reforming of Methane. Chem. Eng. J. 2024, 498, 155602. [Google Scholar] [CrossRef]
  34. Han, D.; Yang, Y.; Wu, T.; Wang, Z.; Xiong, J.; Zou, J.; Wei, Y. Synergistic Effect of Ni and CeOx in NiCeOx/SSZ-13 Catalyst for Boosting Activity and Stability during Dry Reforming of Methane. Catal. Today 2024, 435, 114720. [Google Scholar] [CrossRef]
  35. Al-Fatesh, A.S.; Kumar, R.; Fakeeha, A.H.; Kasim, S.O.; Khatri, J.; Ibrahim, A.A.; Arasheed, R.; Alabdulsalam, M.; Lanre, M.S.; Osman, A.I.; et al. Promotional Effect of Magnesium Oxide for a Stable Nickel-Based Catalyst in Dry Reforming of Methane. Sci. Rep. 2020, 10, 13861. [Google Scholar] [CrossRef]
  36. Liu, K.; Ye, L.; Cao, Z.; Li, M.; Zhang, L.; Liu, W.; Ma, Q.; Peng, H. Synergistic Enhancement of Coke Resistance in Methane Dry Reforming via Oxygen Vacancies and Spatial Confinement on Ni-ZrO2/DMS Catalysts. Fuel 2025, 388, 134588. [Google Scholar] [CrossRef]
  37. Hossain, M.A.; Ayodele, B.V.; Cheng, C.K.; Khan, M.R. Syngas Production from Catalytic CO2 Reforming of CH4 over CaFe2O4 Supported Ni and Co Catalysts: Full Factorial Design Screening. Bull. Chem. React. Eng. Catal. 2018, 13, 57–73. [Google Scholar] [CrossRef]
  38. Yergaziyeva, G.Y.; Kutelia, E.; Dossumov, K.; Gventsadze, D.; Jalabadze, N.; Dzigrashvili, T.; Mambetova, M.M.; Anissova, M.M.; Nadaraia, L.; Tsurtsumia, O.; et al. Comparative Study the Activity in Dry Reforming of Methane of Bioxide NiO-Co3O4 and NiO-Fe2O3 Systems Supported on the Granulated Natural Diatomite. Combust. Plasma Chem. 2023, 21, 89–97. [Google Scholar] [CrossRef]
  39. Azeem, S.; Ashraf, M.; Hussain, S. Role of Stable Ni Nano Catalysts for Dry Reforming of Methane. Adsorpt. Sci. Technol. 2025, 43, 02636174251330376. [Google Scholar] [CrossRef]
  40. Guo, S.; Sun, Y.; Zhang, Y.; Zhang, C.; Li, Y.; Bai, J. Bimetallic Nickel-Cobalt Catalysts and Their Application in Dry Reforming Reaction of Methane. Fuel 2024, 358, 130290. [Google Scholar] [CrossRef]
  41. Mousavinejad, S.A.; Farsi, M.; Rahimpour, M.R. Role of Promoter on the Catalytic Activity of Novel Hollow Bimetallic Ni–Co/Al2O3 Catalyst in the Dry Reforming of Methane Process: Optimization Using Response Surface Methodology. J. Energy Inst. 2024, 113, 101524. [Google Scholar] [CrossRef]
  42. Zafarnak, S.; Rahimpour, M.R. Co-Ni Bimetallic Supported on Mullite as a Promising Catalyst for Biogas Dry Reforming toward Hydrogen Production. Mol. Catal. 2023, 534, 112803. [Google Scholar] [CrossRef]
  43. Zou, Z.; Zhang, T.; Lv, L.; Tang, W.; Zhang, G.; Kumar Gupta, R.; Wang, Y.; Tang, S. Preparation Adjacent Ni-Co Bimetallic Nano Catalyst for Dry Reforming of Methane. Fuel 2023, 343, 128013. [Google Scholar] [CrossRef]
  44. Mohamad Ali, M.; Chaghouri, M.; Tidahy, H.L.; Marinova, M.; Guy, H.; Durécu, S.; Royer, S.; Abi-Aad, E.; Ciotonea, C.; Gennequin, C. Catalytic Performance of Shaped Ni Co Extrudates for Dry Reforming of Methane. Chem. Eng. J. 2025, 522, 167269. [Google Scholar] [CrossRef]
  45. Liu, J.; Li, Y.; Che, Y.; Shi, W.; Wang, D.; Cui, R.; Feng, X.; Zhang, J.; Liu, H.; Tian, Y. In Situ Synthesizable High-Performance Ni-Fe5C2 Composite Catalysts for Dry Reforming of Methane. Chem. Eng. J. 2025, 517, 164441. [Google Scholar] [CrossRef]
  46. Yan, X.; Liang, H.; Wang, R.; Li, D.; Li, J.; Peng, W.; Zhang, J. Preparation of High Stability Hierarchical Bimetallic Ni-Mo Catalysts for CO2 Reforming with Methane. Mol. Catal. 2025, 573, 114835. [Google Scholar] [CrossRef]
  47. Zhao, W.; Fu, Q.; Xie, B.; Ni, Z.; Xia, S. Mechanistic Study of Transition Metal Loaded/Doped Ni − MgO Catalyzed Dry Reforming of Methane: DFT Calculations. Chem. Phys. Lett. 2024, 853, 141538. [Google Scholar] [CrossRef]
  48. Saconsint, S.; Srifa, A.; Koo-Amornpattana, W.; Assabumrungrat, S.; Sano, N.; Fukuhara, C.; Ratchahat, S. Development of Ni–Mo Carbide Catalyst for Production of Syngas and CNTs by Dry Reforming of Biogas. Sci. Rep. 2023, 13, 12928. [Google Scholar] [CrossRef]
  49. Zhang, X.; Deng, J.; Pupucevski, M.; Impeng, S.; Yang, B.; Chen, G.; Kuboon, S.; Zhong, Q.; Faungnawakij, K.; Zheng, L.; et al. High-Performance Binary Mo–Ni Catalysts for Efficient Carbon Removal during Carbon Dioxide Reforming of Methane. ACS Catal. 2021, 11, 12087–12095. [Google Scholar] [CrossRef]
  50. Majewski, A.J.; Singh, S.K.; Labhasetwar, N.K.; Steinberger-Wilckens, R. Nickel–Molybdenum Catalysts for Combined Solid Oxide Fuel Cell Internal Steam and Dry Reforming. Chem. Eng. Sci. 2021, 232, 116341. [Google Scholar] [CrossRef]
  51. Duan, S.; Wang, X.; Ren, R.; Lv, Y. Effects of Preparation Methods on the Size and Interface of Ni–W Bimetallic Catalysts for Dry Reforming of Methane. Int. J. Hydrogen Energy 2024, 54, 1469–1477. [Google Scholar] [CrossRef]
  52. Zhao, R.; Cao, K.; Ye, R.; Tang, Y.; Du, C.; Liu, F.; Zhao, Y.; Chen, R.; Shan, B. Deciphering the Stability Mechanism of Pt-Ni/Al2O3 Catalysts in Syngas Production via DRM. Chem. Eng. J. 2024, 491, 151966. [Google Scholar] [CrossRef]
  53. Ji, R.; Su, W.; Men, X.; Yang, J.; Song, X.; Bai, Y.; Wang, J.; Lv, P.; Yu, G. Ultra-Low Amount of Ag Substantially Improves the Stability for Dry Reforming of Methane on Ni/Ag/MgAlO Bimetallic Catalyst. Appl. Surf. Sci. 2025, 688, 162320. [Google Scholar] [CrossRef]
  54. Kaviani, M.; Rezaei, M.; Alavi, S.M.; Akbari, E. Effect of Promoters and Calcination Temperature on the Performance of Nickel Silica Core-Shell Catalyst in Biogas Dry Reforming. J. Energy Inst. 2025, 120, 102062. [Google Scholar] [CrossRef]
  55. Li, B.; He, J.; Yuan, X. Effect of Lanthanum Modification on Core–Shell Structured Nitrogen-Doped Carbon Coated Nickel Catalysts for Dry Reforming of Methane. Fuel 2025, 388, 134495. [Google Scholar] [CrossRef]
  56. Ren, Q.; Wang, S.; Liu, H.; Zhang, X.; Cui, Y.; Zhang, G.-R.; Nie, L.; Mei, D. Synergistic Ni-Ce-SSZ-13 Catalysts for Enhanced Dry Reforming of Methane. Int. J. Hydrogen Energy 2025, 104, 202–211. [Google Scholar] [CrossRef]
  57. Jin, S.; Ye, D.; Zhang, T.; Lv, L.; Tang, W.; Wang, Y.; Zou, Z.; Tang, S. Selective Sublimation of Zn to Prepare NiZn/SiO2 Catalysts for Dry Reforming of Methane: Insight of Synergistic Catalysis and Anti-Coking. Chem. Eng. J. 2025, 509, 161198. [Google Scholar] [CrossRef]
  58. Hosseinpour, M.; Moser, T.; Klötzer, B.; Penner, S. Regeneration of Ni–Zr Methane Dry Reforming Catalysts in CO2: Reduction of Coking and Ni Redispersion. ACS Catal. 2025, 15, 3314–3327. [Google Scholar] [CrossRef]
  59. Akri, M.; Zhao, S.; Li, X.; Zang, K.; Lee, A.F.; Isaacs, M.A.; Xi, W.; Gangarajula, Y.; Luo, J.; Ren, Y.; et al. Atomically Dispersed Nickel as Coke-Resistant Active Sites for Methane Dry Reforming. Nat. Commun. 2019, 10, 5181. [Google Scholar] [CrossRef]
  60. Menegazzo, F.; Pizzolitto, C.; Ghedini, E.; Di Michele, A.; Cruciani, G.; Signoretto, M. Development of La Doped Ni/CeO2 for CH4/CO2 Reforming. C 2018, 4, 60. [Google Scholar] [CrossRef]
  61. Han, J.W.; Park, J.S.; Choi, M.S.; Lee, H. Uncoupling the Size and Support Effects of Ni Catalysts for Dry Reforming of Methane. Appl. Catal. B Environ. 2017, 203, 625–632. [Google Scholar] [CrossRef]
  62. Chen, C.; Wang, W.; Ren, Q.; Ye, R.; Nie, N.; Liu, Z.; Zhang, L.; Xiao, J. Impact of Preparation Method on Nickel Speciation and Methane Dry Reforming Performance of Ni/SiO2 Catalysts. Front. Chem. 2022, 10, 993691. [Google Scholar] [CrossRef]
  63. Hong Phuong, P.; Cam Anh, H.; Tri, N.; Phung Anh, N.; Cam Loc, L. Effect of Support on Stability and Coke Resistance of Ni-Based Catalyst in Combined Steam and CO2 Reforming of CH4. ACS Omega 2022, 7, 20092–20103. [Google Scholar] [CrossRef]
  64. Visser, N.L.; Daoura, O.; Plessow, P.N.; Smulders, L.C.J.; De Rijk, J.W.; Stewart, J.A.; Vandegehuchte, B.D.; Studt, F.; Van Der Hoeven, J.E.S.; De Jongh, P.E. Particle Size Effects of Carbon Supported Nickel Nanoparticles for High Pressure CO2 Methanation. ChemCatChem 2022, 14, e202200665. [Google Scholar] [CrossRef]
  65. Alli, R.D.; Zhou, R.; Mohamedali, M.; Mahinpey, N. Effect of Thermal Treatment Conditions on the Stability of MOF-Derived Ni/CeO2 Catalyst for Dry Reforming of Methane. Chem. Eng. J. 2023, 466, 143242. [Google Scholar] [CrossRef]
  66. Bian, Z.; Kawi, S. Highly Carbon-Resistant Ni–Co/SiO2 Catalysts Derived from Phyllosilicates for Dry Reforming of Methane. J. CO2 Util. 2017, 18, 345–352. [Google Scholar] [CrossRef]
  67. Manabayeva, A.M.; Mäki-Arvela, P.; Vajglová, Z.; Martinéz-Klimov, M.; Tirri, T.; Baizhumanova, T.S.; Grigor’eva, V.P.; Zhumabek, M.; Aubakirov, Y.A.; Simakova, I.L.; et al. Dry Reforming of Methane over Ni–Fe–Al Catalysts Prepared by Solution Combustion Synthesis. Ind. Eng. Chem. Res. 2023, 62, 11439–11455. [Google Scholar] [CrossRef]
  68. Ibrahim, A.A.; Fakeeha, A.H.; Abasaeed, A.E.; Wazeer, I.; Bentalib, A.; Siva Kumar, N.; Abu-Dahrieh, J.K.; Al-Fatesh, A.S. The Effectiveness of Ni-Based Bimetallic Catalysts Supported by MgO-Modified Alumina in Dry Methane Reforming. Catalysts 2023, 13, 1420. [Google Scholar] [CrossRef]
  69. Álvarez Moreno, A.; Ramirez-Reina, T.; Ivanova, S.; Roger, A.-C.; Centeno, M.Á.; Odriozola, J.A. Bimetallic Ni–Ru and Ni–Re Catalysts for Dry Reforming of Methane: Understanding the Synergies of the Selected Promoters. Front. Chem. 2021, 9, 694976. [Google Scholar] [CrossRef] [PubMed]
  70. Abasaeed, A.E.; Ibrahim, A.A.; Fakeeha, A.H.; Bayazed, M.O.; Amer, M.S.; Abu-Dahrieh, J.K.; Al-Fatesh, A.S. Ni-Co Bimetallic Catalysts Supported on Mixed Oxides (Sc-Ce-Zr) for Enhanced Methane Dry Reforming. ChemistryOpen 2024, 13, e202400086. [Google Scholar] [CrossRef]
  71. Zhang, J.; Xie, K.; Jiang, Y.; Li, M.; Tan, X.; Yang, Y.; Zhao, X.; Wang, L.; Wang, Y.; Wang, X.; et al. Photoinducing Different Mechanisms on a Co-Ni Bimetallic Alloy in Catalytic Dry Reforming of Methane. ACS Catal. 2023, 13, 10855–10865. [Google Scholar] [CrossRef]
  72. Cao, D.; Luo, C.; Tan, Z.; Luo, T.; Shi, Z.; Wu, F.; Li, X.; Zheng, Y.; Zhang, L. Carbon Deposition Properties and Regeneration Performance of La2NiO4 Perovskite Oxide for Dry Reforming of Methane. J. Environ. Chem. Eng. 2023, 11, 111022. [Google Scholar] [CrossRef]
  73. Du, Z.; Petru, C.; Yang, X.; Chen, F.; Fang, S.; Pan, F.; Gang, Y.; Zhou, H.-C.; Hu, Y.H.; Li, Y. Development of Stable La0.9Ce0.1NiO3 Perovskite Catalyst for Enhanced Photothermochemical Dry Reforming of Methane. J. CO2 Util. 2023, 67, 102317. [Google Scholar] [CrossRef]
  74. Jiang, Q.; Xin, Y.; Xing, J.; Cao, Y.; Sun, F.; Xing, X.; Hong, H.; Xu, C.; Jin, H. Enhanced Oxygen Migration in Tailored Lanthanum-Based Perovskite for Solar-Driven Dry Reforming of Methane. Fuel 2024, 363, 130852. [Google Scholar] [CrossRef]
  75. Gil-Muñoz, G.; Alcañiz-Monge, J. Examining the Effect of Zirconium Doping in Lanthanum Nickelate Perovskites on Their Performance as Catalysts for Dry Methane Reforming. J. Environ. Chem. Eng. 2025, 13, 115387. [Google Scholar] [CrossRef]
  76. Li, C.; Li, G.; Zhan, Y.; Liu, C.; Tang, Y.; Yan, D.; Li, J.; Jia, L. An Engineering Approach for Active and Stable Dry Methane Reforming within Solid Oxide Fuel Cell Stacks Using La0.6Sr0.2Cr0.85Ni0.15O3-δ as a Catalyst. Int. J. Hydrog. Energy 2025, 98, 1079–1086. [Google Scholar] [CrossRef]
  77. Nair, M.M.; Kaliaguine, S.; Kleitz, F. Nanocast LaNiO3 Perovskites as Precursors for the Preparation of Coke-Resistant Dry Reforming Catalysts. ACS Catal. 2014, 4, 3837–3846. [Google Scholar] [CrossRef]
  78. Liu, C.; Tang, W.; Zhong, Y.; Abuelgasim, S.; Xu, C.; Abdalazeez, A. Dry Reforming of Methane by LaNiO3 Perovskite Oxide: Cu-Substituted Improving Reactivity and Stability. Int. J. Hydrogen Energy 2025, 101, 617–626. [Google Scholar] [CrossRef]
  79. Muñoz, H.J.; Korili, S.A.; Gil, A. Surface Tuning of a Highly Crystalline Ni/LaAlO3 Perovskite Catalyst Obtained from Aluminum Saline Slags Using Various Synthesis Methods for the Dry Reforming of Methane. Catal. Today 2025, 447, 115158. [Google Scholar] [CrossRef]
  80. Sellam, D.; Ikkour, K.; Dekkar, S.; Messaoudi, H.; Belaid, T.; Roger, A.-C. CO2 Reforming of Methane over LaNiO3 Perovskite Supported Catalysts: Influence of Silica Support. Bull. Chem. React. Eng. Catal. 2019, 14, 568–578. [Google Scholar] [CrossRef]
  81. Oni, B.A.; Tomomewo, O.S.; Sanni, S.E.; Ojo, V.O. Dry Reforming of Methane with CO2 over Co-La1−xCaxNiO3 Perovskite-Type Oxides Supported on ZrO2. Mater. Today Commun. 2023, 36, 106802. [Google Scholar] [CrossRef]
  82. Winterstein, T.F.; Malleier, C.; Mohammadi, A.; Krüger, H.; Kahlenberg, V.; Venter, A.M.; Bekheet, M.F.; Müller, J.T.; Gurlo, A.; Heggen, M.; et al. Lanthanum Nickel Titanate Perovskites as Model Systems for Ni-Perovskite Interfacial Engineering in Methane Dry Reforming. Mater. Today Chem. 2025, 45, 102620. [Google Scholar] [CrossRef]
  83. Bian, Z.; Wang, Z.; Jiang, B.; Hongmanorom, P.; Zhong, W.; Kawi, S. A Review on Perovskite Catalysts for Reforming of Methane to Hydrogen Production. Renew. Sustain. Energy Rev. 2020, 134, 110291. [Google Scholar] [CrossRef]
  84. Qiu, P.; Yuan, J.; Cheng, K.; Xiong, C.; Wu, S. Enhanced Catalytic Activity for Methane Dry Reforming of Ce-Doped Ba0.9Zr0.7Y0.2Ni0.1O3-δ Perovskite Catalyst. Mol. Catal. 2023, 549, 113543. [Google Scholar] [CrossRef]
  85. Osti, A.; Costa, S.; Rizzato, L.; Senoner, B.; Glisenti, A. Photothermal Activation of Methane Dry Reforming on Perovskite-Supported Ni-Catalysts: Impact of Support Composition and Ni Loading Method. Catal. Today 2025, 449, 115200. [Google Scholar] [CrossRef]
  86. Joo, S.; Seong, A.; Kwon, O.; Kim, K.; Lee, J.H.; Gorte, R.J.; Vohs, J.M.; Han, J.W.; Kim, G. Highly Active Dry Methane Reforming Catalysts with Boosted in Situ Grown Ni-Fe Nanoparticles on Perovskite via Atomic Layer Deposition. Sci. Adv. 2020, 6, eabb1573. [Google Scholar] [CrossRef] [PubMed]
  87. Georgiadis, A.G.; Tsiotsias, A.I.; Siakavelas, G.I.; Charisiou, N.D.; Ehrhardt, B.; Wang, W.; Sebastian, V.; Hinder, S.J.; Baker, M.A.; Mascotto, S.; et al. An Experimental and Theoretical Approach for the Biogas Dry Reforming Reaction Using Perovskite-Derived La0.8X0.2NiO3-δ Catalysts (X = Sm, Pr, Ce). Renew. Energy 2024, 227, 120511. [Google Scholar] [CrossRef]
  88. Bhattar, S.; Abedin, M.A.; Kanitkar, S.; Spivey, J.J. A Review on Dry Reforming of Methane over Perovskite Derived Catalysts. Catal. Today 2021, 365, 2–23. [Google Scholar] [CrossRef]
  89. Muñoz, H.J.; Korili, S.A.; Gil, A. Facile Synthesis of an Ni/LaAlO3—Perovskite via an MOF Gel Precursor for the Dry Reforming of Methane. Catal. Today 2024, 429, 114487. [Google Scholar] [CrossRef]
  90. Rudolph, B.; Tsiotsias, A.I.; Ehrhardt, B.; Dolcet, P.; Gross, S.; Haas, S.; Charisou, N.D.; Goula, M.A.; Mascotto, S. Nanoparticle Exsolution from Nanoporous Perovskites for Highly Active and Stable Catalysts. Adv. Sci. 2023, 10, 2205890. [Google Scholar] [CrossRef]
  91. Liu, Z.; Zhou, J.; Sun, Y.; Yue, X.; Yang, J.; Fu, L.; Deng, Q.; Zhao, H.; Yin, C.; Wu, K. Tuning Exsolution of Nanoparticles in Defect Engineered Layered Perovskite Oxides for Efficient CO2 Electrolysis. J. Energy Chem. 2023, 84, 219–227. [Google Scholar] [CrossRef]
  92. Piazzolla, F.; Moraes, T.S.; Figueiredo, S.S.; De Paula, D.F.; Dos Santos Veiga, E.L.; Rodella, C.B.; Fonseca, F.C. Exsolution of Ni Nanoparticles from La0.4Sr0.4Ti0.8Ni0.2O3-δ Perovskite for Ethanol Steam Reforming. Catal. Today 2025, 444, 115011. [Google Scholar] [CrossRef]
  93. Shah, S.; Xu, M.; Pan, X.; Gilliard-Abdulaziz, K.L. Exsolution of Embedded Ni–Fe–Co Nanoparticles: Implications for Dry Reforming of Methane. ACS Appl. Nano Mater. 2021, 4, 8626–8636. [Google Scholar] [CrossRef]
  94. Osazuwa, O.U.; Ng, K.H. The Roles of Oxygen Mobility and Oxygen Vacancy in Metallic Catalysts-Prompted Dry Reforming of Methane: A Review. Renew. Energy 2025, 256, 124074. [Google Scholar] [CrossRef]
  95. Yang, C.; Grimaud, A. Factors Controlling the Redox Activity of Oxygen in Perovskites: From Theory to Application for Catalytic Reactions. Catalysts 2017, 7, 149. [Google Scholar] [CrossRef]
  96. Yu, K.; Lou, L.; Liu, S.; Zhou, W. Asymmetric Oxygen Vacancies: The Intrinsic Redox Active Sites in Metal Oxide Catalysts. Adv. Sci. 2020, 7, 1901970. [Google Scholar] [CrossRef] [PubMed]
  97. An, B.; Ma, Y.; Han, X.; Schröder, M.; Yang, S. Activation and Catalysis of Methane over Metal–Organic Framework Materials. Acc. Mater. Res. 2025, 6, 77–88. [Google Scholar] [CrossRef]
  98. Ong, J.L.; Loy, A.C.M.; Teng, S.Y.; How, B.S. Future Paradigm of 3D Printed Ni-Based Metal Organic Framework Catalysts for Dry Methane Reforming: Techno-Economic and Environmental Analyses. ACS Omega 2022, 7, 15369–15384. [Google Scholar] [CrossRef]
  99. Zheng, K.; Gao, X.; Xie, Y.; He, Z.; Ma, Y.; Hou, S.; Su, D.; Ma, X. Free-Standing Bimetallic Co/Ni-MOF Foams toward Enhanced Methane Dry Reforming under Non-Thermal Plasma Catalysis. J. Colloid Interface Sci. 2025, 683, 564–573. [Google Scholar] [CrossRef]
  100. Wang, J.; Qi, T.; Li, G.; Zhang, Y.; Chen, H.; Li, W. Elucidating the Promoting Mechanism of Coordination-Driven Self-Assembly MOFs/SiO2 Composite Derived Catalyst for Dry Reforming of Methane with CO2. Fuel 2022, 330, 125569. [Google Scholar] [CrossRef]
  101. Zhou, D.; Huang, H.; Cai, W.; Liang, W.; Xia, H.; Dang, C. Immobilization of Ni on MOF-Derived CeO2 for Promoting Low-Temperature Dry Reforming of Methane. Fuel 2024, 363, 130998. [Google Scholar] [CrossRef]
  102. Liu, H.; Dong, M.; Xiong, J.; Huang, Z.; Hou, H.; Liang, Y.; Lu, J. Study on Ce-MOF-Derived Oxides as Morphology-Tunable Catalyst Supports for Dry Reforming of Methane. Appl. Surf. Sci. 2025, 679, 161167. [Google Scholar] [CrossRef]
  103. Figueredo, G.P.; Medeiros, R.L.B.A.; Macedo, H.P.; De Oliveira, Â.A.S.; Braga, R.M.; Mercury, J.M.R.; Melo, M.A.F.; Melo, D.M.A. A Comparative Study of Dry Reforming of Methane over Nickel Catalysts Supported on Perovskite-Type LaAlO3 and Commercial α-Al2O3. Int. J. Hydrogen Energy 2018, 43, 11022–11037. [Google Scholar] [CrossRef]
  104. Salaev, M.A.; Liotta, L.F.; Vodyankina, O.V. Lanthanoid-Containing Ni-Based Catalysts for Dry Reforming of Methane: A Review. Int. J. Hydrogen Energy 2022, 47, 4489–4535. [Google Scholar] [CrossRef]
  105. Shao, J.; Li, C.; Fei, Z.; Liu, Y.; Zhang, J.; Li, L. MOFs-Derived Ni@ZrO2 Catalyst for Dry Reforming of Methane: Tunable Metal-Support Interaction. Mol. Catal. 2024, 558, 114028. [Google Scholar] [CrossRef]
  106. Abdulrasheed, A.; Jalil, A.A.; Gambo, Y.; Ibrahim, M.; Hambali, H.U.; Shahul Hamid, M.Y. A Review on Catalyst Development for Dry Reforming of Methane to Syngas: Recent Advances. Renew. Sustain. Energy Rev. 2019, 108, 175–193. [Google Scholar] [CrossRef]
  107. Chein, R.-Y.; Fung, W.-Y. Syngas Production via Dry Reforming of Methane over CeO2 Modified Ni/Al2O3 Catalysts. Int. J. Hydrogen Energy 2019, 44, 14303–14315. [Google Scholar] [CrossRef]
  108. Chiou, J.Y.Z.; Kung, H.-Y.; Wang, C.-B. Highly Stable and Active Ni-Doped Ordered Mesoporous Carbon Catalyst on the Steam Reforming of Ethanol Application. J. Saudi Chem. Soc. 2017, 21, 205–209. [Google Scholar] [CrossRef]
  109. Zhao, S.; Cai, W.; Li, Y.; Yu, H.; Zhang, S.; Cui, L. Syngas Production from Ethanol Dry Reforming over Rh/CeO2 Catalyst. J. Saudi Chem. Soc. 2018, 22, 58–65. [Google Scholar] [CrossRef]
  110. Cao, D.; Zeng, F.; Zhao, Z.; Cai, W.; Li, Y.; Yu, H.; Zhang, S.; Qu, F. Cu Based Catalysts for Syngas Production from Ethanol Dry Reforming: Effect of Oxide Supports. Fuel 2018, 219, 406–416. [Google Scholar] [CrossRef]
  111. Dewa, M.; Elharati, M.A.; Hussain, A.M.; Miura, Y.; Song, D.; Fukuyama, Y.; Furuya, Y.; Dale, N.; Zhang, X.; Marin-Flores, O.G.; et al. Metal-Supported Solid Oxide Fuel Cell System with Infiltrated Reforming Catalyst Layer for Direct Ethanol Feed Operation. J. Power Sources 2022, 541, 231625. [Google Scholar] [CrossRef]
  112. Xu, Y.; Li, X.; Gao, J.; Wang, J.; Ma, G.; Wen, X.; Yang, Y.; Li, Y.; Ding, M. A Hydrophobic FeMn@Si Catalyst Increases Olefins from Syngas by Suppressing C1 By-Products. Science 2021, 371, 610–613. [Google Scholar] [CrossRef]
  113. Senapati, S.; Ray, K.; Pradhan, N.C. An Energy-Efficient Aspen plus Model for H2-Rich Syngas Production via Dry Reforming of Ethanol: A Thermodynamic Analysis. Int. J. Hydrogen Energy 2025, 98, 1107–1118. [Google Scholar] [CrossRef]
  114. Attili, R.; Viet Nguyen Vo, D.; Sabri Mahmud, M. Kinetic Study of Ethanol Dry Reforming Using Lanthanum Copper Perovskite. Mater. Today Proc. 2023, in press. [Google Scholar] [CrossRef]
  115. Drif, A.; Bion, N.; Brahmi, R.; Ojala, S.; Pirault-Roy, L.; Turpeinen, E.; Seelam, P.K.; Keiski, R.L.; Epron, F. Study of the Dry Reforming of Methane and Ethanol Using Rh Catalysts Supported on Doped Alumina. Appl. Catal. Gen. 2015, 504, 576–584. [Google Scholar] [CrossRef]
  116. Hou, T.; Lei, Y.; Zhang, S.; Zhang, J.; Cai, W. Ethanol Dry Reforming for Syngas Production over Ir/CeO2 Catalyst. J. Rare Earths 2015, 33, 42–45. [Google Scholar] [CrossRef]
  117. Chen, Q.; Cai, W.; Liu, Y.; Zhang, S.; Li, Y.; Huang, D.; Wang, T.; Li, Y. Synthesis of Cu-Ce0.8Zr0.2O2 Catalyst by Ball Milling for CO2 Reforming of Ethanol. J. Saudi Chem. Soc. 2019, 23, 111–117. [Google Scholar] [CrossRef]
  118. Ramkiran, A.; Vo, D.-V.N.; Mahmud, M.S. Syngas Production from Ethanol Dry Reforming Using Cu-Based Perovskite Catalysts Promoted with Rare Earth Metals. Int. J. Hydrogen Energy 2021, 46, 24845–24854. [Google Scholar] [CrossRef]
  119. Wei, Y.; Cai, W.; Deng, S.; Li, Z.; Yu, H.; Zhang, S.; Yu, Z.; Cui, L.; Qu, F. Efficient Syngas Production via Dry Reforming of Renewable Ethanol over Ni/KIT-6 Nanocatalysts. Renew. Energy 2020, 145, 1507–1516. [Google Scholar] [CrossRef]
  120. Zhukova, A.; Fionov, Y.; Chuklina, S.; Mikhalenko, I.; Fionov, A.V.; Isaikina, O.; Zhukov, D.Y.; De Lima, A.M. CO2 Reforming of Ethanol over Ni/Al2 O3 -(Zr–Yb)O2 Catalysts: The Effect of Zr:Al Ratio on Nickel Activity and Carbon Formation. Energy Fuels 2024, 38, 482–498. [Google Scholar] [CrossRef]
  121. Cao, D.; Cai, W.; Li, Y.; Li, C.; Yu, H.; Zhang, S.; Qu, F. Syngas Production from Ethanol Dry Reforming over Cu/Ce0.8Zr0.2O2 Catalyst. Catal. Lett. 2017, 147, 2929–2939. [Google Scholar] [CrossRef]
  122. Shi, G.; Wang, Y.; Cao, Y.; Cai, W.; Tan, F. Synthesis of High-Performance Ni/Ce0.8Zr0.2O2 Catalyst via Co-Nanocasting Method for Ethanol Dry Reforming. Korean J. Chem. Eng. 2020, 37, 2143–2151. [Google Scholar] [CrossRef]
  123. Li, F.; Dong, J.; Wang, M.; Lin, X.; Cai, W.; Liu, X. Ethanol Dry Reforming over Ordered Mesoporous Co-Zn Composite Oxide for Syngas Production. Korean J. Chem. Eng. 2022, 39, 1744–1752. [Google Scholar] [CrossRef]
  124. Li, T.; Li, F.; Nginyo, J.; Cai, W.; Yu, B. Syngas Production through Dry Reforming of Ethanol over Co@SiO2 Catalysts: Effect of SiO2 Shell Thickness. Mol. Catal. 2023, 547, 113307. [Google Scholar] [CrossRef]
  125. Cai, W.; Dong, J.; Chen, Q.; Xu, T.; Zhai, S.; Liu, X.; Cui, L.; Zhang, S. One-Pot Microwave-Assisted Synthesis of Cu-Ce0.8Zr0.2O2 with Flower-like Morphology: Enhanced Stability for Ethanol Dry Reforming. Adv. Powder Technol. 2020, 31, 3874–3881. [Google Scholar] [CrossRef]
  126. Fionov, Y.; Khlusova, K.; Chuklina, S.; Mushtakov, A.; Fionov, A.; Zhukov, D.; Averin, A.; Zhukova, A. High-Performance Ni/Al2O3-(Zr + Ce)O2 Catalysts for Syngas Production via Ethanol Dry Reforming. Fuel 2024, 376, 132685. [Google Scholar] [CrossRef]
  127. Fedorova, V.; Bespalko, Y.; Arapova, M.; Smal, E.; Valeev, K.; Prosvirin, I.; Sadykov, V.; Parkhomenko, K.; Roger, A.; Simonov, M. Ethanol Dry Reforming over Bimetallic Ni-Containing Catalysts Based on Ceria-Zirconia for Hydrogen Production. ChemCatChem 2023, 15, e202201491. [Google Scholar] [CrossRef]
  128. Zhukova, A.; Fionov, Y.; Semenova, S.; Khaibullin, S.; Chuklina, S.; Maslakov, K.; Zhukov, D.; Isaikina, O.; Mushtakov, A.; Fionov, A. Ethanol Dry Reforming for Hydrogen-Rich Syngas Production over Cu-Promoted Ni/Al2 O3 –ZrO2 Catalysts. J. Phys. Chem. C 2024, 128, 20177–20194. [Google Scholar] [CrossRef]
  129. Park, S.-W.; Lee, D.; Kim, S.-I.; Kim, Y.J.; Park, J.H.; Heo, I.; Chang, T.S.; Lee, J.H. Effects of Alkali Metals on Nickel/Alumina Catalyzed Ethanol Dry Reforming. Catalysts 2021, 11, 260. [Google Scholar] [CrossRef]
  130. Li, F.; Wang, S.; Li, T.; Tian, Y.; Wang, M.; Cai, W. Effect of Calcination Temperature on the Performance of SiO2@Co@CeO2 Catalyst in CO2 Reforming With Ethanol. Catal. Lett. 2023, 153, 3712–3723. [Google Scholar] [CrossRef]
  131. Labhasetwar, N.; Saravanan, G.; Kumar Megarajan, S.; Manwar, N.; Khobragade, R.; Doggali, P.; Grasset, F. Perovskite-Type Catalytic Materials for Environmental Applications. Sci. Technol. Adv. Mater. 2015, 16, 036002. [Google Scholar] [CrossRef]
  132. Eremeev, N.F.; Hanna, S.A.; Sadovskaya, E.M.; Leonova, A.A.; Bulavchenko, O.A.; Ishchenko, A.V.; Prosvirin, I.P.; Sadykov, V.A.; Bespalko, Y.N. Catalysts for Ethanol Dry Reforming Based on High-Entropy Perovskites. J. Catal. 2025, 445, 116028. [Google Scholar] [CrossRef]
  133. Tian, Y.; Li, T.; Cai, W. Ethanol Reforming with CO2 over MOF-Derived CoMOx (M:Ce,Zr,Zn): Impact of Support on Activity/Stability. J. Mater. Sci. 2023, 58, 16033–16045. [Google Scholar] [CrossRef]
  134. Wang, M.; Li, T.; Tian, Y.; Zhang, J.; Cai, W. CO2 Reforming with Ethanol Over Bimetallic Co(Ni)/ZnO Catalyst with Enhanced Activity: Synergistic Effect of Ni and Co. Catal. Lett. 2024, 154, 3829–3838. [Google Scholar] [CrossRef]
  135. Rao, Z.; Cao, Y.; Huang, Z.; Yin, Z.; Wan, W.; Ma, M.; Wu, Y.; Wang, J.; Yang, G.; Cui, Y.; et al. Insights into the Nonthermal Effects of Light in Dry Reforming of Methane to Enhance the H2/CO Ratio Near Unity over Ni/Ga2 O3. ACS Catal. 2021, 11, 4730–4738. [Google Scholar] [CrossRef]
  136. Tian, Y.; Li, Y.; Cao, D.; Liu, Q.; Cai, W. Photothermal Dry Reforming of Ethanol over Rod-like Ni/ZrO2 Catalysts. Mol. Catal. 2024, 564, 114308. [Google Scholar] [CrossRef]
  137. Li, T.; Tian, Y.; Nginyo, J.; Difuma Luis, D.I.; Cai, W. Light-Assisted Ethanol Dry Reforming over NiZnOx Hollow Microspheres with Enhanced Activity and Stability. Renew. Energy 2024, 227, 120514. [Google Scholar] [CrossRef]
  138. Shafiqah, M.-N.N.; Tran, H.N.; Nguyen, T.D.; Phuong, P.T.T.; Abdullah, B.; Lam, S.S.; Nguyen-Tri, P.; Kumar, R.; Nanda, S.; Vo, D.-V.N. Ethanol CO2 Reforming on La2O3 and CeO2-Promoted Cu/Al2O3 Catalysts for Enhanced Hydrogen Production. Int. J. Hydrogen Energy 2020, 45, 18398–18410. [Google Scholar] [CrossRef]
  139. Wang, M.; Li, F.; Chen, Q.; Cai, W. Ethanol Dry Reforming over Mn-Doped Co/CeO2 Catalysts with Enhanced Activity and Stability. Energy Fuels 2021, 35, 13945–13954. [Google Scholar] [CrossRef]
  140. Arapova, M.; Smal, E.; Bespalko, Y.; Fedorova, V.; Valeev, K.; Cherepanova, S.; Ischenko, A.; Sadykov, V.; Simonov, M. Ethanol Dry Reforming over Ni Supported on Modified Ceria-Zirconia Catalysts– the Effect of Ti and Nb Dopants. Int. J. Hydrogen Energy 2021, 46, 39236–39250. [Google Scholar] [CrossRef]
  141. Wang, M.; Li, F.; Dong, J.; Lin, X.; Liu, X.; Wang, D.; Cai, W. MOF-Derived CoCeOx Nanocomposite Catalyst with Enhanced Anti-Coking Property for Ethanol Reforming with CO2. J. Environ. Chem. Eng. 2022, 10, 107892. [Google Scholar] [CrossRef]
  142. Choi, D.S.; Kim, N.Y.; Yoo, E.; Kim, J.; Joo, J.B. Enhanced Coke Resistant Ni/SiO2@SiO2 Core–Shell Nanostructured Catalysts for Dry Reforming of Methane: Effect of Metal-Support Interaction and SiO2 Shell. Chem. Eng. Sci. 2024, 299, 120480. [Google Scholar] [CrossRef]
  143. Moradi, M.J.; Iranshahi, D. Investigating the Performance of Ni/Al2O3-ZrO2 Thin Film Nanocatalysts Synthesized by Pvd Method for Methane Dry Reforming. Results Eng. 2025, 27, 106625. [Google Scholar] [CrossRef]
Figure 1. Current distribution of syngas utilization worldwide [9].
Figure 1. Current distribution of syngas utilization worldwide [9].
Applsci 15 10722 g001
Figure 2. Number of publications on the topic of “dry reforming of methane” in the Scopus database (2015–15 May 2025).
Figure 2. Number of publications on the topic of “dry reforming of methane” in the Scopus database (2015–15 May 2025).
Applsci 15 10722 g002
Figure 3. Stability of the 23Ni-7Mo/YSZ catalyst during dry reforming of methane (a); TGA results of tested Ni-Mo-based catalysts (b); XRD diffractograms (c); TEM images of 23Ni-7Mo/YSZ (d). Reproduced with permission [50].
Figure 3. Stability of the 23Ni-7Mo/YSZ catalyst during dry reforming of methane (a); TGA results of tested Ni-Mo-based catalysts (b); XRD diffractograms (c); TEM images of 23Ni-7Mo/YSZ (d). Reproduced with permission [50].
Applsci 15 10722 g003
Figure 4. Number of publications on the topic “CO2 reforming of ethanol” in the Scopus database (2015–15 May 2025).
Figure 4. Number of publications on the topic “CO2 reforming of ethanol” in the Scopus database (2015–15 May 2025).
Applsci 15 10722 g004
Figure 5. Ni/xAl2O3-(Zr+Ce)O2 catalyst after the ethanol dry reforming reaction: (a) TEM images; (b) Raman spectra; (c) Thermogravimetric analysis (TGA) Reproduced with permission from [126].
Figure 5. Ni/xAl2O3-(Zr+Ce)O2 catalyst after the ethanol dry reforming reaction: (a) TEM images; (b) Raman spectra; (c) Thermogravimetric analysis (TGA) Reproduced with permission from [126].
Applsci 15 10722 g005
Figure 6. Stability analysis of Ni/Al2O3 and Ni/M2O-Al2O3 (M = Li, Na, K) catalysts during the EDR process: (1) TEM images of carbon deposits on the catalyst surface: (a) Ni/Al2O3; (b) Ni/Li₂O-Al2O3; (c) Ni/Na2O-Al2O3; (d) Ni/K2O-Al2O3; (2) TGA profiles of post-reaction mass changes; (3) Raman spectra of spent catalysts Reprinted from [129].
Figure 6. Stability analysis of Ni/Al2O3 and Ni/M2O-Al2O3 (M = Li, Na, K) catalysts during the EDR process: (1) TEM images of carbon deposits on the catalyst surface: (a) Ni/Al2O3; (b) Ni/Li₂O-Al2O3; (c) Ni/Na2O-Al2O3; (d) Ni/K2O-Al2O3; (2) TGA profiles of post-reaction mass changes; (3) Raman spectra of spent catalysts Reprinted from [129].
Applsci 15 10722 g006
Figure 7. SEM images of spent catalysts: (a) SiO2@Co@ CeO2-550; (b) SiO2@Co@CeO2 700; (c) SiO2@Co@CeO2-850. Reprinted from [130].
Figure 7. SEM images of spent catalysts: (a) SiO2@Co@ CeO2-550; (b) SiO2@Co@CeO2 700; (c) SiO2@Co@CeO2-850. Reprinted from [130].
Applsci 15 10722 g007
Figure 8. TEM images of catalysts: (a) LaCuO3 and (b) CeCuO3 after the ethanol dry reforming reaction. Reproduced with permission from [118].
Figure 8. TEM images of catalysts: (a) LaCuO3 and (b) CeCuO3 after the ethanol dry reforming reaction. Reproduced with permission from [118].
Applsci 15 10722 g008
Table 1. Main and side reactions occurring in the DRM process.
Table 1. Main and side reactions occurring in the DRM process.
NoReaction NameEquationΔH° 298k (kJ/mol)
1Dry reforming of methaneCO2 + CH4 → 2CO + 2H2+247.0
2Reverse Water-Gas Shift reactionCO2 + H2 → CO + H2O41.0
3Boudouard reaction2CO → C + CO2−171.0
4Decomposition of methaneCH4 → C + 2H275.0
5Carbon monoxide reductionCO + H2 → C + H2O−131.3
6Hydrogenation of CO2CO2 + 2H2 → C + 2H2O−90.0
Table 3. Comparative table of perovskite catalysts for DRM.
Table 3. Comparative table of perovskite catalysts for DRM.
CatalystMethod of SynthesisProcess Conditions (T, GHSV)Feed Gas RatioConversion, % (CH4/CO2)Key FeaturesRef
La2NiO4Sol–gel850 °C,
24,000 h−1
CH2:CO2 = 1:183/92High dispersion metallic Ni0, coke reduction[72]
La0.9Ce0.1NiO3Pechini method700 °C,
300 l·h−1·gcat−1
10% CO2/10% CH4/80% Ar80/71Oxygen vacancies and electron-hole pairs[73]
LaNiO3Sol–gel800 °C,
15 l·h−1·gcat−1
CH2:CO2 = 1:197/95High resistance to coke, Ni dispersion[74]
LaZrxNi1-xO3Citrate method700 °C, 58L/(g.h.)25/25/50 CH4/CO2/He)77/86Resistance to sintering, activity at extreme temperatures[75]
LSCrN@NFSelf-combustion method750 °C,
10 mL min−1
45% CH4–45% CO2–10% N280/83Optimized basicity, increased CO2 adsorption[76]
LaNi0.8Cu0.2O3Sol–gel700 °C,
250 l·h−1·gcat−1
CH4:CO2:Ar = 10:10:183/85Low activation energy, Cu-Ni synergy[78]
10 wt % Ni/LaAlO3Impregnation700 °C,
1.2 10−4 cm3/(h g)
CH2:CO2 = 1:187/86Strong interaction of active centers with the carrier[79]
2%Co- La0.2Ca0.8NiO3-ZrO2Co-precipitation method700 °C,
42,000 h−1
CH4: CO2:N2(Ar) = 1:1:188/90Metal-carrier interaction and metal synergy[81]
La0.8Sm0.2NiO3Citrate sol–gel method750 °C,
200,000 h−1
Ar:CH4:CO2 = 2.5:1.5:140/78Reduced ordering/crystallinity
of deposited coke
[87]
Table 4. Key characteristics of MOF-derived catalysts for dry reforming of methane.
Table 4. Key characteristics of MOF-derived catalysts for dry reforming of methane.
CatalystNi Particle Size (nm)Coke Formation Rate (mg·g−1·h−1)CH4/CO2
Conversion at 700 °C (%)
Stability
(Particle Growth, %)
Specific
Surface Area (m2/g)
Key FeaturesRef
5.3 wt.%Ni-3.1 wt.% MgO@mSiO2 (MOF-derived)7.2 → 18.3 (60 h)1.25 (vs. 7.47 for analog)96 → 75/75 → 69+154%n/aSiO2 mesopores; OH groups; Ni–O–Si anchoring (phyllosilicate interphase)[100]
20 wt.% Ni-Ce-BTC (N2-pyr)2.19 → 18.3 (120 h)2.88 (vs. 65.4 with CO2 treatment)n/a+740%24537% Ce3+, protective carbon layer, high proportion of oxygen vacancies[65]
4.57 wt.% Ni/CeO2-M17.2 → 17.9 (10 h)1.25 (vs. 7.47 for Ni/CeO2-C)30.8/40.1 (550 °C)+4%22.4CeO2 nanorods, ID/IF = 0.18[101]
10 wt.% Ni/LaAlO35.03 → 18.3 (60 h)1.25 (vs. 7.47 for sol–gel analog)75/80+264%33Perovskite structure, oxygen vacancies, mesopores[79]
13.45 wt.% Ni/ZrO2-B20.47 → 16.6 (6 h)3.8 wt.% (vs. 6.9 wt.% for A; 0.7 wt.% for C)n/a−19%12.93Oxygen vacancies ZrO2, 25 nm mesopores, strong metal-support interaction[105]
Notes: Coke accumulation (Ni/ZrO2 series): wt% of coke relative to catalyst mass after reaction (TGA on spent catalysts). Particle growth: percentage increase in Ni particle size over the indicated reaction time (from TEM/XRD).
Table 5. Possible reactions of ethanol dry reforming and their equations.
Table 5. Possible reactions of ethanol dry reforming and their equations.
Reaction NameEquationΔH° 298k (kJ/mol)
1The main reaction of EDRC2H5OH + CO2 → 3CO +3H2+296.7
2Dehydrogenation of ethanolC2H5OH → CH3CHO + H2+68.5
3Decomposition of ethanolC2H5OH → CO + CH4 + H2+49.6
4Acetaldehyde decompositionCH3COH→ CH4 + CO−18.9
5Dry reforming of acetaldehydeCH3COH + CO2 → 2H2 + 3CO−186.3
6Methanation of CO2CO2 + 4H2 → CH4 + 2H2O−153.0
7Methanation of COCO2 + 3H2 → CH4 + H2O−206.2
8Dry reforming of methaneCH4 + CO2 → 2CO + 2H2+247.0
Table 6. Recent advances in EDR catalysts (last 5 years).
Table 6. Recent advances in EDR catalysts (last 5 years).
CatalystMethod of PreparationReaction ConditionsReduction ConditionsActivityStabilityRef
15 wt.% Cu-Ce0.8Zr0.2O2co-precipitationT = 700 °C, EtOH/CO2 = 1:1; volumetric velocity = 14,000 mL/gcat/hin situ T = 500 °C for 1 h, 5 vol.% H2/N2 (30 mL/min)XC2H5OH = 90%,
yield CO = 37%,
yield H2 = 22%,
yield CH4 = 28%
50 h[125]
3wt.%La–10 wt.% Cu/Al2O3wet impregnationT = 750 °C,
EtOH/CO2 = 1:1
n/a-XC2H5OH = 87.6%
XCO2 = 55.1%
yield CO = 27%
yield H2 = 52%
-[138]
Mn-15 wt.% Co/CeO2urea co-precipitationT = 650 °C,
EtOH/CO2/N2 = 1:1:2
volumetric velocity = 14,000 mLgcat−1 h−1).
n/aXC2H5OH = 97%
Yield H2 = 45 mol.%
yield CO = 45 mol.%
50 h[139]
LaCuO3 (bulk perovskite)citrate sol–gelT = 750 °C
volumetric velocity = 42 L/gcat/h
n/aXC2H5OH = 93.7%
yield H2 = 54%,
yield CO = 50%
-[118]
5 wt.%Ni/Ce0.75Zr0.25- x(Nb,Ti)xO2-δsolvothermal methodEtOH/CO2 =1:1T = 650 °C for 1 h,
5 vol.% H2/N2 (100 mL/min)
XCO2 = 60–80%-[140]
8.09 wt.% CoCe-MOFsolvothermal methodEtOH/CO2/N2 = 1:1:1.5,
GHSV = 27,700 mLgcat−1 h−1
in situ T = 500 °C for 1 h,
5 vol.% H2/N2 (40 mL/min)
XC2H5OH = 97%40 h[141]
13.7 wt.% CoZnO-HThard template methodT = 650 °C,
EtOH/CO2/N2 = 1:1:1
T = 500 °C for 1 h, 5 vol.% H2/N2XC2H5OH = 100%,
XCO2 = 25%
yield H2 = 30 mol.%
yield CO = 40 mol.%
40 h[123]
SiO2@Co@CeO2electrostatic adsorptionT = 500 °C,
EtOH/CO2/N2 = 1:1:1.5,
GHSV = 27,700 mLg−1 h−1
T = 500 °C for 1 h, H2/N2 5 vol.% (40 mL/min)XC2H5OH = 92%
yield H2 = 55.1 mol.%
yield CO = 17.8 mol.%
15 h[130]
10 wt.% Ni/ACZsol–gelT = 650 °C,
EtOH:CO2 = 1:1.8
n/aXC2H5OH = 100%,
XCO2 = 63%
7 h[126]
9.3 wt.% NiZnOx-Mtemplate methodEtOH/CO2/N2 = 1:1:3in situ T = 500 °C for 1 h, 5 vol.% H2/N2XC2H5OH = 80%100 h[137]
10 wt.% Ni/35AZsol–gelT = 600 °C,
CO2/C2H5OH = 1.4:1
n/ayield H2 = 47%
yield = CO 34%
-[120]
12 wt.% Ni/CoZn-MOFMOF precursor-based methodEtOH/CO2/N2 = 1:1:1,
14,000 mLg−1 h−1
5 vol.% H2/N2 (40 mL/min)XC2H5OH = 100%
yield H2 = 47%
-[134]
Ni/ZrO2precipitation methodT = 450 °C,
EtOH:CO2 = 1:1
n/aXC2H5OH = 70%40 h[136]
5 wt.% Ni/La(Cr0.2Mn0.2Fe0.2Co0.2Ni0.2)O3 (LCMFCN)citrate methodT = 700 °C,
EtOH/CO2 = 1:1
T = 600 °C for 1 h, 5% H2+Ar
2 vol.% EtOH + 2 vol.% CO2 + Ar, T = 600–750 °C
XC2H5OH = 99.7%
yield H2 = 50%
yield CO = 87%
5 h[132]
Table 7. Advantages and disadvantages of DRM and EDR.
Table 7. Advantages and disadvantages of DRM and EDR.
ProcessAdvantagesDisadvantages
DRMUtilizes two greenhouse gases simultaneously (CH4 and CO2);
Produces syngas with H2/CO ≈ 1 (suitable for Fischer–Tropsch and oxygenate synthesis);
Well-studied with extensive literature and pilot trials
Requires very high temperatures (≥700 °C);
Severe coking via CH4 cracking and Boudouard reaction;
Ni sintering at elevated temperatures; H2/CO ratio often needs adjustment for downstream use
EDRCombines renewable ethanol with CO2 utilization (bio-based, sustainable); Operates at milder temperatures (300–600 °C);
Produces syngas with H2/CO ≈ 1, closer to downstream optimum;
Attractive integration into biorefinery concepts
Complex reaction network (dehydration, dehydrogenation, condensation); Formation of oxygenates and polymeric coke precursors;
Catalyst deactivation through coke and side-products;
Still an emerging technology with limited pilot-scale validation
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mambetova, M.; Anissova, M.; Myltykbayeva, L.; Makayeva, N.; Dossumov, K.; Yergaziyeva, G. Catalyst Development for Dry Reforming of Methane and Ethanol into Syngas: Recent Advances and Perspectives. Appl. Sci. 2025, 15, 10722. https://doi.org/10.3390/app151910722

AMA Style

Mambetova M, Anissova M, Myltykbayeva L, Makayeva N, Dossumov K, Yergaziyeva G. Catalyst Development for Dry Reforming of Methane and Ethanol into Syngas: Recent Advances and Perspectives. Applied Sciences. 2025; 15(19):10722. https://doi.org/10.3390/app151910722

Chicago/Turabian Style

Mambetova, Manshuk, Moldir Anissova, Laura Myltykbayeva, Nursaya Makayeva, Kusman Dossumov, and Gaukhar Yergaziyeva. 2025. "Catalyst Development for Dry Reforming of Methane and Ethanol into Syngas: Recent Advances and Perspectives" Applied Sciences 15, no. 19: 10722. https://doi.org/10.3390/app151910722

APA Style

Mambetova, M., Anissova, M., Myltykbayeva, L., Makayeva, N., Dossumov, K., & Yergaziyeva, G. (2025). Catalyst Development for Dry Reforming of Methane and Ethanol into Syngas: Recent Advances and Perspectives. Applied Sciences, 15(19), 10722. https://doi.org/10.3390/app151910722

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop