Previous Article in Journal
Multiscale Models to Evaluate the Impact of Chemical Compositions and Test Conditions on the Mechanical Properties of Cement Mortar for Tile Adhesive Applications
Previous Article in Special Issue
Growth of Carbon Nanofibers and Carbon Nanotubes by Chemical Vapour Deposition on Half-Heusler Alloys: A Computationally Driven Experimental Investigation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photo-Thermal Dry Reforming of Methane with PGM-Free and PGM-Based Catalysts: A Review

by
Alessio Varotto
1,2,*,
Umberto Pasqual Laverdura
1,
Marta Feroci
2 and
Maria Luisa Grilli
1,*
1
Energy Technologies and Renewable Sources Department, Italian National Agency for New Technologies, Energy and Sustainable Economic Development (ENEA), Via Anguillarese 301, 00123 Rome, Italy
2
Department of Fundamental and Applied Sciences for Engineering (SBAI), Sapienza University of Rome, Via Castro Laurenziano, 7, 00161 Rome, Italy
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(15), 3809; https://doi.org/10.3390/ma17153809 (registering DOI)
Submission received: 17 June 2024 / Revised: 18 July 2024 / Accepted: 22 July 2024 / Published: 1 August 2024

Abstract

:
Dry reforming of methane (DRM) is considered one of the most promising technologies for efficient greenhouse gas management thanks to the fact that through this reaction, it is possible to reduce CO2 and CH4 to obtain syngas, a mixture of H2 and CO, with a suitable ratio for the Fischer–Tropsch production of long-chain hydrocarbons. Two other main processes can yield H2 from CH4, i.e., Steam Reforming of Methane (SRM) and Partial Oxidation of Methane (POM), even though, not having CO2 as a reagent, they are considered less green. Recently, scientists’ challenge is to overcome the many drawbacks of DRM reactions, i.e., the use of precious metal-based catalysts, the high temperatures of the process, metal particle sintering and carbon deposition on the catalysts’ surfaces. To overcome these issues, one proposed solution is to implement photo-thermal dry reforming of methane in which irradiation with light is used in combination with heating to improve the efficiency of the process. In this paper, we review the work of several groups aiming to investigate the pivotal promoting role of light radiation in DRM. Focus is also placed on the catalysts’ design and the progress needed for bringing DRM to an industrial scale.

1. Introduction

Currently, there is an increasing concentration of greenhouse gases (GHGs) in the atmosphere caused by anthropogenic activities, leading to global warming [1], especially due to CO2 and CH4. Correlations between the annual increase in temperature and GHGs have been analyzed over the years by many research centers, including NASA [2,3,4]. Figure 1 shows the chart of global temperature anomalies, which are deviations from a reference value or long-term average, and the increasing CO2 emissions in the period from 1959 to 2022. Despite the ten-year variation of some tenths of degrees, it must be considered that even 1 or 2 °C can have a huge impact on the global environment [5].
The escalating levels of CO2 and CH4 concentrations in the atmosphere have intensified concerns regarding climate change [6]. The Paris Agreement was signed by 196 parties who committed to reduce GHG emissions to limit global warming well below 2 °C above pre-industrial levels, increasing the ability to adapt to the adverse impacts of climate change and create forest climate resilience and low-GHG-emission development [7]. The target goals are to reduce CO2 emissions by at least 30 gigatons ( G t C O 2 )/yr [8] using alternative energy sources and, at the same time, employing carbon capture, utilization and storage (CCUS), reducing costs and increasing energy efficiency. Europe must achieve these goals by 2030 to reduce GHG emissions, and this is part of the “Fit for 55” legislation that is part of the European Green Deal to achieve climate neutrality by 2050. During the last COP28 (Conference of the Parties) in Dubai, more than 200 parties discussed the world’s first “global stock-take” to ratchet up climate change action before the end of the decade, with the overarching aim to keep the global temperature increase below 1.5 °C. One of the most important points of this COP is that the agreement on mitigation measures now includes a fossil fuel phase-out instead of only a phase-down. This is the first COP in which fossil fuels were designated as the principal cause of climate change.
Dry reforming of methane (DRM) constitutes a valuable approach for effective CO2 and CH4 conversion to added-value syngas (H2 and CO), even though it is considered at present to be far from effectively mitigating the environmental issues related to GHG emissions because of the many challenges limiting its widespread application at the industrial scale.
Many review papers on dry reforming of methane reactions can be found in the literature and are focused on particular aspects, such as the catalysts’ active materials, the supports, the catalysts’ design, the catalysts’ deactivation by carbon deposition, etc. With this review paper on photo-thermal dry reforming of methane preceded by only a few previous ones [9,10,11,12,13,14,15,16], we aim to investigate the pivoting role of light irradiation focusing both on platinum group metal-free catalysts and catalysts containing platinum group metals as active elements.
This review will start with a description of the dry reforming of methane reaction, followed by a description of the other reactions forming syngas from methane (bi- and tri-reforming of methane). The principles of photo-thermal catalysis (PTC) will be introduced before the photo-thermal dry reforming of methane (PTC-DRM) description, with PTC-DRM being the focus of the present paper. Then, the catalysts for PTC-DRM, both containing platinum group metals (PGMs) and without PGMs, will be reviewed.

2. Dry Reforming of Methane (DRM) Reaction

In the dry reforming of methane, CO2 and CH4 react to obtain a syngas rich in CO and H2 [17]. The DRM reaction was studied for the first time by Fischer and Tropsch in 1928 using nickel and cobalt catalysts, and nowadays, it is proposed as a potential solution to convert CO2 on a large scale [8,18]. Equation (1) reports the DRM reaction and Figure 2 schematizes the DRM reaction promoted by a source of heat. The main characteristics of the DRM are that, according to the thermodynamics [19], it may yield a H2/CO ratio close to unity and is highly endothermic, so high temperatures (>650 °C) must be used to promote the DRM reaction, even in the presence of a catalyst [18].
CH 4 + CO 2     2   H 2 + 2   CO Δ H 298   K 0 = 247   k J / m o l
As the variation in enthalpy is positive, for Le Chatelier’s principle, high temperatures are needed to transform the reagents into syngas, as evident from the Gibbs Equation (2) [19,20,21]:
Δ G f 0 = Δ H f 0 Δ S f 0 × T
where Δ H f 0 is the variation in enthalpy formation, equal to 247 kJ/mol, and Δ S f 0 is the variation in entropy formation, equal to 0.26 kJ/(mol·K) [22]. Δ G f 0 is negative when T > 950 K [23].
It is important to underline that the H2/CO ratio in the DRM reaction is always lower than unity due to the reverse water gas shift (RWGS) equilibrium reaction (3):
CO 2 + H 2     CO + H 2 O   Δ H 298   K 0 = 41   k J / m o l
where part of the hydrogen formed during the DRM reaction reacts with carbon dioxide to form carbon monoxide and water [24]. Equations (1) and (3) are spontaneous when the temperature is above 650 °C, and this means that the RWGS is inevitable [25].
High temperatures are needed because of the high energies required to break the first C-H bond in methane (750 kJ/mol), which is the rate-determining step of the entire reaction, and the first C=O bond in carbon dioxide (476 kJ/mol).
To decrease the activation energies and thus the temperature needed to promote the DRM reaction, catalysts are generally used. The reaction can also be promoted by using external sources of light. Thus, as schematized in Figure 3, the DRM reaction can occur thermocatalytically (TC), photocatalytically (PC) and photothermocatalytically (PTC).
The first classical method employs heat to promote the DRM reaction, the second one uses a light source, typically a Xe or a Hg lamp or a concentrated sunlight, and in the third one, both heat and light are combined to increase the efficiency of the reaction. The last approach, i.e., the synergistic effect of light and heat, seems to be more powerful than the sum of its parts [26,27,28,29].
In the following sections, the fundamentals of photo-thermal (applying the principles of PC and TC) dry reforming of methane will be discussed [24,30,31,32,33]. Photo-thermal dry reforming of methane is superior compared to the other two pathways proposed; for thermocatalysis occurring at high temperatures (the reaction is endothermic), byproducts such as coke may be generated, and only an opportune catalyst design can mitigate the effect of catalyst deactivation by carbon deposition. At low temperatures, even if carbon deposition is reduced, to achieve high catalytic efficiencies, the assistance of a light source is necessary. On the other hand, a light source without heat assistance does not lead to high catalytic efficiency. Therefore, temperature plays a significant role in the DRM reaction, as reported by many authors. As an example, Sokefun et al. [33] have investigated the effect of Ru loading and reduction temperature of Ru-Ni-Mg/Ceria-Zirconia catalysts, achieving dry reforming activity between 450 and 510 °C, even though conversion efficiencies were lower in comparison to the values generally obtained at higher temperatures.
Figure 3. Photo-catalysis and thermo-catalysis contributions to PTC. During photo-catalysis, the activation energy barrier (Ea, red line) is reduced by the effect of light (Ea*, blue line), while the temperature effect is to provide additional thermal energy (kBT, green zone) to the energy path of an endothermic reaction. Figure reprinted with permission from [34]. Adapted from [35]. Copyright 2008 Royal Society of Chemistry.
Figure 3. Photo-catalysis and thermo-catalysis contributions to PTC. During photo-catalysis, the activation energy barrier (Ea, red line) is reduced by the effect of light (Ea*, blue line), while the temperature effect is to provide additional thermal energy (kBT, green zone) to the energy path of an endothermic reaction. Figure reprinted with permission from [34]. Adapted from [35]. Copyright 2008 Royal Society of Chemistry.
Materials 17 03809 g003

3. Bi- and Tri-Reforming of Methane

The DRM reaction converts the two GHGs CH4 and CO2 into syngas; however, two other main reactions can be used to convert methane into syngas: Steam Reforming of Methane (SRM) and Partial Oxidation of Methane (POM), which are reported in Equations (4) and (5):
C H 4 + H 2 O   CO + 3 H 2 Δ H 298 K 0 = 206   k J / m o l
C H 4 + 0.5   O 2   CO + 2 H 2 Δ H 298 K 0 = 36   k J / m o l
The SRM reaction provides syngas from methane and water, while the POM reaction uses oxygen in place of H2O [36].
When Equations (1) and (4) are combined together, in one step (6), the process is called bi-reforming of methane (BRM) or CSDRM (i.e., combined SRM and DRM). When Equations (1), (4) and (5) are combined together, in one step, the process is called tri-reforming of methane (TRM) (7).
3 C H 4 + 2 H 2 O + C O 2   4 CO + 8 H 2 Δ H 298 K 0 = 206   k J / m o l
3 C H 4 + 0.5   O 2 + C O 2 + H 2 O   4 CO + 7 H 2 Δ H 298 K 0 = 104   k J / m o l
BRM enhances the H2 production using SRM, due to the equilibrium in the water gas shift reaction. The combination of dry reforming with steam reforming leads to a syngas with flexible H2/CO ratios via easy adjustment of feed-stock composition and suitable for the Fisher–Tropsch production of methanol, oxo-alcohols and other chemicals. In addition, in BRM, carbon deposition on catalysts’ surface [37] is reduced.
Tahir et al. [38] synthesized a Co/HC3N4 catalyst with different Co loading and performed a photo-activity test using a Hg lamp with light intensity of 100 mW/cm2 at different feed mixtures of CH4/CO2. HC3N4 was used because of its ability to exploit solar energy, while cobalt was used to increase the active sites and capture the charge carriers. The highest production rates of H2 and CO were obtained in the case when 2% wt. of Co was used, reaching 689 and 95.8 µmol/g, respectively, after 4 h of irradiation time. In both DRM and BRM reactions, production rate decreased in the presence of water.
As in a DRM reaction, in a TRM reaction, oxides such as CeO2, ZrO2 and MgO are also generally used [39]. Due to the presence of CO2 as a reactant, support needs a high oxygen storage capacity, as it increases the conversion of carbon dioxide due to the basic character of these oxides [40]. Oxide supports are chosen due to their high thermal stability and their high oxygen storage capacity, and it is even possible to use them as photocatalysts for one of these reactions.
Despite BRM and TRM showing very promising potential because of their higher H2/CO ratios and reduced carbon deposition on catalysts’ surface, several challenges, as in DRM, exist for their scalability [37].
Figure 4 reports the trends of publications about DRM, PTC-DRM, BRM and TRM starting from 1994, and ref. [41] reports a bibliometric study on dry reforming of methane.

4. Fundamentals of Photo-Thermal Catalysis

In this section, the focus is on the fundamentals of Photo-Thermal Catalysis (PTC) to analyze how catalysts’ performances are improved by light absorption. In PTC, absorption of photons by a material promotes electrons from the valence to the conduction band and generates electron–hole pairs, which will participate in the subsequent redox reactions.
PTC involves the interaction between photons, electrons and phonons, giving rise to photo-thermal and/or Localized Surface Plasmon Resonance (LSPR) effects [42,43].
While PTC utilizes lattice vibrations induced by light to achieve the photo-thermal conversion, the LSPR mechanism generates an electric field enhancement by plasmons and it utilizes electron dispersion to generate hot carriers, i.e., charge carriers (electrons) that dissipate energy in the form of heat. The LSPR mechanism occurs when resonant conditions are satisfied, i.e., when the energy of the external source of light matches the work function of the metal nanoparticles on the surface. In this way, resonant frequencies of conduction electrons and incident electromagnetic waves are enhanced on the surface of plasmonic nanoparticles, in so-called hot spots. The excess energy can decay through several pathways, radiative or non-radiative. In non-radiative decay, electrons will generate hot charge carriers in the plasmonic structure through intra- (s-to-s transitions) or interband (d-to-s transitions) mechanisms (this phenomenon is called “Landau damping”). In this case, electron–hole pair excitations and electron–electron collisions occur. To facilitate the promotion of electrons from the valence band to the conduction band, it is important to use materials with a narrow band gap. In fact, one of the main challenges in photo-catalysis is to avoid electron–hole recombination to facilitate the formation of products [44,45]. For plasmon-induced catalysis, several reaction mechanisms have been proposed on the basis of their time scale with or without an adsorbate or a semiconductor (Figure 5) [46].
Indirect and direct hot electron transfer (HET) into a semiconductor is the topic of a very recent review [45]. In Figure 6, the indirect and direct charge transfer (CT) mechanisms are represented in the case of a plasmon-induced catalysis. The indirect electron transfer is a two-step process. In the first step, hot carriers (electrons and holes) are generated, while in the second one, there is the transfer of electrons and, depending on the existence of a single-component or a hetero-structured plasmonic photocatalyst, this can happen through a metal/adsorbate system (Figure 6a) or a metal/semiconductor system (Figure 6b,c). In the latter case, the electron transfer occurs through a Schottky barrier and the trapping of the transferred hot electrons increases their lifetime, because they do not travel back to the metal nanostructure.
The disadvantage of indirect electron transfer is the necessity to have electrons near the Fermi level. On the contrary, direct electron transfer is a single-step process, and it is assumed that energy loss is lower. However, high electric fields are needed for the direct electron transfer process to occur.
More insights about PTC are available in the literature [28,44,45,46,47,48,49,50].
In a reaction process involving light excitation, the efficiency η defined in (8) and the quantum efficiency Q.E. defined in (9) must be considered:
η = ( r C O × Δ c H C O 0 + r H 2 × Δ c H H 2 0 r C H 4 × Δ c H C H 4 0 )   P i l l u m i n a t i o n × 100 %
where P i l l u m i n a t i o n is the light power used, r i represents the reaction rate, Δc H i 0 is the standard combustion heat and i represents products and reagents, with Δc H C O 2 0 being equal to zero.
Q . E . = the   number   of   reacted   electron   s   ( n ) the   number   of   incident   photons   ( N ) × 100 %
The quantum efficiency Q.E. is the ratio of the reacted electrons and the number of incident photons N, which can be calculated using (10):
N = I × S × t × λ h c
where I is the light power per area,   S the surface area, t the time during which the catalyst is exposed to the light source, λ the wavelength, h the Planck’s constant and c the velocity of light [51].
In the next paragraph, the synergistic effect of light and heat on the DRM reaction will be discussed.
Figure 6. (a) Indirect and direct charge transfer mechanisms (HET—hot electron transfer; CT—charge transfer). A: metal, B: adsorbate. Figure modified from [49] (Creative Commons CC BY). (b) Indirect electron transfer and (c) direct electron transfer in a metal/semiconductor interface. Figures modified with permission from [46].
Figure 6. (a) Indirect and direct charge transfer mechanisms (HET—hot electron transfer; CT—charge transfer). A: metal, B: adsorbate. Figure modified from [49] (Creative Commons CC BY). (b) Indirect electron transfer and (c) direct electron transfer in a metal/semiconductor interface. Figures modified with permission from [46].
Materials 17 03809 g006

5. Photo-Thermal Dry Reforming of Methane

Solar-driven photo-thermal dry reforming of methane is considered among the most promising processes to convert GHGs into valuable chemicals. However, it suffers from some drawbacks that currently limit its application at the industrial scale, such as the high cost of electricity, the carbon tax, the limited number of hours during which natural sunlight is used to meet the energetic requirements of the reaction, the cost of the photo-catalyst and the catalyst lifetime [52,53,54,55]. Considering these issues [48], the requirements to obtain an efficient PTC-DRM are:
-
high visible light absorption;
-
good match between the band structure of the catalyst and the redox potentials of the reaction;
-
good thermocatalytic activity.
In Figure 7, the description of a proposed simplified PTC-DRM mechanism is presented in which the principal steps are shown. Methane is dissociated by the active metal of the catalyst into CHx and H* species, and H* species combine to produce H2 (Figure 7a). CO2 is activated by spillover on the catalyst’s surface (Figure 7b) and reacts with H* and CH* species to produce CO molecules and to release H2O, respectively. Methane activation is promoted by holes, and electrons promote the cleavage of the C-O bond in carbon dioxide.
In the literature, two main PTC mechanisms are proposed: monofunctional and bifunctional. It is hypothesized that in the monofunctional PTC mechanism, methane and carbon dioxide are activated on metal sites, following the typical Langmuir–Hinshelwood–Hougen–Watson mechanism; instead, in the bifunctional PTC mechanism, methane is activated on the metal sites and carbon dioxide on the surface of the support [56,57].
According to Langmuir–Hinshelwood–Hougen–Watson mechanism, in a bimolecular reaction such as DRM, three elementary steps are considered. The first one is the adsorption of the gases (A(g) and B(g)) on adjacent sites, the second step is the reaction of molecules at the surface of the support and the third one is the desorption of the products (11):
A(g) + B(g) →  A(ads) + B(ads) →  P
Light irradiation enhances, as heat, every step of the PTC mechanism, with the main differences that, compared to dark conditions, methane dissociation is faster [58].
In the study reported by Liu et al. [27], a catalyst with magnesium/aluminum-layered double hydroxides (Mg/Al-LDH) was used with Ni and Co alloys as active species. After light irradiation, methane decomposes into CHx and H* species, and in a second step, C* and H* species are obtained before the desorption of the products. The activation energy required by using NiCo alloy was higher compared to that for catalysts employing only Ni or Co as active metals, to prevent the formation of carbon on the surface of the catalyst. This is one of the most accepted pathways for the DRM reaction; low activation energies of CH* oxidation compared to the oxidation of C* to CO* disfavor the deposition of carbon on the surface of catalysts.
Going deep into the DRM reaction, the value of the GHSV (Gas Hourly Space Velocity), i.e., the volume of a gas entering a reactor per hour per unit volume of catalyst, must also be considered. GHSV is calculated by the following Equation (12):
GHSV = [Ftot·3600·(s·h−1)]/(Vcat)
where Ftot is the total flow rate used in the experiment and Vcat is the volume of the catalyst. GHSV is expressed as h−1. The inverse of this value is the space time τ (13), which represents the residence time of the gasses with catalyst:
τ = GHSV−1
The yield of the DRM reaction increases by increasing the residence time (decreasing the GHSV) due to the higher number of adsorbed reactants on the catalytic surface.
PTC is performed, usually, in a fixed bed reactor or in a stainless-steel reactor in which the catalyst is contained in a quartz tube to guarantee homogeneous heating. After the introduction of the sample into the reactor, the feed ratio of the gases is fixed at a certain temperature and the catalyst, if necessary, is reduced by hydrogen.
A simplified scheme of a type of reactor for PTC-DRM is presented in Figure 8.
In this representation, a reactor containing the catalyst is positioned inside a tubular vertical furnace and a source of light is used to promote the DRM reaction. After conversion, the produced gases are analyzed by a gas analyzer.
When the DRM is carried out at “low” temperatures, i.e., at T < 600 °C, the yields of the products are low and, in addition, at temperatures lower than 325 °C, the disproportion of carbon dioxide is favored. Disproportionation of CO2 produces different types of carbon species, which cause the deactivation of the catalysts [21,59]. On the other hand, starting from T > 650 °C, methane cracking is favored, especially on Ni-based catalysts. The formation of solid carbon C(s) during CO2 reforming of CH4 may therefore occur either via CH4 cracking (14) or CO disproportionation (i.e., the Boudouard Reaction (15)) [24]:
CH 4     2   H 2   + C ( s ) Δ H 298 K 0 = 75   k J / m o l
2 CO     CO 2   + C ( s ) Δ H 298 K 0 = 172.5   k J / m o l
Figure 9 illustrates the formation of a coke deposit on a catalyst surface. It is evident how carbon deposits hinder the catalytic activity, blocking the active sites, which cannot be used by active metals [59].
Another crucial point is to understand how the carbon, derived from CO2 and CH4 adsorption, affects the lowering of the activation energies of the elementary steps [23]. In a recent review from Xu and Park [59], the main deactivation mechanisms of Ni-based catalysts are discussed, together with strategies to limit coke formation to improve the catalysts’ performance and stability.

6. Catalysts for PTC-DRM Reaction

In this last section, catalysts for the DRM reaction have been divided into two main classes: PGM-based and PGM-free catalysts, according to the active metal components.
As already said, PGM stands for Platinum Group Metal, i.e., platinum, palladium, rhodium, ruthenium, iridium and osmium. These metals have high catalytic activity due to their high heat and corrosion resistance, which make them almost unique for a vast range of industrial, medical and electronic applications. However, PGMs are listed among the critical raw materials (CRMs) for the EU [61], and many efforts are being devoted to their recovery from end-of-life products or to their partial or total substitution. It is worth noting that the support also plays a fundamental role in determining catalysts’ overall performance, because the support can provide additional catalytic sites, good dispersion of the active metal, increased surface area, etc., and the interaction between the support and the active component is fundamental in determining the catalytic activity and the selectivity of the catalysts [62].

6.1. PGM-Free Catalysts

The PGM-free catalysts for DRM are mostly based on Ni or Co as monometallic or bimetallic active compounds dispersed on different types of supports, mainly oxides [63,64,65,66]. It is worth nothing that Ni was recently added to the CRM list, while Co, which has been on the CRM list since 2011, presents additional cancerogenic issues [67]. Ni and Co cost is, however, much lower than that of PGMs [68].
In a recent study, Hu et al. [69] synthesized a catalyst for PTC-DRM containing Ni supported on Al2O3, which showed 142.8 and 160.2 mmol g N i 1 h−1 yields of H2 and CO, respectively. Photo-thermal conditions helped in the formation of Ni(0) and in obtaining hydrogen molecules starting from methane. However, catalysts supported on acidic materials, such as alumina, rapidly deactivate due to coke deposition, and this is one of the reasons why supports like CaO are often used [57]. In fact, CaO is a promising support for the DRM reaction, because of its high coke resistance, good dispersion of metal and strong metal–support interaction. Moreover, it is cheap, easily available, non-corrosive and environmentally friendly [70]. CaO is used to provide basic sites for the adsorption of CO2, but rarely, CaO is used alone, because it reacts with CO2 to form CaCO3. This is the reason why CaO is generally combined with other oxides to enhance the catalytic activity. ZrO2 is used with CaO to increase the activity of the catalyst, and it exhibits a strong metal–support interaction and stabilizes oxygen vacancies in the ZrOx lattice. A very recent study on the modification of CaO-based adsorbents and Ni-based catalysts for DRM and CaL (calcium looping) was recently reported by Wang et al. [71]. Even SiO2 is an excellent choice because of its inertness, its low cost and low toxicity [72].
A catalyst with (Ni/CeO2)-SiO2 was demonstrated to be a useful obstacle to carbon deposition and sintering, due to its silica shell. From the Scherrer equation, the Ni sizes were 3.5 nm in the catalyst submitted to the DRM test. These values were close to the ones for the reduced catalyst, confirming the non-occurrence of Ni sintering. These findings were also confirmed by TEM images. Despite the fact that carbon deposition was confirmed by TGA analysis, its amount was very low, about 0.5% of the total amount of catalyst. The loss of carbon occurred in the temperature range of 800–1073 K and the reduction of nickel ion to metal was confirmed by XPS analysis. The fast CO2 activation was attributed to the reduction of Ce4+ to Ce3+, which led to a higher number of oxygen vacancies [73].
The creation of oxygen vacancies can be explained easily using cerium oxide as an example. Ce4+ leads to a re-distribution of the charge during the reduction of the ions; when CO2 molecules interact with the surface, the electronic density increases, CO2 accepts an electron and the adjacent oxygen vacancy acts as a sink to detach and accommodate the oxygen anion, replenishing the vacancy. At the end, the CO2 energy activation barrier for the breaking of the bond between C and O is reduced [55,74,75,76].
Perovskites are being extensively investigated for the DRM reaction and have been demonstrated to have high activity, even though challenges still exist for their application at the industrial scale due to carbon deposition and thermal stability. Perovskites are based on the general formula ABO3/A2BO4, where A is usually a rare earth, alkaline earth or alkali metal ion, and B is a transition metal cation. The catalytic properties of perovskites are significantly affected by the choice of A and B metal cations, by the preparation techniques and by the partial substitution of A and B cations. A recent review paper on dry reforming of methane over perovskite-derived catalysts is reported in [77]. As oxides, perovskites can handle high-temperature processes [57,78]. The insertion of calcium in a perovskite structure can improve oxygen storage capacity by creating oxygen defects, leading to a relatively high surface-to-volume ratio.
To study oxygen vacancies, two important techniques can be used: Electron Paramagnetic Resonance (EPR), in which, at a g value (specific value calculated for this technique) of 2.004, a high density of oxygen vacancies in the catalyst is evidenced, and X-ray Photoelectron Spectroscopy (XPS), by analyzing the O 1s region, where the three peaks around 530 eV can be attributed to oxygen lattice, vacancies and adsorbates. The oxygen vacancy-to-lattice oxygen ratio gives insights into the diffusion during the charge transfer [79,80,81].
Bimetallic alloys are very effective in enhancing the performances of the catalysts during heterogeneous catalysis. Zhang et al. [26] synthesized NiCo alloys on SiO2 supports and observed that CO2 conversion increased when light changed from high to low wavelengths, i.e., at higher energy. In their study, Field Emission Microscopy (FEM) was very useful to understand the role of hot carriers in PTC. 1.2 Ni0.3Co/SiO2 showed the highest values of quantum efficiency (Equation (9)) of 67 and 58% at 420 and 500 nm, respectively.
Takami et al. achieved conversion rates of 21% for CO2 and 20% for CH4 even at 473 K in a DRM reaction under visible light with a plasmonic Ni/Al2O3 photocatalyst. The production rates of CO and H2 were 1.87 and 1.20 mmol h−1, respectively [82].
Production rate can be calculated using (15) [51]:
r i = ( C o u t l e t × V o u t l e t × 60 ) / ( 22.4 × m c a t )
where C o u t l e t and V o u t l e t are, respectively, concentration and volume of the gas, m c a t is the mass of the catalyst and r i is the outlet from the reactor.
In [28], Xie et al. compared three different experimental conditions, thermo-, photo- and photo-thermal, and used TiO2 as support, which is commonly used for photothermal catalysts. In their study, the authors reported how the synergistic effect of light and heat improves the formation rates of the products and the conversions of CO2 and CH4, mostly due to the separation of the excited electron/hole pairs.
Doping with alkaline metals or their oxide can increase the activity of catalysts due to their abilities to enhance the adsorption and activation of CO2 molecules. From the Arrhenius plot, apparent activation energy is reduced, passing from 57.2 kJ/mol under dark conditions to 29.5 kJ/mol under light conditions.
Fertout et al. [83] developed Ni catalysts on γ alumina doped with La2O3 and alkaline earth oxides (MgO, CaO and SrO), and they promoted the formation of carbonates, instead of hydroxides on these modified supports. Catalytic activity increased with the basicity of metals: Mg < Ca < Sr, probably because of a faster activation of CO2. Strontium has strong basicity, but magnesium improves the dispersion of Ni by surface rearrangement and calcium enhances the thermal stability and suppresses carbon deposition, even though it decreases the catalytic activity.
Khan et Tahir [84] developed TiO2 NP-embedded 2D Ti3C2 exfoliated sheets coupled with g-C3N4 to construct a 2D/2D g-C3N4/Ti3C2 heterojunction. The performance of the semiconductor was examined to analyze CO and H2 production under visible light. The graphitic carbon nitride (g-C3N4) after irradiation with an artificial light source acted as an electron donor. CO decomposition of methane was achieved by electrons donated by methane; holes, CHx species and H+ ions reacted with methane to form a H2 molecule.
Tahir et al. [85] used g-C3N4 combined with Cu and observed that an increase in the amount of copper led to an increase of yields of the products. It is obvious from this study that the introduction of copper in the structure of graphitic carbon nitride is the key to obtain a H2/CO ratio close to unity. The catalyst’s structure was successful in promoting the DRM reaction because Cu/g-C3N4 decreased the CO selectivity due to a more prolific production of electrons.
Table 1 reports the DRM performances of PGM-free catalysts on several types of supports. In Table 2, the performances of catalysts exposed to an external source of light are reported.
To better compare the different types of catalysts, the turnover frequency (T.O.F.) should also be considered. In [93], T.O.F. is expressed by two different Equations (17) and (18):
T.O.F. = specific rate (converted molecules·s−1·g−1)/L (active sites·g−1)
T.O.F. = areal site (converted molecules·s−1·nm−2)/ds (active sites·nm−2)
Specific rate refers to converted molecules per second per mass of sample. L is expressed as active sites per mass of sample and ds is expressed as active sites per surface area of sample (19).
L = d s   ( active   sites / nm 2 ) × A   ( nm 2 / g )
where ds is the surface density of active sites and A is the surface area of the catalyst.
The difficulty in calculating T.O.F. lies in the exact determination of active sites, which is why the definition of T.O.F. is still debated [94].
If T.O.F. increases, the conversions of reactants also increase, and this means that photo-thermal conditions are more efficient than thermal conditions. As an example, in [73], it was found that at the same temperature of 750 °C, in the presence of (Ni/CeO2)-SiO2 catalyst, the conversions of CO2 and CH4 using the photo-thermal approach were 42.8 and 19%, respectively, while using only the thermal approach, conversion rates were much lower, i.e., 10.5 and 6.4 for carbon dioxide and methane, respectively.
In the case of a structure-insensitive reaction, T.O.F. is constant. If the catalytic reaction is structure-sensitive, T.O.F. depends on the size of the catalyst particles [95].
In the work from Vogt et al. [96], the catalyst is formed by different-sized Ni nanoparticles on a SiO2 support to prevent the formation of carbon on its surface. By changing the morphology of the catalyst, two different conditions for SRM and DRM are found. In the first one, T.O.F. increases with the increase in Ni particle size. The explanation is that bonds are cleaved preferentially over highly under-coordinated atoms in the metal nanoparticles. A change in the morphology of the sample will contribute to a change in the value of T.O.F., which depends directly on the rate-determining step, and this is also related to the intermediates involved during the DRM reaction. As an example, FTIR experiments evidenced that there are peaks attributed to the presence of CO molecules adsorbed in two different positions, top and bridge, and the ratio of the two peaks corresponding to COads-top and COads-bridge is correlated with the T.O.F trend.
In Table 3 the effect of different supports on Ni-based catalysts is reported.

6.2. PGM-Based Catalysts for Photo-Thermal DRM

In this section, performances of catalysts containing Platinum Group Metals as active metals for PTC-DRM will be discussed. These precious metals can improve radically the catalytic activity of the DRM reaction, and they have a strong resistance to carbon deposition [19,99].
However, as already mentioned, also for these catalysts the support plays a fundamental role in the photocatalytic response, as we can see in the next two examples.
Yao et al. [74] developed a highly efficient photothermal Rh/LaNiO3 catalyst for solar-driven DRM, obtaining high generation rates for H2 and CO of 452.3 and 527.26 mmol·gRh−1·h−1, respectively, under the irradiation of a 300 W Xe lamp of 1.5 W/cm2, without external heating.
In a previous work from the same group, Yang et al. [100] developed an all-in-one photothermal and photoelectric catalytic DRM process employing a Rh/CexWO3 catalyst. Also, in this case, the experiment was carried out under irradiation of a Xe lamp without external heating. For 1.8 W/cm2 irradiation, the generation rates of H2 and CO were 88.5 and 152.3 mmol·gRh−1·h−1, respectively.
In both studies, Rh was uniformly dispersed on the oxide supports and Eact (H2) and Eact (CO) were 43.89 and 39.99 kJ/mol for Rh/CexWO3 and 115.3 and 87.5 kJ/mol for Rh/LaNiO3 catalyst, respectively. In both situations, Rh provided a low electronic density and this led to the promotion of breaking of methane bonds and to the formation of H2. The formation of oxo-bridges lowered the start-up temperature of the DRM reaction, the temperature at which the reaction becomes spontaneous, according to Equation (2). In both studies, a migration of oxygen was reported, from Ce to W in the Rh/CexWO3 catalyst, and from La to Ni in the Rh/LaNiO3 catalyst. The weaker the binding in the oxo-bridge, the higher the migration of oxygen atoms. This is the main reason why Eact was lower in the case that CexWO3 was used as a support. The higher production rates obtained in the case of a catalyst with LaNiO3 support were attributed to the high number of oxygen defects able to effectively activate CO2 and enhance the interaction between the support and the Rh nanoparticles.
Yin and coworkers [101], who used a catalyst made of Ru and Ni alloys on an Al2O3 support, obtained very high values of production rates of H2 and CO: 539 and 642 mmol·g−1·h−1, respectively. In their work, the light irradiation and the thermal decomposition of carbonates are used to promote the DRM reaction. Methane reacts with CO2 produced by thermal decomposition of MgCO3, producing MgO and syngas. MgCO3 decomposes completely into MgO at 400 °C in 5% CH4/Ar, as evident from the XRD analysis. Generation rates of H2 and CO increase during the first 20 min up to 0.68 and 0.44 mmol·g−1·h−1, respectively, and selectivity of carbon dioxide decreases by 20% after one hour with the consumption of MgCO3. The increase in generation rates is due to the formation of carbonates, external light irradiation and localized surface plasmon resonance effect.
Song et al. [102], who used a Pt-Au/SiO2 catalyst (Pt content= 0.5–1.5 wt.%), obtained H2 and CO generation rates of, respectively, 5.5 and 7.2 mmol·g−1h−1 at 400 °C using a light intensity of 0.6 μW·cm−2. This catalyst presents a high surface area of 436 m2/g and the aggregate structure made of Pt and Au is thermodynamically stable. The temperature is comparable to the one used for the case of Yin et al. [101], but there is a huge difference in the value of light intensity used in these two studies. In the study of Yin, in fact, the intensity of light (13.5 W/cm2) is much higher than that used in the study of Song (0.6 μW/cm2). This explains the difference in the values of generation rates of products.
In all of the above-mentioned studies, SEM analysis has evidenced that catalytic metals are well dispersed on each support and no agglomeration is found on the surface of the catalysts used.
One of the most used supports to perform PTC-DRM is anatase TiO2 due to its narrow band gap, its ability to suppress carbon deposition and its good redox properties for oxygen mobility [99].
Zhang et al. [58] have investigated the performances of a Pt/mesoporous-TiO2 catalyst under different experimental conditions: reaction temperature, intensity of light and CO2/CH4 molar ratio. Authors found that a moderate increase in CO2 partial pressure was beneficial to the further dissociation of CH4, while the total amount of dissociated H species diminished with CO2/CH4 greater than 1.08/1. For CO formation, the authors proposed two possible pathways; the first one consists of the decomposition of HCOO- and the second one consists of the dissociation of the CO2 molecule.
Wu et al. [103] reported two different situations for Pt nanocrystals partially confined in mesoporous CeO2 nanorods (Pt/CeO2-MNR). In the first one, under focused UV-Vis-IR illumination, r (H2) and r (CO) were, respectively, 5.7 and 6.0 mmol g−1 min−1, while under visible-IR illumination, r (H2) and r (CO) were 4.9 and 5.2 mmol g−1 min−1, respectively. In the latter case, authors found an improved photoactivation which inhibits CO disproportion as a major side reaction of coke formation, promoting the oxidation of carbon species produced by CH4 dissociation and thus increasing the H2/CO ratio.
Another study shows the effect of the structure of the support [104]. In this work, Pt is combined with two different supports, i.e., a mesoporous TiO2 and P25. Higher production rates are obtained in the first case because of the mesoporous structure, which shows a different light absorption response, a different recombination of electrons and holes and a narrower band gap. This explains why at the same temperature of 500 °C and at the same intensity of light of 4.76 W/cm2, the production rates of H2 and CO are 178.6 and 281 mmol· g P t 1 ·h−1 and 39.8 and 105.6 mmol· g P t 1 ·h−1 in the case of mesoporous titania and P25, respectively.
In the study from Zhang et al. [105], the photo-thermal properties of a Pt-Au/P25 composite catalyst have been enhanced with the LSPR effect. The authors found that plasmonic nanoparticles act as electron traps, inhibiting the recombination of electron–hole pairs and promoting the PTC-DRM reaction. In the case of Pt and Au, their behavior was identical due to their similar atomic number. Even if Pt-Au/P25 had the smallest value of surface area, the catalyst had a good catalytic activity due to the Fermi energy levels of both metals, which helped to lower the valence band (2.67 eV). This led to an increase in generation rates, which increased linearly with the light intensity increase, and to the promotion of a better migration of holes. In this study the Pt-Au/P25 catalyst had a strong metal–surface interaction, and no deactivation during a 10 h DRM test was observed. The DRM results show how light can improve yields of products, 85.38 and 201.92 mmol· g P t 1 ·h−1 for H2 and CO, respectively, favoring the dissociation of methane and CO2, while lower values were obtained in the case that only heat was used, 27.60 mmol· g P t 1 ·h−1 for H2 and 97 mmol· g P t 1 ·h−1 for CO.
Platinum is a very efficient catalytic metal for the DRM reaction [32,80,106,107]. The DRM reaction can be better performed on the active sites of precious metals due to their higher electron cloud density being able to reinforce the redox reactions on the surface of the catalysts. By combining precious metals and metals such as nickel and cobalt, high yields of products can be achieved [64]. It can be noted that the sample “0.3 Ni”, reported in Table 4, has the highest value of BET surface area but deactivates very rapidly when compared to the other catalysts. Comparison of 0.2, 0.3 and 0.4 PdNi bimetallic catalysts shows that increasing the amount of Pd, two situations arise:
-
the dispersion of Ni decreases;
-
sintering becomes more significant, and this leads, inevitably, to a decrease in the BET surface area.
Table 4. BET results on different catalysts. Feed ratio of CH4/CO2 is 1/1.
Table 4. BET results on different catalysts. Feed ratio of CH4/CO2 is 1/1.
Name SampleTemperature
°C
Light Intensity
W/cm2
BET Surface Area
m2/g
Pore Volume
cm3/g
Average Pore Diameter
nm
Catalytic PerformancesReferences
P255004.663.00.127.9N.A.[105]
Pt/P255004.662.00.149.1r (H2) = 65 mmol·g−1·h−1
r (CO) = 162 mmol·g−1·h−1
H2/CO = 0.40
[105]
Au/P255004.660.00.149.3N.A.[105]
Pt-Au/P255004.655.00.128.7r (H2) = 86 mmol·g−1·h−1
r (CO) = 203 mmol·g−1·h−1
H2/CO = 0.42
[105]
mes-TiO25003.891.80.2510.7N.A.[104]
P255003.862.00.127.5N.A.[104]
Pt/P255003.854.30.118.4r (H2) = 40 mmol·g−1·h−1
r (CO) = 106 mmol·g−1·h−1
H2/CO = 0.38
[104]
Ni/mes-TiO25003.885.10.2411.1r (H2) = 31 mmol·g−1·h−1
r (CO) = 81 mmol·g−1·h−1
H2/CO = 0.38
[104]
Pt/mes-TiO25003.877.50.2211.4r (H2) = 211 mmol·g−1·h−1
r (CO) = 309 mmol·g−1·h−1
H2/CO = 0.68
[104]
0.3 Ni6500580.70.392.2CH4 = 16%
CO2 = 25%
H2/CO = 0.69
[64]
0.2 PdNi6500523.60.382.4CH4 = 23%
CO2 = 31%
H2/CO = 0.83
[64]
0.3 PdNi6500514.70.352.2CH4 = 36%
CO2 = 52%
H2/CO = 0.81
[64]
0.4 PdNi6500506.60.352.1CH4 = 30%
CO2 = 42%
H2/CO = 0.81
[64]
Ni/CaO-Al2O3700076.10.2111.1CH4 conv. = 54%
CO2 conv. = 59%
[81]
CeO2-Ni/CaO-Al2O3700090.70.2913.2CH4 conv. = 82%
CO2 conv. = 84%
[81]
High surface area leads to a higher contact area for the reactants, leading to a higher activity. These last two examples underline the importance of the interaction between support and metal, as mentioned before, and it is evident in the two cases of Ni/CaO-Al2O3 and CeO2-Ni/CaO-Al2O3 that high surface areas resulted in high dispersion of Ni on the support, thus increasing the catalytic activity [81].
Moreover, as occurs in the case of PGM-free catalysts, an appropriate choice of support, doping and active metal is needed. On the one hand, in fact, doping of the support can induce localized states in the band gap, which may increase the absorption ability of the photocatalyst, but on the other hand, localized band states may increase the charge carrier recombination, decreasing the catalytic activity.
Table 5 reports the catalytic performances under different conditions for PGM-based catalysts.
In Table 6, the performance of catalysts for the photo-thermal dry reforming of methane reaction using different values of flow rates of feed gases is presented.

7. Conclusions

This paper presents an overview of the current trends in PGM-based and PGM-free catalysts for photo-thermal dry reforming of methane, also highlighting drawbacks for large-scale PTC-DRM applications. With appropriate catalyst design, dry reforming of methane at high temperatures produces a syngas with a high conversion efficiency, generally higher than that obtained in solar-driven DRM reactions. However, if on the one hand, DRM at high temperature does not meet the needs of further industrial applications due to the high costs of materials for the catalysts and reactor, then on the other hand, the limited utilization of visible light limits the scalability of PTC-DRM at the industrial level. The synergistic effect of heat and light irradiation can therefore be a valuable solution to lower the temperature of the DRM process and increase process yields, making PTC-DRM a possible competitive technology for converting greenhouse gases into fuels at a large scale. The yields of the DRM products may, in fact, be enhanced by the photoelectric effect, which leads to a decrease in activation energies under lower-temperature conditions. Supported precious metal are widely used in DRM; however, several non- precious metals show great potential for reducing DRM costs. A suitable combination of support and catalytic component is a key factor for enhancing the conversion efficiency of the PTC-DRM process; however, issues related to catalyst deactivation due to carbon deposition still need to be completely solved. Coke management is at present one of the main issues limiting the widespread application of DRM technology.

Author Contributions

A.V. and M.L.G. conceived the work. Data curation: A.V., U.P.L., M.F. and M.L.G. Original manuscript writing: A.V. and M.L.G. Review and editing: A.V., U.P.L., M.F. and M.L.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The following projects: (i) “Chemistry of Platinum Group Metals (CHemPGM)”, H2020-MSCA-RISE-2020 101007669, (ii) Electric System Research Programme, Project 1.6 “Energy Efficiency of Industrial Products and Processes”, Work Package 4 “Production of green H2 from biomass gasification using efficient CO2 capture, storage and reuse processes”, financed by the Ministry for the Environment and Energy Security, and (iii) “Development of ECCSEL—R.I. ItaLian facilities: usEr access, services and loNg-Term sustainability”, NATIONAL RECOVERY AND RESILIENCE PLAN (PNRR) MISSION 4 “Education and research”, COMPONENT 2 “From research to business”, INVESTMENT 3.1.1 “Fund for the creation of an integrated system of research and innovation infrastructures” (Infrastructures of research), are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shiraiwa, M. Facing Global Climate and Environmental Change. ACS Environ. Au 2023, 3, 121–122. [Google Scholar] [CrossRef] [PubMed]
  2. Yoro, K.O.; Daramola, M.O. CO2 Emission Sources, Greenhouse Gases, and the Global Warming Effect. In Advances in Carbon Capture: Methods, Technologies and Applications; Elsevier: Amsterdam, The Netherlands, 2020; pp. 3–28. ISBN 9780128196571. [Google Scholar]
  3. Global Climate Change Impacts in the United States: Highlights—Citations, Rights, Re-Use—UNT Digital Library. Available online: https://digital.library.unt.edu/ark:/67531/metadc11959/citation/#responsibilities-of-use (accessed on 4 June 2024).
  4. McCulloch, M.T.; Winter, A.; Sherman, C.E.; Trotter, J.A. 300 Years of Sclerosponge Thermometry Shows Global Warming Has Exceeded 1.5 °C. Nat. Clim. Change 2024, 14, 171–177. [Google Scholar] [CrossRef]
  5. Pearce, J.M.; Parncutt, R. Quantifying Global Greenhouse Gas Emissions in Human Deaths to Guide Energy Policy. Energies 2023, 16, 6074. [Google Scholar] [CrossRef]
  6. Painter, D.S. Burning Up: A Global History of Fossil Fuel Consumption. J. Interdiscip. Hist. 2019, 50, 442–443. [Google Scholar] [CrossRef]
  7. Hulme, M. 1.5 °C and climate research after the Paris Agreement. Nature Clim Change 2016, 6, 222–224. [Google Scholar] [CrossRef]
  8. Parsapur, R.K.; Chatterjee, S.; Huang, K.W. The Insignificant Role of Dry Reforming of Methane in CO2 emission Relief. ACS Energy Lett. 2020, 5, 2881–2885. [Google Scholar] [CrossRef]
  9. Li, M.; Sun, Z.; Hu, Y.H. Thermo-Photo Coupled Catalytic CO2 Reforming of Methane: A Review. Chem. Eng. J. 2022, 428, 131222. [Google Scholar] [CrossRef]
  10. Ji, G.; Wu, S.; Song, X.; Meng, L.; Jia, Y.; Tian, J. Recent Progress in Photo-Thermal Synergistic Catalysis for Methane Dry Reforming. Int. J. Hydrogen Energy 2024, 57, 696–708. [Google Scholar] [CrossRef]
  11. Lei, Y.; Ye, J.; García-Antón, J.; Liu, H. Recent Advances in the Built-in Electric-Field-Assisted Photocatalytic Dry Reforming of Methane. Chin. J. Catal. 2023, 53, 72–101. [Google Scholar] [CrossRef]
  12. Kotkowski, T.; Cherbański, R.; Stankiewicz, A.I. Electrifying the Dry Reforming of Methane. Shall We Target the Chemistry or the Heat Supply? Chem. Eng. Process. 2024, 202, 109875. [Google Scholar] [CrossRef]
  13. Cho, Y.; Shoji, S.; Yamaguchi, A.; Hoshina, T.; Fujita, T.; Abe, H.; Miyauchi, M. Visible-Light-Driven Dry Reforming of Methane Using a Semiconductor-Supported Catalyst. Chem. Commun. 2020, 56, 4611–4614. [Google Scholar] [CrossRef] [PubMed]
  14. Jabbour, K. Tuning combined steam and dry reforming of methane for “metgas” production: A thermodynamic approach and state-of-the-art catalysts. J. Energy 2020, 48, 54–91. [Google Scholar] [CrossRef]
  15. He, Z.; Huang, M.; Lin, T.; Zhong, L. Recent Advances in Dry Reforming of Methane via Photothermocatalysis. Acta Phys. Chim. Sin. 2023, 39, 2212060. [Google Scholar] [CrossRef]
  16. He, C.; Wu, S.; Wang, L.; Zhang, J. Recent Advances in Photo-Enhanced Dry Reforming of Methane: A Review. J. Photochem. Photobiol. C 2022, 51, 100468. [Google Scholar] [CrossRef]
  17. Aramouni, N.A.K.; Touma, J.G.; Tarboush, B.A.; Zeaiter, J.; Ahmad, M.N. Catalyst Design for Dry Reforming of Methane: Analysis Review. Renew. Sustain. Energy Rev. 2018, 82, 2570–2585. [Google Scholar] [CrossRef]
  18. Luyben, W.L. Design and Control of the Dry Methane Reforming Process. Ind. Eng. Chem. Res. 2014, 53, 14423–14439. [Google Scholar] [CrossRef]
  19. 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]
  20. Wang, S.; Lu, G.Q.; Millar, G.J. Carbon Dioxide Reforming of Methane to Produce Synthesis Gas over Metal-Supported Catalysts: State of the Art. Energy Fuels 1996, 10, 896–904. [Google Scholar] [CrossRef]
  21. Pakhare, D.; Spivey, J. A Review of Dry (CO2) Reforming of Methane over Noble Metal Catalysts. Chem. Soc. Rev. 2014, 43, 7813–7837. [Google Scholar] [CrossRef] [PubMed]
  22. Standard Thermodynamic Quantities for Chemical Substances at 25 °C. In CRC Handbook of Chemistry and Physics, 84th ed.; CRC Press: Boca Raton, FL, USA, 2004.
  23. Kulandaivalu, T.; Mohamed, A.R.; Ali, K.A.; Mohammadi, M. Photocatalytic Carbon Dioxide Reforming of Methane as an Alternative Approach for Solar Fuel Production—A Review. Renew. Sustain. Energy Rev. 2020, 134, 110363. [Google Scholar] [CrossRef]
  24. Bradford, M.C.J.; Vannice, M.A. CO2 Reforming of CH4. Catal. Rev. Sci. Eng. 1999, 41, 1–42. [Google Scholar] [CrossRef]
  25. Teh, L.P.; Setiabudi, H.D.; Timmiati, S.N.; Aziz, M.A.A.; Annuar, N.H.R.; Ruslan, N.N. Recent Progress in Ceria-Based Catalysts for the Dry Reforming of Methane: A Review. Chem. Eng. Sci. 2021, 242, 116606. [Google Scholar] [CrossRef]
  26. 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]
  27. Liu, X.; Shi, H.; Meng, X.; Sun, C.; Zhang, K.; Gao, L.; Ma, Y.; Mu, Z.; Ling, Y.; Cheng, B.; et al. Solar-Enhanced CO2 Conversion with CH4 over Synergetic NiCo Alloy Catalysts with Light-to-Fuel Efficiency of 33.8%. Sol. RRL 2021, 5, 2100185. [Google Scholar] [CrossRef]
  28. Xie, T.; Zhang, Z.Y.; Zheng, H.Y.; Xu, K.D.; Hu, Z.; Lei, Y. Enhanced Photothermal Catalytic Performance of Dry Reforming of Methane over Ni/Mesoporous TiO2 Composite Catalyst. Chem. Eng. J. 2022, 429, 132507. [Google Scholar] [CrossRef]
  29. Fang, S.; Hu, Y.H. Thermo-Photo Catalysis: A Whole Greater than the Sum of Its Parts. Chem. Soc. Rev. 2022, 51, 3609–3647. [Google Scholar] [CrossRef] [PubMed]
  30. De Caprariis, B.; De Filippis, P.; Palma, V.; Petrullo, A.; Ricca, A.; Ruocco, C.; Scarsella, M. Rh, Ru and Pt Ternary Perovskites Type Oxides BaZr(1-x)MexO3 for Methane Dry Reforming. Appl. Catal. A Gen. 2016, 517, 47–55. [Google Scholar] [CrossRef]
  31. Rego de Vasconcelos, B.; Pham Minh, D.; Martins, E.; Germeau, A.; Sharrock, P.; Nzihou, A. Highly-Efficient Hydroxyapatite-Supported Nickel Catalysts for Dry Reforming of Methane. Int. J. Hydrogen Energy 2020, 45, 18502–18518. [Google Scholar] [CrossRef]
  32. Tang, Y.; Li, Y.; Bao, W.; Yan, W.; Zhang, J.; Huang, Y.; Li, H.; Wang, Z.; Liu, M.; Yu, F. Enhanced Dry Reforming of CO2 and CH4 on Photothermal Catalyst Ru/SrTiO3. Appl. Catal. B 2023, 338, 123054. [Google Scholar] [CrossRef]
  33. Sokefun, Y.O.; Trottier, J.; Yung, M.M.; Joseph, B.; Kuhn, J.N. Low Temperature Dry Reforming of Methane Using Ru-Ni-Mg/Ceria-Zirconia Catalysts: Effect of Ru Loading and Reduction Temperature. Appl. Catal. A Gen. 2022, 645, 118842. [Google Scholar] [CrossRef]
  34. Fresno, F.; Iglesias-Juez, A.; Coronado, J.M. Photothermal Catalytic CO2 Conversion: Beyond Catalysis and Photocatalysis. Top. Curr. Chem. 2023, 381, 21. [Google Scholar] [CrossRef] [PubMed]
  35. Ghoussoub, M.; Xia, M.; Duchesne, N.P.; Segal, D.; Ozin, G. Principles of photothermal gas-phase heterogenous CO2 catalysis. Energy Environ. Sci. 2019, 12, 1122–1142. [Google Scholar] [CrossRef]
  36. Entesari, N.; Goeppert, A.; Prakash, G.K.S. Renewable Methanol Synthesis through Single Step Bi-Reforming of Biogas. Ind. Eng. Chem. Res. 2020, 59, 10542–10551. [Google Scholar] [CrossRef]
  37. Farooqi, A.S.; Yusuf, M.; Mohd Zabidi, N.A.; Saidur, R.; Sanaullah, K.; Farooqi, A.S.; Khan, A.; Abdullah, B. A Comprehensive Review on Improving the Production of Rich-Hydrogen via Combined Steam and CO2 Reforming of Methane over Ni-Based Catalysts. Int. J. Hydrogen Energy 2021, 46, 31024–31040. [Google Scholar] [CrossRef]
  38. Tahir, M.; Ali Khan, A.; Bafaqeer, A.; Kumar, N.; Siraj, M.; Fatehmulla, A. Highly Stable Photocatalytic Dry and Bi-Reforming of Methane with the Role of a Hole Scavenger for Syngas Production over a Defective Co-Doped g-C3N4 Nanotexture. Catalysts 2023, 13, 1140. [Google Scholar] [CrossRef]
  39. Lino, A.V.P.; Colmenares Calderon, Y.N.; Mastelaro, V.R.; Assaf, E.M.; Assaf, J.M. Syngas for Fischer-Tropsch Syn-thesis by Methane Tri-Reforming Using Nickel Supported on MgAl2O4 Promoted with Zr, Ce and Ce-Zr. Appl. Surf. Sci. 2019, 481, 747–760. [Google Scholar] [CrossRef]
  40. Minh, D.P.; Siang, T.J.; Vo, D.V.N.; Phan, T.S.; Ridart, C.; Nzihou, A.; Grouset, D. Hydrogen Production from Biogas Reforming: An Overview of Steam Reforming, Dry Reforming, Dual Reforming, and Tri-Reforming of Methane. Hydrog. Supply Chain Des. Deploy. Oper. 2018, 111–166. [Google Scholar] [CrossRef]
  41. 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]
  42. Wang, Z.; Yang, Z.; Kadirova, Z.C.; Guo, M.; Fang, R.; He, J.; Yan, Y.; Ran, J. Photothermal Functional Material and Structure for Photothermal Catalytic CO2 Reduction: Recent Advance, Application and Prospect. Coord. Chem. Rev. 2022, 473, 214794. [Google Scholar] [CrossRef]
  43. Linic, S.; Aslam, U.; Boerigter, C.; Morabito, M. Photochemical Transformations on Plasmonic Metal Nanoparticles. Nat. Mater. 2015, 14, 567–576. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, J.; Chen, H.; Duan, X.; Sun, H.; Wang, S. Photothermal Catalysis: From Fundamentals to Practical Applications. Mater. Today 2023, 68, 234–253. [Google Scholar] [CrossRef]
  45. Mateo, D.; Cerrillo, J.L.; Durini, S.; Gascon, J. Fundamentals and Applications of Photo-Thermal Catalysis. Chem. Soc. Rev. 2021, 50, 2173–2210. [Google Scholar] [CrossRef]
  46. Zhang, Y.; He, S.; Guo, W.; Hu, Y.; Huang, J.; Mulcahy, J.R.; Wei, W.D. Surface-Plasmon-Driven Hot Electron Photochemistry. Chem. Rev. 2018, 118, 2927–2954. [Google Scholar] [CrossRef] [PubMed]
  47. Song, C.; Wang, Z.; Yin, Z.; Xiao, D.; Ma, D. Principles and Applications of Photothermal Catalysis. Chem. Catal. 2022, 2, 52–83. [Google Scholar] [CrossRef]
  48. Iglesias-Juez, A.; Fresno, F.; Coronado, J.M.; Highfield, J.; Ruppert, A.M.; Keller, N. Emerging High-Prospect Applications in Photothermal Catalysis. Curr. Opin. Green. Sustain. Chem. 2022, 37, 100652. [Google Scholar] [CrossRef]
  49. Zhang, Y.; Yan, L.; Guan, M.; Chen, D.; Xu, Z.; Guo, H.; Hu, S.; Zhang, S.; Liu, X.; Guo, Z.; et al. Indirect to Direct Charge Transfer Transition in Plasmon-Enabled CO2 Photoreduction. Adv. Sci. 2022, 9, 102978. [Google Scholar] [CrossRef]
  50. Zhu, Z.; Guo, W.; Zhang, Y.; Pan, C.; Xu, J.; Zhu, Y.; Lou, Y. Research Progress on Methane Conversion Coupling Photocatalysis and Thermocatalysis. Carbon. Energy 2021, 3, 519–540. [Google Scholar] [CrossRef]
  51. Zhang, J.; Li, Y.; Sun, J.; Chen, H.; Zhu, Y.; Zhao, X.; Zhang, L.C.; Wang, S.; Zhang, H.; Duan, X.; et al. Regulation of Energetic Hot Carriers on Pt/TiO2 with Thermal Energy for Photothermal Catalysis. Appl. Catal. B 2022, 309, 121263. [Google Scholar] [CrossRef]
  52. Kattel, S.; Liu, P.; Chen, J.G. Tuning Selectivity of CO2 Hydrogenation Reactions at the Metal/Oxide Interface. J. Am. Chem. Soc. 2017, 139, 9739–9754. [Google Scholar] [CrossRef]
  53. Wang, N.; Qian, W.; Chu, W.; Wei, F. Crystal-Plane Effect of Nanoscale CeO2 on the Catalytic Performance of Ni/CeO2 Catalysts for Methane Dry Reforming. Catal. Sci. Technol. 2016, 6, 3594–3605. [Google Scholar] [CrossRef]
  54. Tavasoli, A.V.; Preston, M.; Ozin, G. Photocatalytic Dry Reforming: What Is It Good For? Energy Environ. Sci. 2021, 14, 3098–3109. [Google Scholar] [CrossRef]
  55. Pan, F.; Xiang, X.; Du, Z.; Sarnello, E.; Li, T.; Li, Y. Integrating Photocatalysis and Thermocatalysis to Enable Efficient CO2 Reforming of Methane on Pt Supported CeO2 with Zn Doping and Atomic Layer Deposited MgO Overcoating. Appl. Catal. B 2020, 260, 118189. [Google Scholar] [CrossRef]
  56. Ranjekar, A.M.; Yadav, G.D. Dry Reforming of Methane for Syngas Production: A Review and Assessment of Catalyst Development and Efficacy. J. Indian Chem. Soc. 2021, 98, 100002. [Google Scholar] [CrossRef]
  57. Zhang, Z.; Zhang, Y.; Liu, L. Role and Mechanism of Calcium-Based Catalysts for Methane Dry Reforming: A Review. Fuel 2024, 355, 129329. [Google Scholar] [CrossRef]
  58. Zhang, Z.Y.; Li, T.; Yao, J.L.; Xie, T.; Xiao, Q. Mechanism and Kinetic Characteristics of Photo-Thermal Dry Reforming of Methane on Pt/Mesoporous-TiO2 Catalyst. Mol. Catal. 2023, 535, 112828. [Google Scholar] [CrossRef]
  59. Xu, Z.; Park, E.D. Recent Advances in Coke Management for Dry Reforming of Methane over Ni-Based Catalysts. Catalysts 2024, 14, 176. [Google Scholar] [CrossRef]
  60. Grams, J.; Ruppert, A.M. Catalyst Stability—Bottleneck of Efficient Catalytic Pyrolysis. Catalysts 2021, 11, 265. [Google Scholar] [CrossRef]
  61. Study on the Critical Raw Materials for the EU 2023—Publications Office of the EU. Available online: https://op.europa.eu/en/publication-detail/-/publication/57318397-fdd4-11ed-a05c-01aa75ed71a1/language-en (accessed on 3 June 2024).
  62. Anderson, C.G. The Production of Critical Materials as By Products. Asp. Min. Miner. Sci. 2018, 2, 532. [Google Scholar] [CrossRef]
  63. García-Diéguez, M.; Pieta, I.S.; Herrera, M.C.; Larrubia, M.A.; Alemany, L.J. Improved Pt-Ni Nanocatalysts for Dry Reforming of Methane. Appl. Catal. A Gen. 2010, 377, 191–199. [Google Scholar] [CrossRef]
  64. Damyanova, S.; Pawelec, B.; Arishtirova, K.; Fierro, J.L.G.; Sener, C.; Dogu, T. MCM-41 Supported PdNi Catalysts for Dry Reforming of Methane. Appl. Catal. B 2009, 92, 250–261. [Google Scholar] [CrossRef]
  65. Pinheiro, A.N.; Valentini, A.; Sasaki, J.M.; Oliveira, A.C. Highly Stable Dealuminated Zeolite Support for the Production of Hydrogen by Dry Reforming of Methane. Appl. Catal. A Gen. 2009, 355, 156–168. [Google Scholar] [CrossRef]
  66. Alabi, W.O. CO2 Reforming of CH4 on Ni-Al-Ox Catalyst Using Pure and Coal Gas Feeds: Synergetic Effect of CoO and MgO in Mitigating Carbon Deposition. Environ. Pollut. 2018, 242, 1566–1576. [Google Scholar] [CrossRef] [PubMed]
  67. Grilli, M.L.; Slobozeanu, A.E.; Larosa, C.; Paneva, D.; Yakoumis, I.; Cherkezova-Zheleva, Z. Platinum Group Metals: Green Recovery from Spent Auto-Catalysts and Reuse in New Catalysts—A Review. Crystals 2023, 13, 550. [Google Scholar] [CrossRef]
  68. Daily Metal Price: Cobalt Price (USD/Pound) Chart for the Last Year. Available online: https://www.dailymetalprice.com/metalpricecharts.php?c=co&u=lb&d=240 (accessed on 7 January 2024).
  69. Hu, Q.; Li, Y.; Wu, J.; Hu, Y.; Cao, H.; Yang, Y. Extraordinary Catalytic Performance of Nickel Half-Metal Clusters for Light-Driven Dry Reforming of Methane. Adv. Energy Mater. 2023, 13, 2300071. [Google Scholar] [CrossRef]
  70. Banković-Ilić, I.B.; Miladinović, M.R.; Stamenković, O.S.; Veljković, V.B. Application of Nano CaO–Based Catalysts in Biodiesel Synthesis. Renew. Sustain. Energy Rev. 2017, 72, 746–760. [Google Scholar] [CrossRef]
  71. Wang, L.; Pu, Z.; Shi, Y.; Wu, M.; Shi, W.; Wang, F. Photothermal Methane Dry Reforming Catalyzed by Multifunctional (Ni-Cu/CeO2)@SiO2 Catalyst. ACS Sustain. Chem. Eng. 2023, 11, 17384–17399. [Google Scholar] [CrossRef]
  72. Wang, E.; Zhu, Z.; Li, R.; Wu, J.; Ma, K.; Zhang, J. Ni/CaO-Based Dual-Functional Materials for Calcium-Looping CO2 Capture and Dry Reforming of Methane: Progress and Challenges. Chem. Eng. J. 2024, 482, 148476. [Google Scholar] [CrossRef]
  73. Han, K.; Wang, Y.; Wang, S.; Liu, Q.; Deng, Z.; Wang, F. Narrowing Band Gap Energy of CeO2 in (Ni/CeO2)@SiO2 Catalyst for Photothermal Methane Dry Reforming. Chem. Eng. J. 2021, 421, 129989. [Google Scholar] [CrossRef]
  74. Yao, Y.; Li, B.; Gao, X.; Yang, Y.; Yu, J.; Lei, J.; Li, Q.; Meng, X.; Chen, L.; Xu, D. Highly Efficient Solar-Driven Dry Reforming of Methane on a Rh/LaNiO3 Catalyst through a Light-Induced Metal-To-Metal Charge Transfer Process. Adv. Mater. 2023, 35, 2303654. [Google Scholar] [CrossRef] [PubMed]
  75. Lorber, K.; Djinović, P. Accelerating Photo-Thermal CO2 Reduction to CO, CH4 or Methanol over Metal/Oxide Semiconductor Catalysts. iScience 2022, 25, 104107. [Google Scholar] [CrossRef]
  76. Liu, H.; Li, M.; Dao, T.D.; Liu, Y.; Zhou, W.; Liu, L.; Meng, X.; Nagao, T.; Ye, J. Design of PdAu Alloy Plasmonic Nanoparticles for Improved Catalytic Performance in CO2 Reduction with Visible Light Irradiation. Nano Energy 2016, 26, 398–404. [Google Scholar] [CrossRef]
  77. 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]
  78. Moogi, S.; Hyun Ko, C.; Hoon Rhee, G.; Jeon, B.H.; Ali Khan, M.; Park, Y.K. Influence of Catalyst Synthesis Methods on Anti-Coking Strength of Perovskites Derived Catalysts in Biogas Dry Reforming for Syngas Production. Chem. Eng. J. 2022, 437, 135348. [Google Scholar] [CrossRef]
  79. Zhao, S.; Luo, Y.; Li, C.; Ren, K.; Zhu, Y.; Dou, W. High-Performance Photothermal Catalytic CO2 Reduction to CH4 and CO by ABO3 (A = La, Ce; B = Ni, Co, Fe) Perovskite Nanomaterials. Ceram. Int. 2023, 49, 20907–20919. [Google Scholar] [CrossRef]
  80. Costilla, I.O.; Sánchez, M.D.; Gigola, C.E. Palladium Nanoparticle’s Surface Structure and Morphology Effect on the Catalytic Activity for Dry Reforming of Methane. Appl. Catal. A Gen. 2014, 478, 38–44. [Google Scholar] [CrossRef]
  81. Taherian, Z.; Shahed Gharahshiran, V.; Khataee, A.; Meshkani, F.; Orooji, Y. Comparative Study of Modified Ni Catalysts over Mesoporous CaO-Al2O3 Support for CO2/Methane Reforming. Catal. Commun. 2020, 145, 106100. [Google Scholar] [CrossRef]
  82. Takami, D.; Ito, Y.; Kawaharasaki, S.; Yamamoto, A.; Yoshida, H. Low Temperature Dry Reforming of Methane over Plasmonic Ni Photocatalysts under Visible Light Irradiation. Sustain. Energy Fuels 2019, 3, 2968–2971. [Google Scholar] [CrossRef]
  83. Fertout, R.I.; Ghelamallah, M.; Helamallah, M.; Kacimi, S.; López, P.N.; Corberán, V.C. Nickel Supported on Alkaline Earth Metal–Doped γ-Al2O3–La2O3 as Catalysts for Dry Reforming of Methane. Russ. J. Appl. Chem. 2020, 93, 289–298. [Google Scholar] [CrossRef]
  84. Khan, A.A.; Tahir, M. Well-Designed 2D/2D Ti3C2TA/R MXene Coupled g-C3N4 Heterojunction with in-Situ Growth of Anatase/Rutile TiO2 Nucleates to Boost Photocatalytic Dry-Reforming of Methane (DRM) for Syngas Production under Visible Light. Appl. Catal. B 2021, 285, 119777. [Google Scholar] [CrossRef]
  85. Tahir, B.; Tahir, M.; Amin, N.A.S. Photo-Induced CO2 Reduction by CH4/H2O to Fuels over Cu-Modified g-C3N4 Nanorods under Simulated Solar Energy. Appl. Surf. Sci. 2017, 419, 875–885. [Google Scholar] [CrossRef]
  86. Taherian, Z.; Gharahshiran, V.S.; Fazlikhani, F.; Yousefpour, M. Catalytic Performance of Samarium-Modified Ni Catalysts over Al2O3–CaO Support for Dry Reforming of Methane. Int. J. Hydrogen Energy 2021, 46, 7254–7262. [Google Scholar] [CrossRef]
  87. Ranjbar, A.; Rezaei, M. Preparation of Nickel Catalysts Supported on CaO.2Al2O3 for Methane Reforming with Carbon Dioxide. Int. J. Hydrogen Energy 2012, 37, 6356–6362. [Google Scholar] [CrossRef]
  88. Wang, C.; Sun, N.; Zhao, N.; Wei, W.; Sun, Y.; Sun, C.; Liu, H.; Snape, C.E. Coking and Deactivation of a Mesoporous Ni-CaO-ZrO2 Catalyst in Dry Reforming of Methane: A Study under Different Feeding Compositions. Fuel 2015, 143, 527–535. [Google Scholar] [CrossRef]
  89. Dama, S.; Ghodke, S.R.; Bobade, R.; Gurav, H.R.; Chilukuri, S. Active and Durable Alkaline Earth Metal Substituted Perovskite Catalysts for Dry Reforming of Methane. Appl. Catal. B 2018, 224, 146–158. [Google Scholar] [CrossRef]
  90. Zhao, J.; Guo, X.; Shi, R.; Waterhouse, G.I.N.; Zhang, X.; Dai, Q.; Zhang, T. NiFe Nanoalloys Derived from Layered Double Hydroxides for Photothermal Synergistic Reforming of CH4 with CO2. Adv. Funct. Mater. 2022, 32, 2204056. [Google Scholar] [CrossRef]
  91. Tahir, B.; Tahir, M.; Amin, N.A.S. Tailoring Performance of La-Modified TiO2 Nanocatalyst for Continuous Photocatalytic CO2 Reforming of CH4 to Fuels in the Presence of H2O. Energy Convers. Manag. 2018, 159, 284–298. [Google Scholar] [CrossRef]
  92. 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/Ga2O3. ACS Catal. 2021, 11, 4730–4738. [Google Scholar] [CrossRef]
  93. Boudart, M. Turnover Rates in Heterogeneous Catalysis. Chem. Rev. 1995, 95, 661–666. [Google Scholar] [CrossRef]
  94. Kozuch, S.; Martin, J.M.L. “Turning over” Definitions in Catalytic Cycles. ACS Catal. 2012, 2, 2787–2794. [Google Scholar] [CrossRef]
  95. Crampton, A.S.; Rötzer, M.D.; Ridge, C.J.; Yoon, B.; Schweinberger, F.F.; Landman, U.; Heiz, U. Assessing the Concept of Structure Sensitivity or Insensitivity for Sub-Nanometer Catalyst Materials. Surf. Sci. 2016, 652, 7–19. [Google Scholar] [CrossRef]
  96. Vogt, C.; Kranenborg, J.; Monai, M.; Weckhuysen, B.M. Structure Sensitivity in Steam and Dry Methane Reforming over Nickel: Activity and Carbon Formation. ACS Catal. 2020, 10, 1428–1438. [Google Scholar] [CrossRef]
  97. 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 2017, 203, 625–632. [Google Scholar] [CrossRef]
  98. Wang, Y.; Yao, L.; Wang, Y.; Wang, S.; Zhao, Q.; Mao, D.; Hu, C. Low-Temperature Catalytic CO2 Dry Reforming of Methane on Ni-Si/ZrO2 Catalyst. ACS Catal. 2018, 8, 6495–6506. [Google Scholar] [CrossRef]
  99. Usman, M.; Wan Daud, W.M.A.; Abbas, H.F. Dry Reforming of Methane: Influence of Process Parameters—A Review. Renew. Sustain. Energy Rev. 2015, 45, 710–744. [Google Scholar] [CrossRef]
  100. Yang, Y.; Chai, Z.; Qin, X.; Zhang, Z.; Muhetaer, A.; Wang, C.; Huang, H.; Yang, C.; Ma, D.; Li, Q.; et al. Light-Induced Redox Looping of a Rhodium/Cex WO3 Photocatalyst for Highly Active and Robust Dry Reforming of Methane. Angew. Chem. Int. Ed. Engl. 2022, 61, e202200567. [Google Scholar] [CrossRef] [PubMed]
  101. Yin, Q.; Shen, T.; Li, J.; Ning, C.; Xue, Y.; Chen, G.; Xu, M.; Wang, F.; Song, Y.F.; Zhao, Y.; et al. Solar-Driven Dry Reforming of Methane Using RuNi Single-Atom Alloy Catalyst Coupled with Thermal Decomposition of Carbonates. Chem. Eng. J. 2023, 470, 144416. [Google Scholar] [CrossRef]
  102. Song, H.; Meng, X.; Dao, T.D.; Zhou, W.; Liu, H.; Shi, L.; Zhang, H.; Nagao, T.; Kako, T.; Ye, J. Light-Enhanced Carbon Dioxide Activation and Conversion by Effective Plasmonic Coupling Effect of Pt and Au Nanoparticles. ACS Appl. Mater. Interfaces 2018, 10, 408–416. [Google Scholar] [CrossRef] [PubMed]
  103. Wu, S.; Li, Y.; Hu, Q.; Wu, J.; Zhang, Q. Photothermocatalytic Dry Reforming of Methane for Efficient CO2 Reduction and Solar Energy Storage. ACS Sustain. Chem. Eng. 2021, 9, 11635–11651. [Google Scholar] [CrossRef]
  104. Zhang, Z.Y.; Zhang, T.; Liang, W.P.; Bai, P.W.; Zheng, H.Y.; Lei, Y.; Hu, Z.; Xie, T. Promoted Solar-Driven Dry Reforming of Methane with Pt/Mesoporous-TiO2 Photo-Thermal Synergistic Catalyst: Performance and Mechanism Study. Energy Convers. Manag. 2022, 258, 115496. [Google Scholar] [CrossRef]
  105. Zhang, Z.-Y.; Zhang, T.; Wang, R.-K.; Yu, B.; Tang, Z.-Y.; Zheng, H.-Y.; He, D.; Xie, T.; Hu, Z. Photo-Enhanced Dry Reforming of Methane over Pt-Au/P25 Composite Catalyst by Coupling Plasmonic Effect. J. Catal. 2022, 413, 829–842. [Google Scholar] [CrossRef]
  106. Zhang, J.; Wang, L.; Zhao, X.; Shi, L.; Chen, H.; Zhang, S.; Zhang, P.; Wang, S.; Zhang, L.; Wang, Y.; et al. The Nature of Active Sites for Plasmon-Mediated Photothermal Catalysis and Heat-Coupled Photocatalysis in Dry Reforming of Methane. Energy Environ. Mater. 2023, 6, e12416. [Google Scholar] [CrossRef]
  107. Zhao, Y.X.; Yang, B.; Li, H.F.; Zhang, Y.; Yang, Y.; Liu, Q.Y.; Xu, H.G.; Zheng, W.J.; He, S.G. Photoassisted Selective Steam and Dry Reforming of Methane to Syngas Catalyzed by Rhodium–Vanadium Bimetallic Oxide Cluster Anions at Room Temperature. Angew. Chem. Int. Ed. 2020, 59, 21216–21223. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The link between increasing global temperature and increasing CO2 emissions from 1859 to 2022. The data are taken from ref. [5].
Figure 1. The link between increasing global temperature and increasing CO2 emissions from 1859 to 2022. The data are taken from ref. [5].
Materials 17 03809 g001
Figure 2. DRM reaction promoted by a source of heat.
Figure 2. DRM reaction promoted by a source of heat.
Materials 17 03809 g002
Figure 4. Publications trends over the years of “DRM”, “PTC-DRM”, “BRM” and “TDRM” (source: Scopus: 5 July 2024).
Figure 4. Publications trends over the years of “DRM”, “PTC-DRM”, “BRM” and “TDRM” (source: Scopus: 5 July 2024).
Materials 17 03809 g004
Figure 5. Time scales of plasmon-induced hot carrier generation, hot electron transfer, and thermalization processes. Figure reprinted with permission from [46].
Figure 5. Time scales of plasmon-induced hot carrier generation, hot electron transfer, and thermalization processes. Figure reprinted with permission from [46].
Materials 17 03809 g005
Figure 7. Simplified mechanism of PTC-DRM. (a) Activation of methane; (b) activation of carbon dioxide. The symbol * indicates the species adsorbed on the surface of the catalyst. Ov stands for oxygen vacancy.
Figure 7. Simplified mechanism of PTC-DRM. (a) Activation of methane; (b) activation of carbon dioxide. The symbol * indicates the species adsorbed on the surface of the catalyst. Ov stands for oxygen vacancy.
Materials 17 03809 g007
Figure 8. Exemplified scheme of a fixed bed reactor for PTC-DRM.
Figure 8. Exemplified scheme of a fixed bed reactor for PTC-DRM.
Materials 17 03809 g008
Figure 9. Formation of a coke deposit on a surface of a catalyst. Reprinted from [60]. Copyright MDPI 2021.
Figure 9. Formation of a coke deposit on a surface of a catalyst. Reprinted from [60]. Copyright MDPI 2021.
Materials 17 03809 g009
Table 1. Gas hourly space velocity of state-of-the-art DRM catalysts at different temperatures and feed ratio CH4/CO2 equal to 1/1.
Table 1. Gas hourly space velocity of state-of-the-art DRM catalysts at different temperatures and feed ratio CH4/CO2 equal to 1/1.
CatalystsGHSV
h−1
τ
s
Temperature
°C
Catalytic PerformancesReferences
CeO2-Ni/CaO-Al2O390000.4700CH4 conv. = 81%
CO2 conv. = 84%
[81]
Sm/Ni/Al2O3-CaO12,0000.3700CH4 conv. = 68%
CO2 conv. = 43%
H2/CO = 0.64
[86]
7%Ni/CaO-Al2O318000.5700CH4 conv. = 66%
CO2 conv. =62%
H2/CO = 0.85
[87]
Ni-CaO-ZrO248,0000.075750CH4 conv. = 82%
H2/CO = 0.94
[88]
CaZr0.8Ni0.2O3-δ28,8000.125800CH4 conv. = 95%
CO2 conv. = 96%
H2/CO = 0.98
[89]
Table 2. Catalytic performances of PGM-free catalysts. Production rates have been calculated using (16). For solar irradiation, a Xe lamp was used. Feed ratio of CH4/CO2 is 1/1.
Table 2. Catalytic performances of PGM-free catalysts. Production rates have been calculated using (16). For solar irradiation, a Xe lamp was used. Feed ratio of CH4/CO2 is 1/1.
CatalystSurface Temperature
°C
Light Intensity
mW·cm−2
Catalytic PerformancesReference
CN/TNT2520r (H2) = 49 μmol·g−1·h−1
r (CO) = 75 μmol·g−1·h−1
H2/CO = 0.64
[84]
(Ni/CeO2)-SiO2750-CH4 conv. = 66%
CO2 conv. = 80%
H2/CO = 0.90
[73]
Ni3Fe1 nanoalloy3503.62r (H2) = 326 × 103 μmol·g−1·h−1
r (CO) = 632 × 103 μmol·g−1·h−1
H2/CO = 0.52
[90]
La/TiO2100150r(H2) = 74 μmol·h−1·gcat−1
r (CO) = 183 μmol·h−1·gcat−1
H2/CO = 0.40
[91]
Cu-g-C3N4100100r(H2) = 76 μmol·h−1·gcat−1
r(CO) = 142 μmol·h−1·gcat−1
H2/CO = 0.54
[85]
Ni/Ga2O33911.9–3r (H2) = 194 μmol·g−1·h−1
r (CO) = 206 μmol·g−1·h−1
H2/CO = 0.94
[92]
Table 3. Characteristics of Ni-based catalysts with different supports and coatings: turnover frequencies and conversion of methane and carbon dioxide. Feed ratio of CH4/CO2 is 1/1.
Table 3. Characteristics of Ni-based catalysts with different supports and coatings: turnover frequencies and conversion of methane and carbon dioxide. Feed ratio of CH4/CO2 is 1/1.
SamplesTemperature
°C
T.O.F. (CH4)
s−1
T.O.F. (CO2)
s−1
CH4conv
%
CO2conv
%
References
Ni/SiO2@Al2O3800135.2177.262.882.3[97]
Ni/SiO2@MgO800120.0171.150.071.3[97]
Ni/SiO2@ZrO280029.537.829.137.3[97]
Ni/SiO2@TiO280012.918.117.023.8[97]
Ni-Zr/SiO24000.320.3222[98]
Ni-Zr/SiO24501.061.480.81.2[98]
Ni-Si/ZrO24000.500.444.33.8[98]
Ni-Si/ZrO24501.381.301.62.4[98]
Table 5. Catalytic performances of PGM-based catalysts for photo-thermal dry reforming of methane under different working conditions. Xe lamp was used, except for [55], in which concentrated sunlight was used. Feed ratio of CH4/CO2 is 1/1.
Table 5. Catalytic performances of PGM-based catalysts for photo-thermal dry reforming of methane under different working conditions. Xe lamp was used, except for [55], in which concentrated sunlight was used. Feed ratio of CH4/CO2 is 1/1.
CatalystSurface Temperature °CLight Intensity
W·cm−2
Catalytic PerformancesReference
Rh/La2O33401.5r (H2) = 452 mmol·g−1·h−1
r (CO) = 527 mmol·g−1·h−1
H2/CO = 0.86
[74]
Pt/TiO25004.67r (H2) = 134 mmol·g−1·h−1
r (CO) = 221 mmol·g−1·h−1
H2/CO = 0.61
[58]
Pt/TiO23504.67r (H2) = 7 mmol·g−1·h−1
r (CO) = 22 mmol·g−1·h−1
H2/CO = 0.32
[58]
Pt/TiO25003.72r (H2) = 121 mmol·g−1·h−1
r (CO) = 200 mmol·g−1·h−1
H2/CO = 0.61
[58]
MgO/Pt/Zn-CeO26003r (H2) = 356 mmol·g−1·h−1
r (CO) = 516 mmol·g−1·h−1
H2/CO = 0.69
[55]
Ru-Al/LDH35013.5r (H2) = 9 mmol·g−1·h−1
r (CO) = 11 mmol·g−1·h−1
H2/CO = 0.76
[101]
Pt/TiO25000.4r (H2) = 598 mmol·g−1·h−1
r (CO) = 902 mmol·g−1·h−1
H2/CO = 0.66
[51]
Pt/TiO27000.4r (H2) = 1103 mmol·g−1·h−1
r (CO) = 1495 mmol·g−1·h−1
H2/CO = 0.74
[51]
Pt/TiO25003.76r(H2) = 211 mmol·g−1·h−1
r(CO) = 309 mmol·g−1·h−1
H2/CO = 0.68
[104]
Pt/TiO25002.2r (H2) = 103 mmol·g−1·h−1
r (CO) = 179 mmol·g−1·h−1
H2/CO = 0.58
[104]
Table 6. Performance comparison of catalysts for photo-thermal dry reforming of methane using different values of flow rates (GHSV). – the value of light intensity is not available.
Table 6. Performance comparison of catalysts for photo-thermal dry reforming of methane using different values of flow rates (GHSV). – the value of light intensity is not available.
CatalystTemperature
°C
Light Intensity
mW·cm−2
GHSV
h−1
Catalytic PerformancesReferences
Ru/SrTiO36004592r(H2) = 320 mmol·g−1·h−1
r(CO) = 387 mmol·g−1·h−1
H2/CO = 0.83
[32]
Pt/TiO25003.2240,000r(H2) = 211 mmol·g−1·h−1
r(CO) = 309 mmol·g−1·h−1
H2/CO = 0.68
[104]
Pt/P255003.2240,000r(H2) = 40 mmol·g−1·h−1
r(CO) = 106 mmol·g−1·h−1
H2/CO = 0.38
[104]
Pt/TiO25004.67320,000r (H2) = 55 mmol·g−1·h−1
r (CO) = 73 mmol·g−1·h−1
H2/CO = 0.75
[58]
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

Varotto, A.; Pasqual Laverdura, U.; Feroci, M.; Grilli, M.L. Photo-Thermal Dry Reforming of Methane with PGM-Free and PGM-Based Catalysts: A Review. Materials 2024, 17, 3809. https://doi.org/10.3390/ma17153809

AMA Style

Varotto A, Pasqual Laverdura U, Feroci M, Grilli ML. Photo-Thermal Dry Reforming of Methane with PGM-Free and PGM-Based Catalysts: A Review. Materials. 2024; 17(15):3809. https://doi.org/10.3390/ma17153809

Chicago/Turabian Style

Varotto, Alessio, Umberto Pasqual Laverdura, Marta Feroci, and Maria Luisa Grilli. 2024. "Photo-Thermal Dry Reforming of Methane with PGM-Free and PGM-Based Catalysts: A Review" Materials 17, no. 15: 3809. https://doi.org/10.3390/ma17153809

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

Article Metrics

Back to TopTop