Next Article in Journal
Continuous-Thrust Circular Orbit Phasing Optimization of Deep Space CubeSats
Previous Article in Journal
Assessment of Feldkamp-Davis-Kress Reconstruction Parameters in Overall Image Quality in Cone Beam Computed Tomography
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Direct Solar Thermal Water-Splitting Using Iron and Iron Oxides at High Temperatures: A Review

1
Centro de Desarrollo Energético Antofagasta, Universidad de Antofagasta, Antofagasta 1240000, Chile
2
Departamento de Ingeniería Mecánica, Universidad de Tarapacá, Arica 1100000, Chile
3
Departamento de Ingeniería en Metalurgia, Universidad de Atacama, Copiapó 1531772, Chile
4
Facultad de Ingeniería y Arquitectura, Universidad Arturo Prat, Iquique 1100000, Chile
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2024, 14(16), 7056; https://doi.org/10.3390/app14167056 (registering DOI)
Submission received: 11 July 2024 / Revised: 2 August 2024 / Accepted: 6 August 2024 / Published: 12 August 2024

Abstract

:
Green hydrogen is poised to play a crucial role in the energy-transition process in developed countries over the coming years, particularly in those countries aiming to achieve net-zero emissions. Consequently, the for green hydrogen is expected to rise significantly. This article explores the fundamental methods of producing hydrogen, focusing on the oxidation reaction within a thermochemical solar cycle for the dissociation of steam. Solar thermochemical cycles have been extensively researched, yet they remain in the development stage as research groups strive to identify optimal materials and conditions to enhance process efficiency, especially at high temperatures. The article analyses theoretical foundations drawn from exhaustive scientific studies related to the oxidation of iron in steam, the relationship with the activation energy of the corrosive process, thermodynamic aspects, and the kinetic model of a heterogeneous reaction. Additionally, it presents various mechanisms of high-temperature oxidation, pH effects, reactors, and materials (including fluidized beds). This scientific review suggests that hydrogen production via a thermochemical cycle is more efficient than production via electrochemical processes (such as electrolysis), provided the limitations of the cycle’s reduction stage can be overcome.

1. Introduction

The highest levels of solar radiation are typically found in arid and hyperarid climates. The Atacama Desert in northern Chile offers ideal solar radiation, making it widely recognised for its potential usefulness in producing solar hydrogen (H2). Additionally, the region boasts a stable and predictable climate, with minimal rainfall and a lack of extreme temperature variations. This climatic consistency is crucial for the long-term performance and reliability of solar energy systems. Moreover, the desert’s vast expanses of land provide ample space for large-scale solar installations, facilitating the deployment of cutting-edge solar technologies and enabling the generation of significant amounts of clean energy. The knowledge gained from studying solar resources in this unique environment not only contributes to advancements in solar technology but also serves as a valuable blueprint for harnessing solar power in other regions with similar radiation characteristics. There are numerous reasons that explain why the thermochemical process [1,2] plays an important role in producing hydrogen (H2) in arid and hyperarid zones. Zones with high solar radiation have the following advantage: the high levels of solar radiation present in arid regions, such as the Atacama Desert, make them ideal for harnessing solar power.
Water electrolysis is a promising method for hydrogen production due to its ability to produce high-purity hydrogen, its integration with renewable energy sources, and its adaptability to different scales of operation. However, its high energy consumption and the challenge of reducing the use of exotic materials remain significant challenges for this type of technology, affecting energy-conversion efficiency and posing a major barrier to its large-scale adoption. Although thermochemical solar cycles using ceria have 25% annual efficiency compared to 14% for the electrolysis process [3,4], the cycles are still impeded by a low level of technological preparation for the production of low-emission hydrogen [5,6,7,8].
The utilization of abundant solar energy in thermochemical processes can result in the generation of hydrogen and has the following advantages. (i) The efficient splitting of water molecules into hydrogen and oxygen is facilitated using concentrated solar energy to achieve high temperatures in thermochemical water splitting. When high solar irradiance is available, this method can achieve higher efficiency than is achieved with other hydrogen-production methods. (ii) Hydrogen production is made sustainable and environmentally friendly by utilizing renewable solar energy in the process. This approach contributes to cleaner energy production by reducing dependence on fossil fuels and decreasing greenhouse gas emissions. (iii) Iron or iron oxide can be used as abundant and low-cost materials for the thermochemical process, which makes it economically viable. Cost-effective production is guaranteed in regions with high solar potential thanks to the high availability of these materials globally. (iv) Arid zones tend to have predictable and stable weather patterns with little rainfall and small variations in temperature. Solar thermochemical systems’ performance and reliability are enhanced by the consistent climatic conditions, which ensure reliable and continuous operation, and (v) technological advancements can be achieved by researching and developing solar thermochemical hydrogen production in high-irradiation zones with high solar potential. Through the application of this knowledge, global hydrogen-production efforts can benefit from improved efficiency, cost reduction, and the development of new materials. The thermochemical solar process (TSP) for recovering solar hydrogen is considered environmentally favourable and inexpensive. The electrodes used in this process are generally made from iron or iron oxide, materials that are abundant worldwide. The global size of the market for iron oxide is estimated to be valued at around USD 2629.4 million in 2023, with projections reaching USD 3.6 billion by the end of 2030 [9]. The thermocatalytic material used in the TSP is ceria [3,4]. This technology today yields low production of solar hydrogen due to technological problems [8]. Regarding production costs, it is estimated that hydrogen from solar photovoltaic energy and wind energy is between 4.0–9.0 USD/Kg H2 and could be below 1.5 USD/Kg H2, with the costs of photovoltaic solar electricity at 14 USD/MWh in 2030 [8].
Regarding thermal water-splitting using iron and iron oxides at high temperatures, the number of articles published has shown significant growth between the years 2007 and 2023, as shown in Figure 1, which illustrates the numbers of publications that contain the specific words, “thermal water splitting using iron and oxides at high temperatures” in Web of Science.
This review is essential, as it provides the fundamental theoretical information necessary to understand the challenges associated with using pure iron or iron oxides as thermocatalytic materials in thermochemical cycles for hydrogen production. It is crucial to comprehend the mechanisms of these materials’ reactions in the presence of water vapour from a thermodynamic perspective, identify the heterogeneous reactions that occur on non-catalytic surfaces through electrochemistry, and consider the oxidation mechanisms associated with reactions at high temperatures with steam.

2. Theoretical Background

Under specific thermodynamic conditions, the interaction between steam (H2O(g)) and pure iron metal (Fe) can produce hydrogen (H2) through the oxidation of Fe at temperatures ranging from 200 to 1000 °C. This oxidation reaction occurs in an environment rich in H2O(g) and can be expressed as follows [10]:
3 F e + 4 H 2 O   F e 3 O 4 + 4 H 2  
In 1926, Dunn et al. [11] proposed that high-temperature oxidation is governed solely by the diffusion of O2 through the oxide layer. The presence of H2O(g) accelerates the oxidation process of mild steel to twice the rate associated with Fe pure [12,13].
The production of hydrogen (H2) using magnetite (Fe3O4) was proposed by Nakamura et al. in 1977 [14,15]. However, the production of H2 through the oxidation of H2O(g) using concentrated solar radiation was proposed by Lede et al. [16] in 1983. Recent studies have introduced notable innovations in the application of solar thermochemical reactors for converting concentrated solar energy into chemical fuels. These advancements involve a two-step thermochemical process specifically designed to produce H2 from H2O(g). By harnessing the immense potential of solar energy, this innovative approach offers a sustainable and highly efficient method for generating H2. The exothermic reaction using H2O(g) to produce H2 is expressed as follows [14,15]:
3 F e O + H 2 O   F e 3 O 4 + H 2
The Fe3O4 produced in this reaction is transformed into FeO in a solar oven through an endothermic subprocess [15,17]. However, a significant barrier arises in this subprocess due to the premature fusion of FeO, which occurs at a temperature lower than the threshold temperature required for the complete thermodynamic cycle. This inherent challenge presents difficulties in effectively employing a thermochemical solar reactor for H2 production, limiting its efficiency and feasibility. The reaction is expressed as follows [14,18]:
F e 3 O 4 l 3 F e O   l + 1 2 O 2   ;   T > 1875   K  
High-temperature corrosion is influenced by (i) temperature, (ii) gas composition, (iii) exposure time and (iv) system pressure and can be characterised by the reduction in thickness (penetration) and the rate of thickness growth. The oxidation rate increases significantly with rising temperatures [19], creating a homogeneous medium and enhancing thermodynamic equilibrium [20]. The Gibbs free energy is expressed as follows:
G o = R T   l n ( K )
When a metal is exposed to corrosion at high temperatures (corroding metal, oxidation or thermal degradation), it yields a value of ΔG°, as determined by Equation (5), which is based on the reaction xM + yH2O → MxOy + yH2 [19]. The expression for the Gibbs free energy is as follows:
G o = y R T ln P H 2 P H 2 O
Meredig & Wolverton et al. [21] indicated that during the corrosion process, the contact of a metal (M) with steam can produce H2 by the following reaction [21]:
M O ( x 1 ) + H 2 O M O x + H 2
where the expression for the Gibbs free energy is as follows:
G o = H M O x H M O ( x 1 ) H H 2 O T S H 2 S H 2 O   0
In the above equation, MOx is the oxidized metal and MOx−1 is the reduced oxide.
The water-splitting reaction using FeO as an electrode is favourable only at temperatures below 800 °C, when ΔG° < 0. The theoretical conversion of FeO decreases with increasing temperature [22]. Kodama and Gokon et al. [23] indicated that the oxidation of FeO in the presence of water vapour is a spontaneous reaction (ΔG < 0) at temperatures below 1000 K.
In the use of iron (Fe) and its oxides at high temperatures for hydrogen (H2) production, there are various mechanisms that would allow material agglomeration to occur. The adhesion or clumping of particles notably affects the efficiency and stability of the thermochemical process. Agglomeration is influenced by three factors: (i) at high temperatures, it is normal for Fe and FenOm particles to fuse together, forming larger particles; (ii) changes in the crystalline structure of Fe and FenOm at high temperatures cause physical changes that lead to agglomeration; and (iii) the thermal expansion of these materials causes mechanical stresses that facilitate agglomeration. During the chemical loop process, the morphology and support materials of the iron oxides can influence the stability and effectiveness of the process, as oxygen carriers play a crucial role during reactions, causing changes in their physical structures across several high-temperature redox cycles. Moreover, factors such as the reducing atmosphere and the behaviour of specific iron oxides such as hematite and magnetite under certain high-temperature conditions can contribute to different agglomeration behaviours, which affect the mechanism of hydrogen evolution. [24,25]
Young et al. [26] suggested that various metal oxides can form volatile compounds through direct reactions with water vapour and that hydroxides and oxy-hydroxides can be produced by hydration [27]. On the other hand, Belton & Richardson et al. [28] demonstrated the existence of a volatile iron hydroxide at temperatures exceeding 1300 °C. Compared to non-volatile metal oxides, volatile metal oxides exhibit a higher oxygen-exchange capacity, which directly correlates with their hydrogen-production capacity [29].
Iron hydroxides modify their reactive properties under specific conditions at high temperatures. The volatile iron hydroxides are (i) Fe(OH)2 and (ii) Fe(OH)3. These hydroxides can form under reducing conditions. Fe(OH)2 is less stable than Fe(OH)3 and can easily dehydrate or decompose at high temperatures, potentially forming iron oxides and releasing water vapour. On the other hand, Fe(OH)3 can also undergo dehydration or decomposition when it is subjected to high temperatures. The stability, reactivity, and transformations of these hydroxides under elevated thermal conditions are key to optimizing the efficiency of hydrogen production in thermochemical cycles [30].
Table 1 displays the thermodynamic parameters (∆G°, ∆H°, ∆S°, and T) used to determine the activation energy for the reaction of water vapour with iron and the products of vapour dissociation into molecules, ions, and iron-oxidation products (oxides and hydroxides). Table 2 presents the typical thermodynamic parameters generated during the oxidation/reduction processes of metals in the presence of H2O(g) at high temperatures.

Kinetic Model

The mechanisms by which a pure metal or alloy is oxidized at high temperatures require a series of successive steps [27,41]. The steps are as follows:
  • Step (a) adsorption of a gaseous component,
  • Step (b) dissociation of the gaseous molecule and transfer of electrons,
  • Step (c) nucleation and growth of crystals,
  • Step (d) diffusion and transport of cations, anions, and electrons through the oxide layer.
Nevertheless, the reaction between a solid (Fe) and a gas (H2O(g)) corresponds to a heterogeneous reaction wherein the kinetics are totally or partially controlled for the mass-transport process. The product species, generated by convection and diffusion, can be gaseous or insoluble solids [42]. Diffusion and heat transfer intertwine with the chemical kinetics of non-catalytic heterogeneous reactions, with their respective diffusion equations extending to fluidized solids [43]. The reduction of iron oxide with hydrogen or the oxidation of iron with steam are two typical examples of such reactions. These reactions occur under conditions of a highly compacted unreacted solid with limited porosity, where the chemical reaction progresses rapidly while diffusion proceeds relatively slowly [44].
Wen et al. [44] consider that a superficial non-catalytic heterogeneous reaction consists of the following steps (Figure 2):
  • Step (a) diffusion of fluid reactants through the fluid film surrounding the solid.
  • Step (b) diffusion of the fluid reagents through the porous solid layer,
  • Step (c) adsorption of the fluid reagents on the surface of the solid reagent,
  • Step (d) chemical reaction with the solid surface,
  • Step (e) desorption of the fluid products from the solid reaction surface,
  • Step (f) diffusion of the product far from the reaction surface through the porous surface, the solid media, and the fluid film surrounding the solid.
A study by Surman et al. [45]. involved the oxidation of pure Fe at 500 °C using hydrogen, water vapour, or various mixtures of the two. The oxidation kinetics were compared with models of chemical diffusion and solid-state reactions. The diffusion of volatile iron in the form of Fe(OH)2 was identified as the rate-controlling step. This diffusion occurred from the metal-oxide interface through the porous oxide of the inner layer and ended with the deposition of Fe3O4 on the crystals of the outer layer, which acted as sinks.
The oxidation kinetics occur according to the parabolic law. The expression of the parabolic rate constant (Kw) is represented by the following expression:
K w = 4.8 · 10 4 · F p · T 1 / 2 · P H 2 O P H 2 · e 38,900 / R T  
where Kw is the parabolic rate constant in (g2/m4s), Fp is the porosity factor of the inner layer, and PH2O and PH2 are the partial pressures of H2O and H2, respectively.
Park et al. [46] studied the production of hydrogen using iron oxide and inert silica with H2O(g) at temperatures ranging from 460 to 700 °C and vapour partial pressures between 0.002 and 0.02 MPa. Initially, Fe is oxidized to Fe3O4; this is followed by the gradual oxidation of Fe3O4 to Fe2O3 in the second step. The researchers determined the equilibrium state of H2O(g) adsorption and found that the rate-determining step of the reaction is the desorption of hydrogen from the active sites.
The proposed initial oxidation rate when the partial pressure of water is fixed is represented by the following expression:
d X B d t t = 0 = b k 1 + K e _ a d s · P H 2 O P H 2 O b k 1 + K e _ a d s · P H 2 O K e P H 2
where b is the stoichiometric coefficient of the oxidation reaction from Fe to Fe2O3, k is the overall rate constant based on unit solid volume, Ke is the chemical equilibrium constant, Ke_ads is the equilibrium adsorption constant, and PH2O and PH2 are the partial pressures of water and hydrogen, respectively.
The heterogeneous kinetic models have been detailed extensively by Donovan & Berra et al. [47], Levenspiel et al. [48], and Klaewkla et al. [49] However, Wen & Wang et al. [50] proposed a kinetic model for the non-catalytic heterogeneous reaction in a solid-gas system, considering both heat and mass transfer as combined effects.
Valipour & Saboohi et al. [51] analysed multiple mathematical models and developed a comprehensive model for a multi-reactant system using porous granules within a sample of a mobile packed bed. Our model accounts for various factors, including external mass transfer, internal diffusion through the pores, chemical reactions, heat generation or consumption due to reactions, and heat transfer through effective conduction within the solid matrix.
The three main types of growth laws that have been established experimentally, the linear, parabolic and logarithmic laws, are illustrated in Table 3.
The Fe oxidation kinetics follow a linear−parabolic behaviour in the presence of H2O(g), suggesting a transitional process wherein the diffusion of the hydroxyl ion is a determining factor during the oxidation [52]. Fujii & Meussner et al. [56] indicated that the kinetic oxidation of iron at 1100 °C can be adjusted to a parabolic law during the first 3 h, after which a linear behaviour is established, increasing the weight per unit area of all specimens (the rate of weight gain was 6.2 mg/cm² per hour). The effect of temperature on the reaction rate is obtained by applying the Arrhenius equation, which is represented by the following expression [54,55]:
K = K o e G o R T
where K is the equilibrium constant, Ko is the reaction constant, ΔG° is the activation energy, R is the gas constant, and T is the absolute temperature.
If the reaction rate is governed by diffusion through the oxide layer in the solid state, the oxide thickness (rust) increases during the diffusion process. However, at the same time, the kinetics slow down over time [41]. In the cathodic subprocess, the reaction 2H2O(g) + 2e → 2OH + H2 occurs as a product of the high-temperature dissociation process. The scientific literature indicates that the formation of OH from H2O(g) is proportional to the concentration of H2O(g) and that the appearance of OH cannot be interpreted in terms of the simple dissociation of H2O(g). It is necessary to consider the pathway involved in the formation of intermediate products generated during ORR such as (i) H2O2 and (ii) HO2 [57].

3. Mechanisms of High-Temperature Oxidation

Despite extensive scientific knowledge about the oxidation of iron at both environmental and high temperatures due to the presence of oxygen ions and despite efforts to reduce or avoid iron corrosion, the electrochemical reaction mechanism that explains the presence of oxides such as FeO, Fe3O4, and Fe2O3 [58], especially when iron interacts with the hydroxyl ion (OH), remains insufficiently understood.
Shrinivasan et al. [59] developed a multi-step optical system that enables water decomposition for OH ion detection, demonstrating that rate constants can be accurately measured at lower temperatures, such as 500 K. This approach contrasts with alternative methods that require temperatures exceeding 2570 K [60].
Lede et al. [16] proposed a kinetic mechanism of thermal dissociation of H2O between 2000–3000 K, as represented by the following expressions:
H 2 O + G H + O H + G
H 2 O + H H 2 + O H
O H + H H 2 + O
O H + O O 2 + H
where G is the extinguishing gas (argon at room temperature or steam at 400–450 K and between 2.5–9.4 bar), H is the H+ ions, and OH is the OH ions.
Fe corrosion is a process of electrochemical dissolution involving electron transfer to an intermediate species formed by the interaction between ferrous ions and water. In the reduction of Fe, this intermediate species is found on the iron surface and must undergo a subsequent reaction to form Fe [61].
The initial rate of oxidation of an Fe or steel surface newly exposed to H2O(g) is always lower compared to the rate of corrosion induced by O2. Researchers Tuck et al. [12,52] suggested that while H2 tends to dissolve in solid metals in atomic rather than molecular form, it is unlikely to dissolve in an oxide network due to its large size, which impedes diffusion. Therefore, they propose that the cathodic reaction could occur within the scales.
The general reaction for the production of H2 is shown below:
2 H 2 O + 2 e   2 O H + H 2
The mechanism of evolution of H2 (HER) would occur on the surface of the pores and micropores, where the hydroxyls ( O H ) formed by the dissolution of the vapour are discharged through a cathodic reaction according to Figure 3.
Rahmel & Tobolski et al. [62] proposed that the oxide layer is still thin and flexible in the first step the of oxidation process. In the presence of H2O(g), the transport mechanisms are likely to occur specifically through cavity or pore formation [63]. The mechanism is represented by Figure 4.
Schuetze et al. [37] proposed that the H2O(g) is involved in the transport processes that lead to scale growth in the oxide layer. The researchers indicate that the Fe cations (Fe2+/Fe3+) can simultaneously diffuse towards the outer surface in this process. Based on the knowledge of the structure and diffusion of iron oxides, the oxidation mechanism of pure iron above 570 °C in an iron−oxygen systems can be better understood. In particular, in the large pores in the inner part of the scale of FeO, a circulation mechanism is assumed that consists of the oxidation of Fe in contact with H2O(g), releasing H2 to can move back in the pore towards the outer part of the scale, where it reduces the oxide, thus forming an H2O(g) molecule again [64].
The first reaction that occurs is from the formation of OH ions, which would increase cation vacancies and would also be the main diffusion species [11,12,52]; this reaction is expressed as follows:
F e + 2 H 2 O   F e 2 + + 2 O H + H 2
Yuan et al. [52] suggested the formation of ferrous oxide (FeO) as shown in Figure 5; that is, iron in the presence of water vapour at high temperatures will precipitate FeO and Fe3O4 [52] or Fe2O3 [36], as follows:
F e 2 + + O H F e O + H +
However, FeO in contact with Fe(OH)2 [65] is generated at temperatures above 1300 °C [28] and can form Fe2O3 in conjunction with the release of H2 [52].
Rahmel & Tobolski et al., in 1965 [62], proposed the existence of pores at the Fe/FeO interface during the oxidation process, where they form oxide bridges from the metal to the scale. This structure allows for further oxidation of the metal without substantial inhibition. A mixture of H2/H2O is formed in these pores, and through an oxidation/reduction mechanism, O2 is transported to the Fe surface.
An oxide of the FenOm type will typically contain a variety of defects. These defects are crucial for the transport of material through the oxide layer and thus play a critical role in the oxidation process [12,66].
Yuan et al. [52] suggested that during the initial period (before 5 h), the presence of a single layer of columnar grains may facilitate relatively fast transport of OH ions across grain boundaries. It is therefore reasonable to assume that the hydroxyl ions interact with the surface hematite and magnetite, forming FeO.
However, certain aspects remain under investigation. Stehle et al. [67] highlighted that at typical reaction temperatures (above 327 °C), the mobility of oxygen and metal ions is expected to be very high, so diffusion limitations are generally not considered significant. This exception applies to oxides that exceed a thickness of several microns, where the potential of Fe3+ must be taken into account [52,67,68].
4 O 2 + 2 F e 3 + + F e 2 +   F e 3 O 4
Another aspect that should be considered, as stated by Saunders et al. [27], is that oxide growth is influenced by the simultaneous processes of adsorption, dissociation, and diffusion of reagents, which are altered in the presence of water vapour. Table 4 displays the standard reduction potentials that have been considered in the electrochemical analyses of the oxidation of iron and iron oxides in steam.
From these reactions, the kinetics of electrochemical reactions can be understood, highlighting their significance for the application of metal oxide redox reactions in energy-conversion systems such as chemical loop systems and hydrogen storage [70].
In an activation regime, the speed of the electrochemical process is dictated solely by electron transfer, which controls the pace of the overall process [41]. Hence, the corrosion reaction can be expressed by the ionization of a metal. However, the possibility that this reaction occurs spontaneously under real conditions necessitates the study of the energy changes associated with the reaction. Moreover, some H+ ions migrate into the metal, forming H2. Consequently, the presence of H+ ions can promote stress-corrosion cracking through the process of hydrogen embrittlement [71].
FeO nucleation and growth are enhanced by increasing oxygen pressure [66,72]. Kogan et al. [73] studied the dissociation of water at temperatures of 2000, 2200, 2500, and 2800 K at a constant pressure of 0.05 bar. They determined that 25% of the water vapour dissociates at 2500 K and 55% at 2800 K, with the rate of dissociation increasing at higher temperatures.
Kodama et al. [74] reported that the separation of water through a thermochemical process at a pressure of 0.01 bar and a temperature of 2000 K was barely noticeable, but when the temperature was increased to 2500 K, the yield of H2 exceeded 15% at the same pressure. Young et al. [26] proposed that in pure steam or mixtures of water vapour and inert gas, the equilibrium value of PO2 (oxygen pressure) is determined by the degree of dissociation of H2O. In the case of pure steam, the dissociation of one mole of water produces x mol of H2 and x/2 mol of O2, with x calculated from the equilibrium expression shown in Equation (19), as follows:
K 1 2 = x 3 2 1 x 2 1 + x 2 P T
where PT is the total pressure and K1 is small (x << 1), such that the above expression approaches the following:
x = 2 K 1 2 P T 1 3
p O 2 = x · P T
Ehlers et al. [75] proposed that Fe(OH)2, formed within oxide scales at low oxygen pressure (PO2), migrates to the outer surface where, due to higher pressure, hematite (Fe2O3) is formed. Khanna et al. [64] suggested that iron, in the presence of oxygen, forms a mixed scale of three oxides (FeO, Fe2O3, and Fe3O4), with the composition varying according to temperature and the PO2.
Ketteler et al. [75], in their research, considered a Fe-H2O system to determine the stability of iron oxide as a function of the partial pressure of water, identifying the phase limit for water condensation as ranging from 125 K (1 × 10−11 mbar) to 373 K (1 bar). They focus on the functions of oxygen pressure and temperature, according to the equilibrium constant for the dissociation of water (2H2O → 2H2 + O2) (see Equation (22)). The researchers concludes that water in its gaseous state acts as an oxidizing agent and that the stability of iron oxides is determined by the partial pressure from the decomposition of water into hydrogen and oxygen.
k 1 = p O 2 p H 2 2 p H 2 O 2

4. Effect of pH on Oxidation and Temperature

The pH value is normally based on the equilibrium reaction of the dissociation of water (H2O → H+ + OH), which has an endothermic character, and the equilibrium of which shifts to the right with increasing temperature. Research indicates that at 300 °C (25 MPa), the concentrations of both H+ (stable in acidic solutions) [71] and OH (stable in basic solutions) are approximately three orders of magnitude higher than their concentrations in water at ambient temperature. Consequently, water can be considered both acidic and alkaline [76].
To produce green hydrogen via a thermochemical process (thermolysis), temperatures ranging from 800 to 1400 °C are necessary in the reactor [77]. This temperature range has been achieved using a solar thermal concentrator system with a down-beam configuration. Iron is considered for hydrogen production in the thermochemical solar process because the conversion of Fe3O4 to FeO improves significantly at higher thermal-reduction temperatures [52]. However, its application is challenging because in the thermal-reduction stage of the thermochemical cycle, Fe3O4 melts at temperatures above 2227 °C (Figure 6) [14,78] and FeO melts at temperatures as low as 1370 °C [18] or 1400 °C [23]. These factors complicate the design and operation of thermochemical reactors based on Fe3O4/FeO [18]. However, the thermal-oxidation stage (hydrolysis) is carried out at temperatures ranging from 200 to 1000 °C [10,14,17,38,67]. According to the scientific literature, the temperature affects the process because the oxidation of iron by steam is thermodynamically favourable within the temperature ranges of 127–527 °C [18,34], 650–750 °C [52], and 700–850 °C [36], although reaction temperatures ranging from 717 to 1127 °C [68] have also been investigated.

5. Fluidized Bed Reactors for Thermochemical Water Splitting and Their Materials

It is important to highlight that iron oxide is selected because it is an environmentally friendly base material that is relatively low-cost and is commonly used in redox reactions due to its high oxygen-by-weight ratio, which ensures sufficient reaction time [54].
According to the reviewed scientific literature, temperatures above 1300 °C can be achieved using a thermochemical solar reactor with metallic materials such as cerium oxide or zinc oxide in a fluidized bed [77,79]. The thermal outputs obtained in reactors using water steam and various materials in a fluidized bed are presented in Table 5. Table 5 presents the materials used in the production of H2 via fluidized beds, along with pertinent details such as (i) thermal potential, (ii) temperature ranges within the reactors, (iii) type of fluidized material, and (iv) particle size. Additionally, it includes information on the concentration relationships referenced in the literature. The iron-oxide materials investigated for water separation in the oxidation stage at the laboratory level are detailed in Table 6. Table 6 displays the materials used, focusing on iron-based compounds such as magnetite, which have been doped with materials including zircon, nickel, caesium, and gadolinium. It also details the sample type, partial pressure, environment type, and exposure time. Gokon et al. [39] observed that the rate of hydrogen production after steam injection reached a peak of 12.3 Ncm3/min at 25 min, after which point the rate of hydrogen production rapidly decreased. This demonstrates that hydrogen production through a fluidized bed is feasible. Table 7 below presents a list of thermochemical solar reactors that have utilized fluidized-bed materials and different gases for fluidization, considering the geometry of the reactor and the materials used in its construction. Additionally, the concentration ratio is provided

6. Discussion

The latest technological devices that have been designed and manufactured for the recovery of solar hydrogen using photovoltaic technology, such as solar concentration, direct water-splitting using solar thermal energy with Fe electrodes, and FenOm-type oxides operating at high temperatures are an example of the multiple types of sustainable alternatives to address the global energy crisis. This method directly harnesses solar energy, enhancing sustainability and scalability. By improving efficiency, solar thermal water-splitting presents a viable pathway for large-scale hydrogen production, contributing to the broader adoption of sustainable hydrogen technologies. Furthermore, using Fe or FenOm electrodes at high temperatures offers a complementary approach by potentially reducing overpotential through high-temperature operation; additionally, these materials are abundant and low-cost, whether they are obtained through direct extraction or recycling. The experimental setup for direct solar thermal water-splitting using Fe electrodes involves the use of concentrated solar energy to achieve and maintain the high temperatures necessary for the reaction. A solar concentrator focuses sunlight onto a reactor chamber, heating it to temperatures between 800 °C and 1000 °C. Within this high-temperature environment, the iron electrodes act as thermochemical catalysts, facilitating the water-splitting reaction by cycling between oxidation and reduction states. The abundance and low cost of iron make it an attractive choice for this process. High temperatures improve the catalytic activity and stability of iron, addressing some common challenges associated with non-precious-metal catalysts. This setup exemplifies a sustainable and efficient method for hydrogen production, leveraging solar energy and abundant materials to contribute to the development of renewable-energy technologies.
Non-precious transition metals such as cobalt and nickel are attractive for electrocatalysis due to their low cost. However, challenges related to conductivity, catalytic activity, maintenance of active surface sites, and geopolitical issues (with regard to cobalt) hamper their widespread application. Advances in materials science have focused on the improvement of iron-based alloys or compounds and on surface modifications with high corrosion resistance. These improvements can increase the catalytic properties, active-site density, and durability of the electrodes, overcoming problems of material degradation. High-temperature operation improves the conductivity and catalytic activity of Fe and FenOm, generating a higher density of active surface sites. The oxidation of iron in steam at high temperatures produces a significant amount of H2, demonstrating the high efficiency and reactivity of Fe under these thermal conditions. The high-temperature environment accelerates the reaction rate and improves the purity of the produced H2, making the process suitable for various applications. These results support the viability of using Fe and FenOm catalysts in solar thermochemical cycles.
To advance the technology of direct solar thermal water-splitting using Fe electrodes at high temperatures, several specific areas of future research have been recommended. Improvements in materials science focus on developing Fe-based alloys or composites, surface modifications, and corrosion-resistance techniques. These advancements can improve catalytic properties and increase active-site density and electrode durability, overcoming issues of material degradation. Economic-viability research involves cost-reduction strategies, life-cycle analysis, and scalability studies. These efforts aim to make the technology economically feasible by reducing manufacturing costs, understanding the full economic and environmental impacts, and exploring possibilities for large-scale production.
One of the main advantages of this method is the use of a thermochemical iron (Fe) catalyst, which is not only cost-effective but also widely available, primarily in copper slag—an industrial waste. This waste contains a variety of metal oxides that can be utilized. Despite the lack of evidence for copper slag’s use as a catalyst material for hydrogen production through thermolysis, it was found that copper slag in Chile contains a significant concentration of fayalite, comprising two moles of iron oxide (FeO) and one mole of silicon oxide (SiO2).
This utilization not only reduces the overall cost of the process but also enhances its sustainability and feasibility for large-scale implementation. It’s noteworthy that the United States Geological Survey (USGS) estimates there are vast iron-ore reserves, further emphasizing the accessibility of this material. For instance, in 2017, gross iron ore reserves were estimated at an impressive 170,000 million metric tons, with an iron content of approximately 82,000 million metric tons. However, the 2016 production was only 2106 million metric tons, highlighting the vast untapped potential.
Moreover, the technology’s capability to efficiently split water into hydrogen (H2) and oxygen (O2) at high temperatures is crucial. It offers an effective solution for storing excess solar energy as hydrogen, which can be utilized in fuel cells or other applications during periods of reduced or no sunlight. This effectively tackles one of the key issues with solar power—intermittency—and contributes to a more reliable and consistent energy supply. Nonetheless, it is critical to recognize that the high operating temperatures place considerable stress on the materials, especially the electrodes and electrolyzers. This necessitates the development of advanced engineering solutions to ensure the technology’s longevity, durability, and cost-effectiveness over the long term. Additionally, optimizing the system design and energy conversion efficiency remains a significant challenge that must be overcome to enhance the competitiveness of this process relative to other hydrogen-production methods.
While this technology has immense potential, its practicality and impact are inherently linked to the solar resources and local conditions of the region in which it is deployed. Geographical areas with abundant sunlight stand to benefit the most, making this innovation particularly well-suited for regions with high solar potential. As we continue to refine and develop this technology, it has the potential to play a crucial role in transitioning towards a more sustainable and environmentally friendly energy landscape.

7. Conclusions

Direct solar thermal water-splitting using Fe (catalyst metal) at high temperatures is a promising development in the field of renewable energy and hydrogen production. By harnessing solar power and affordable materials, it offers a sustainable and scalable solution to the challenges of clean hydrogen production. As this technology matures and addresses its current challenges, it could come to play a vital role in transitioning to a more sustainable and greener energy future, reducing our reliance on fossil fuels and lowering carbon emissions across various industries.
A review of the scientific literature suggests that iron will precipitate in the forms of ferrous oxide, hematite, and magnetite during the oxidation of iron in steam at high temperatures, generating hydrogen. It is therefore reasonable to assume that iron, iron oxide, and other metal oxides present in slags, such as copper slag, will also precipitate rust and hydrogen. The theoretical thermodynamic and electrochemical aspects of the reaction of metal oxides with water vapour are fundamental to the design of the solar reactor, as well as to the adequate selection of the catalyst metal.
Future recommendations will focus on determining whether mineral waste (copper slag) containing metal oxides (iron oxides) is feasible to use as a thermocatalytic material for hydrogen production. For this purpose, a morphological and thermochemical characterization of the copper slag produced in Chile will be conducted to determine its potential use in water-splitting using a thermochemical solar reactor.
Future research holds significant promise for advancing this technology and addressing various challenges. Here are some key areas of research that can help further develop this innovative approach: (i) materials science and corrosion resistance, (ii) efficiency enhancement, (iii) thermochemical-cycle optimization, (iv) integrated energy storage, (v) economic viability, (vi) long-term durability and reliability, (vii) scalability and modular design, (viii) hydrogen purity and quality, and (ix) market integration and policy support. Incentives to encourage the adoption of solar thermal water/steam-splitting could make this technology more efficient, cost-effective, and environmentally friendly. By addressing these research areas, we can potentially overcome the current challenges and pave the way for its widespread adoption as a clean and sustainable method for hydrogen production.

Author Contributions

Conceptualization, M.F. and F.M.G.M.; methodology, M.F. and F.M.G.M.; validation, A.S. (Alvaro Soliz), E.F., D.P., A.S. (Atul Sagade) and N.T.; investigation, M.F., N.T., D.P. and F.M.G.M.; writing—original draft preparation, M.F. and F.M.G.M.; writing—review and editing, M.F., A.S. (Atul Sagade), E.F., D.P., A.S. (Alvaro Soliz), N.T. and F.M.G.M.; visualization M.F., A.S. (Alvaro Soliz) and F.M.G.M.; supervision, F.M.G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors would like to thank the Programa de Doctorado en Energía Solar of the Universidad de Antofagasta, Chile. The authors are grateful for the support of ANID-Chile through the research projects FONDECYT Iniciación 11230550 and ANID/ FONDAP 1522A0006 Solar Energy Research Center SERC-Chile.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Budama, V.K.; Duarte, J.P.R.; Roeb, M.; Sattler, C. Potential of Solar Thermochemical Water-Splitting Cycles: A Review. Sol. Energy 2023, 249, 353–366. [Google Scholar] [CrossRef]
  2. Guo, Y.; Chen, J.; Song, H.; Zheng, K.; Wang, J.; Wang, H.; Kong, H. A Review of Solar Thermochemical Cycles for Fuel Production. Appl. Energy 2024, 357, 122499. [Google Scholar] [CrossRef]
  3. Siegel, N.P.; Miller, J.E.; Ermanoski, I.; Diver, R.B.; Stechel, E.B. Factors Affecting the Efficiency of Solar Driven Metal Oxide Thermochemical Cycles. Ind. Eng. Chem. Res. 2013, 52, 3276–3286. [Google Scholar] [CrossRef]
  4. Monnerie, N. Green Hydrogen Production at DLR; German Aerospace Center: Cologne, Germany, 2023. [Google Scholar]
  5. Poudel, M.B.; Lohani, P.C.; Acharya, D.; Kandel, D.R.; Kim, A.A.; Yoo, D.J. MOF Derived Hierarchical ZnNiCo-LDH on Vapor Solid Phase Grown CuxO Nanowire Array as High Energy Density Asymmetric Supercapacitors. J. Energy Storage 2023, 72, 108220. [Google Scholar] [CrossRef]
  6. Elayappan, V.; Shanmugam, R.; Chinnusamy, S.; Yoo, D.J.; Mayakrishnan, G.; Kim, K.; Noh, H.S.; Kim, M.K.; Lee, H. Three-Dimensional Bimetal TMO Supported Carbon Based Electrocatalyst Developed via Dry Synthesis for Hydrogen and Oxygen Evolution. Appl. Surf. Sci. 2020, 505, 144642. [Google Scholar] [CrossRef]
  7. Jeong, S.Y.; Song, J.; Lee, S. Photoelectrochemical Device Designs toward Practical Solar Water Splitting: A Review on the Recent Progress of BiVO4 and BiFeO3 Photoanodes. Appl. Sci. 2018, 8, 1388. [Google Scholar] [CrossRef]
  8. International Energy Agency. Global Hydrogen Review 2022; International Energy Agency: Paris, France, 2022. [Google Scholar]
  9. Noda, S.; Yamaguchi, Y. Estimation of Surface Iron Oxide Abundance with Suppression of Grain Size and Topography Effects. Ore Geol. Rev. 2017, 83, 312–320. [Google Scholar] [CrossRef]
  10. Whitney, W.R. The Corrosion of Iron. J. Am. Chem. Soc. 1903, 25, 394–406. [Google Scholar] [CrossRef]
  11. Dunn, J.S. The High Temperature Oxidation of Metals. Proc. R. Soc. London. Ser. A Contain. Pap. A Math. Phys. Character 1926, 111, 203–209. [Google Scholar] [CrossRef]
  12. Tuck, C.W.; Odgers, M.; Sachs, K. The oxidation of iron at 950 °C in oxygen/water vapour mixtures. Corros. Sci. 1969, 9, 271–285. [Google Scholar] [CrossRef]
  13. Harikrishna, R.B.; Sharma, S.; Deka, H.; Sundararajan, T.; Rao, G.R. Thermochemical Production of Green Hydrogen Using Ferrous Scrap Materials. Int. J. Hydrogen Energy 2024, 52, 1488–1497. [Google Scholar] [CrossRef]
  14. Nakamura, T. Hydrogen Production from Water Utilizing Solar Heat at High Temperatures. Sol. Energy 1977, 19, 467–475. [Google Scholar] [CrossRef]
  15. Steinfeld, A.; Sanders, S.; Palumbo, R. Design Aspects of Solar Thermochemical Engineering-a Case Study: Two-Step Water-Splitting Cycle Using the Fe3O4/FeO Redox System. Sol. Energy 1999, 65, 43–53. [Google Scholar] [CrossRef]
  16. Lede, J.; Lapicque, F.; Villermaux, J. Production of Hydrogen by Direct Thermal Decomposition of Water. Int. J. Hydrogen Energy 1983, 8, 675–679. [Google Scholar] [CrossRef]
  17. Fernandez Saavedra, R. Bibliographic Review about Solar Hydrogen Production Through Thermochemical Cycles. In Revision Bibliografica sobre la Produccion de Hidrogeno Solar Mediante Ciclos Termoquimicos; CIEMAT: Madrid, Spain, 2008. [Google Scholar]
  18. Scheffe, J.R.; Steinfeld, A. Oxygen Exchange Materials for Solar Thermochemical Splitting of H2O and CO2: A Review. Mater. Today 2014, 17, 341–348. [Google Scholar] [CrossRef]
  19. Perez, N. Electrochemistry and Corrosion Science; Springer: Berlin/Heidelberg, Germany, 2004. [Google Scholar]
  20. Etievant, C. Solar Energy Materials Solar High-Temperature Direct Water Splitting a Review of Experiments in France. Sol. Energy Mater. 1991, 24, 413–440. [Google Scholar] [CrossRef]
  21. Meredig, B.; Wolverton, C. First-Principles Thermodynamic Framework for the Evaluation of Thermochemical H2O-Or CO2-Splitting Materials. Phys. Rev. B Condens. Matter Mater. Phys. 2009, 80, 245119. [Google Scholar] [CrossRef]
  22. Abanades, S. Metal Oxides Applied to Thermochemical Water-Splitting for Hydrogen Production Using Concentrated Solar Energy. ChemEngineering 2019, 3, 63. [Google Scholar] [CrossRef]
  23. Kodama, T.; Gokon, N. Thermochemical Cycles for High-Temperature Solar Hydrogen Production. Chem. Rev. 2007, 107, 4048–4077. [Google Scholar] [CrossRef]
  24. Barustan, M.I.A.; Jung, S.-M. Morphology of Iron and Agglomeration Behaviour during Reduction of Iron Oxide Fines. Met. Mater. Int. 2019, 25, 1083–1097. [Google Scholar] [CrossRef]
  25. Gao, Z.; Fu, F.; Niu, L.; Jin, M.; Wang, X. Effect of Support on Hydrogen Generation over Iron Oxides in the Chemical Looping Process. RSC Adv. 2021, 11, 37552–37558. [Google Scholar] [CrossRef] [PubMed]
  26. Young, J. Effects of Water Vapour on Oxidation. Corros. Ser. 2008, 1, 455–495. [Google Scholar]
  27. Saunders, S.R.J.; Monteiro, M.; Rizzo, F. The Oxidation Behaviour of Metals and Alloys at High Temperatures in Atmospheres Containing Water Vapour: A Review. Prog. Mater. Sci. 2008, 53, 775–837. [Google Scholar] [CrossRef]
  28. Belton, G.R.; Richardson, F.D. A Volatile Iron Hydroxide. Trans. Faraday Soc. 1962, 58, 1562–1572. [Google Scholar] [CrossRef]
  29. Mao, Y.; Gao, Y.; Dong, W.; Wu, H.; Song, Z.; Zhao, X.; Sun, J.; Wang, W. Hydrogen Production via a Two-Step Water Splitting Thermochemical Cycle Based on Metal Oxide—A Review. Appl. Energy 2020, 267, 114860. [Google Scholar] [CrossRef]
  30. Naikoo, G.A.; Arshad, F.; Hassan, I.U.; Tabook, M.A.; Pedram, M.Z.; Mustaqeem, M.; Tabassum, H.; Ahmed, W.; Rezakazemi, M. Thermocatalytic Hydrogen Production through Decomposition of Methane-A Review. Front. Chem. 2021, 9, 736801. [Google Scholar] [CrossRef]
  31. Robie, R.A.; Hemingway, B.S. Thermodynamic Properties of Minerals and Related Substances at 298.15 K and 1 Bar (105 Pascals) Pressure and at Higher Temperatures (No. 2131); U.S. G.P.O.: Washington, DC, USA, 1995.
  32. Barin, I.; Platzhi, G. Thermochemical Data of Pure Substances; VCh: Weinheim, Germany, 1989; Volume 304, ISBN 3527287450. [Google Scholar]
  33. Navrotsky, A.; Mazeina, L.; Majzlan, J. Size-Driven Structural and Thermodynamic Complexity in Iron Oxides. Science 2008, 319, 1635–1638. [Google Scholar] [CrossRef]
  34. Svoboda, K.; Slowinski, G.; Rogut, J.; Baxter, D. Thermodynamic Possibilities and Constraints for Pure Hydrogen Production by Iron Based Chemical Looping Process at Lower Temperatures. Energy Convers. Manag. 2007, 48, 3063–3073. [Google Scholar] [CrossRef]
  35. Birks, N.; Meier, G.H.; Pettit, F.S. High-Temperature Corrosion Resistance. JOM 1987, 39, 28–31. [Google Scholar] [CrossRef]
  36. Jeong, M.H.; Lee, D.H.; Bae, J.W. Reduction and Oxidation Kinetics of Different Phases of Iron Oxides. Int. J. Hydrogen Energy 2015, 40, 2613–2620. [Google Scholar] [CrossRef]
  37. Schuetze, M. Fundamentals of High Temperature Corrosion; Materials Science and Technology: A Comprehensive Treatment: Corrosion and Environmental Degradation; Wiley: Hoboken, NJ, USA, 2000; Volume I+II, pp. 67–130. [Google Scholar] [CrossRef]
  38. Kodama, T.; Bellan, S.; Gokon, N.; Cho, H.S. Particle Reactors for Solar Thermochemical Processes. Sol. Energy 2017, 156, 113–132. [Google Scholar] [CrossRef]
  39. Gokon, N.; Mataga, T.; Kondo, N.; Kodama, T. Thermochemical Two-Step Water Splitting by Internally Circulating Fluidized Bed of NiFe2O4 Particles: Successive Reaction of Thermal-Reduction and Water-Decomposition Steps. Int. J. Hydrogen Energy 2011, 36, 4757–4767. [Google Scholar] [CrossRef]
  40. Palumbo, R.; Diver, R.B.; Larson, C.; Coker, E.N.; Miller, J.E.; Guertin, J.; Schoer, J.; Meyer, M.; Siegel, N.P. Solar Thermal Decoupled Water Electrolysis Process I: Proof of Concept. Chem. Eng. Sci. 2012, 84, 372–380. [Google Scholar] [CrossRef]
  41. Genescá, J.; Meas, Y.; Rodríguez, F.; Mendoza, J.; Durán, R.; Uruchurtu, J.; González, J.G. Técnicas Electroquímicas Para el Control y Estudio de la Corrosión; Faculty of Chemistry, UNAM: Mexico City, Mexico, 2002; Volume 244. [Google Scholar]
  42. Bircumshaw, L.L.; Riddiford, A.C. Transport Control in Heterogeneous Reactions. Q. Rev. Chem. Soc. 1952, 6, 157–185. [Google Scholar] [CrossRef]
  43. Hougen, O.A. Diffusion and Heat Exchange in Chemical Kinetics. J. Am. Chem. Soc. 1956, 78, 885–886. [Google Scholar] [CrossRef]
  44. Wen, C.Y. Noncatalytic Heterogeneous Solid-Fluid Reaction Models. Ind. Eng. Chem. 1968, 60, 34–54. [Google Scholar] [CrossRef]
  45. Surman, P.L. The oxidation of iron at controlled oxygen partial pressures. hydrogen/water vapour. Corros. Sci. 1973, 13, 113–124. [Google Scholar] [CrossRef]
  46. Park, H.C.; Sakai, Y.; Kimura, S.; Tone, S.; Otake, T. A Rate Analysis for Oxidation of Porous Reduced Iron with Water Vapor. J. Chem. Eng. Jpn. 1984, 17, 395–399. [Google Scholar] [CrossRef]
  47. Donovan, R.J.; Berra, Y. External Diffusion Effects on Heterogeneous Reactions. In Elements of Chemical Reaction Engineering; Pearson: London, UK, 1999; pp. 686–738. [Google Scholar]
  48. Levenspiel, O. Ingeniería de Las Reacciones Químicas, 3rd ed.; Editorial Limusa Wiley: Mexico City, Mexico, 2010. [Google Scholar]
  49. Klaewkla, R.; Arend, M.; Hoelderich, W.F. A Review of Mass Transfer Controlling the Reaction Rate in Heterogeneous Catalytic Systems; INTECH Open Access Publisher: Rijeka, Croatia, 2011; Volume 5. [Google Scholar]
  50. Wen, C.Y.; Wang, S.C. Thermal and Diffusional Effects in Noncatalytic Solid Gas Reactions. Ind. Eng. Chem. 1970, 62, 30–51. [Google Scholar] [CrossRef]
  51. Valipour, M.S.; Saboohi, Y. Modeling of Multiple Noncatalytic Gas-Solid Reactions in a Moving Bed of Porous Pellets Based on Finite Volume Method. Heat Mass Transf. Waerme Stoffuebertragung 2007, 43, 881–894. [Google Scholar] [CrossRef]
  52. Yuan, J.; Wang, W.; Zhu, S.; Wang, F. Comparison between the Oxidation of Iron in Oxygen and in Steam at 650–750 °C. Corros. Sci. 2013, 75, 309–317. [Google Scholar] [CrossRef]
  53. Young, D.J.; Watson, S. High-Temperature Corrosion in Mixed Gas Environments. Oxid. Met. 1995, 44, 239–264. [Google Scholar] [CrossRef]
  54. Go, K.S.; Son, S.R.; Kim, S.D. Reaction Kinetics of Reduction and Oxidation of Metal Oxides for Hydrogen Production. Int. J. Hydrogen Energy 2008, 33, 5986–5995. [Google Scholar] [CrossRef]
  55. Li, P.; Guo, H.; Gao, J.; Min, J.; Yan, B.; Chen, D.; Seetharaman, S. Novel Concept of Steam Modification towards Energy and Iron Recovery from Steel Slag: Oxidation Mechanism and Process Evaluation. J. Clean. Prod. 2020, 254, 119952. [Google Scholar] [CrossRef]
  56. Fujji, C.T.; Meussner, R.A. The Mechanism of the High-Temperature Oxidation of Iron-Chromium Alloys in Water Vapor. J. Electrochem. Soc. 1964, 111, 1215. [Google Scholar] [CrossRef]
  57. Bauer, S.H.; Schott, G.L.; Duff, R.E. Kinetic Studies of Hydroxyl Radicals in Shock Waves. I. The Decomposition of Water between 2400° and 3200 °K. J. Chem. Phys. 1958, 28, 1089–1096. [Google Scholar] [CrossRef]
  58. Zhang, H.; Zhang, X.; Yang, D.; Shuai, Y.; Lougou, B.G.; Pan, Q.; Wang, F. Selection of Iron-Based Oxygen Carriers for Two-Step Solar Thermochemical Splitting of Carbon Dioxide. Energy Convers. Manag. 2023, 279, 116772. [Google Scholar] [CrossRef]
  59. Srinivasan, N.K.; Michael, J.V. The Thermal Decomposition of Water. Int. J. Chem. Kinet. 2006, 38, 211–219. [Google Scholar] [CrossRef]
  60. Homer, J.B.; Hurle, I.R. The Dissociation of Water Vapour behind Shock Waves. Proc. R. Soc. London. A. Math. Phys. Sci. 1970, 314, 585–598. [Google Scholar] [CrossRef]
  61. Brockris, J.O.; Reddy, A.K.N. Electroquimica Moderna; Editorial Reverté: Barcelona, Spain, 2003; Volume 2. [Google Scholar]
  62. Rahmel, A.; Tobolsk, J. Einfluss von wasserdampf und kohlendioxyd auf die oxydation von eisen in sauerstoff bei hohen temperaturen. Corros. Sci. 1965, 5, 333. [Google Scholar] [CrossRef]
  63. Gareev, K.G. Diversity of Iron Oxides: Mechanisms of Formation, Physical Properties and Applications. Magnetochemistry 2023, 9, 119. [Google Scholar] [CrossRef]
  64. Khanna, A.S. Introduction to High Temperature Oxidation and Corrosion; ASM International: Detroit, MI, USA, 2002; ISBN 0871707624. [Google Scholar]
  65. Sanchez-Caldera, L.E.; Griffith, P.; Asm, F.; Rabinowicz, E.E.; Asme, F. The Mechanism of Corrosion-Erosion in Steam Extraction Lines of Power Stations. J. Eng. Gas Turbines Power 1988, 110, 180–184. [Google Scholar] [CrossRef]
  66. Lawless, K.R. The Oxidation of Metals. Rep. Prog. Phys. 1974, 37, 231. [Google Scholar] [CrossRef]
  67. Stehle, R.C.; Bobek, M.M.; Hooper, R.; Hahn, D.W. Oxidation Reaction Kinetics for the Steam-Iron Process in Support of Hydrogen Production. Int. J. Hydrogen Energy 2011, 36, 15125–15135. [Google Scholar] [CrossRef]
  68. Stehle, R.C.; Bobek, M.M.; Hahn, D.W. Iron Oxidation Kinetics for H2 and CO Production via Chemical Looping. Int. J. Hydrogen Energy 2015, 40, 1675–1689. [Google Scholar] [CrossRef]
  69. Ávila, J.; Genescá, J. Más Allá de La Herrumbre 1; Fondo de Cultura Económica: Mexico City, Mexico, 2013; ISBN 6071603196. [Google Scholar]
  70. Kosaka, F.; Hatano, H.; Oshima, Y.; Otomo, J. Iron Oxide Redox Reaction with Oxide Ion Conducting Supports for Hydrogen Production and Storage Systems. Chem. Eng. Sci. 2015, 123, 380–387. [Google Scholar] [CrossRef]
  71. McCafferty, E. Thermodynamics of Corrosion: Pourbaix Diagrams. In Introduction to Corrosion Science; Springer: New York, NY, USA, 2010; pp. 95–117. [Google Scholar]
  72. Pignocco, A.J.; Pellissier, G.E. Leed studies of oxygen adsorption and oxide formation on an (011) iron surface. Surf. Sci. 1967, 7, 261–278. [Google Scholar] [CrossRef]
  73. Kogan, A. Direct solar thermal splitting of water and on-site separation of the products. experimental feasibility study. J. Hydrogen Energy 1998, 23, 9–98. [Google Scholar] [CrossRef]
  74. Kodama, T. High-Temperature Solar Chemistry for Converting Solar Heat to Chemical Fuels. Prog. Energy Combust. Sci. 2003, 29, 567–597. [Google Scholar] [CrossRef]
  75. Ehlers, J.; Young, D.J.; Smaardijk, E.J.; Tyagi, A.K.; Penkalla, H.J.; Singheiser, L.; Quadakkers, W.J. Enhanced Oxidation of the 9%Cr Steel P91 in Water Vapour Containing Environments. Corros. Sci. 2006, 48, 3428–3454. [Google Scholar] [CrossRef]
  76. Kritzer, P. Corrosion in High-Temperature and Supercritical Water and Aqueous Solutions: A Review. J. Supercrit. Fluids 2004, 29, 1–29. [Google Scholar] [CrossRef]
  77. Siegel, N.P.; Ermanoski, I. A Beam-Down Central Receiver for Solar Thermochemical Hydrogen Production; American Solar Energy Society: Baltimore, MD, USA, 2013. [Google Scholar]
  78. Miyaoka, H. Thermochemical Water Splitting by Concentrated Solar Power. In Lecture Notes in Energy; Springer: Berlin/Heidelberg, Germany, 2016; Volume 32, pp. 137–151. [Google Scholar]
  79. Kodama, T.; Gokon, N.; Matsubara, K.; Yoshida, K.; Koikari, S.; Nagase, Y.; Nakamura, K. Flux Measurement of a New Beam-down Solar Concentrating System in Miyazaki for Demonstration of Thermochemical Water Splitting Reactors. In Proceedings of the Energy Procedia; Elsevier Ltd.: Amsterdam, The Netherlands, 2014; Volume 49, pp. 1990–1998. [Google Scholar]
  80. Kodama, T.; Gokon, N.; Enomoto, S.I.; Itoh, S.; Hatamachi, T. Coal Coke Gasification in a Windowed Solar Chemical Reactor for Beam-down Optics. J. Sol. Energy Eng. Trans. ASME 2010, 132, 041004. [Google Scholar] [CrossRef]
  81. Gokon, N.; Izawa, T.; Kodama, T. Steam Gasification of Coal Cokes by Internally Circulating Fluidized-Bed Reactor by Concentrated Xe-Light Radiation for Solar Syngas Production. Energy 2015, 79, 264–272. [Google Scholar] [CrossRef]
  82. Bellan, S.; Gokon, N.; Matsubara, K.; Cho, H.S.; Kodama, T. Heat Transfer Analysis of 5kWth Circulating Fluidized Bed Reactor for Solar Gasification Using Concentrated Xe Light Radiation. Energy 2018, 160, 245–256. [Google Scholar] [CrossRef]
  83. Koepf, E.; Advani, S.G.; Steinfeld, A.; Prasad, A.K. A Novel Beam-down, Gravity-Fed, Solar Thermochemical Receiver/Reactor for Direct Solid Particle Decomposition: Design, Modeling, and Experimentation. Int. J. Hydrogen Energy 2012, 37, 16871–16887. [Google Scholar] [CrossRef]
  84. Kodama, T.; Gokon, N.; Cho, H.S.; Bellan, S.; Matsubara, K.; Inoue, K. Particle Fluidized Bed Receiver/Reactor with a Beam-down Solar Concentrating Optics: Performance Test of Two-Step Water Splitting with Ceria Particles Using 30-KWth Sun-Simulator. In Proceedings of the AIP Conference Proceedings; American Institute of Physics Inc.: College Park, MD, USA, 2018; Volume 2033. [Google Scholar]
  85. Bellan, S.; Kodama, T.; Seok Cho, H.; Kim, J.S. Hydrogen Production by Solar Fluidized Bed Reactor Using Ceria: Euler-Lagrange Modelling of Gas-Solid Flow to Optimize the Internally Circulating Fluidized Bed. J. Therm. Sci. Technol. 2022, 17, 22-00076. [Google Scholar] [CrossRef]
  86. Gokon, N.; Hasegawa, T.; Takahashi, S.; Kodama, T. Thermochemical Two-Step Water-Splitting for Hydrogen Production Using Fe-YSZ Particles and a Ceramic Foam Device. Energy 2008, 33, 1407–1416. [Google Scholar] [CrossRef]
  87. Kodama, T.; Gokon, N.; Cho, H.S.; Matsubara, K.; Kaneko, H.; Senuma, K.; Itoh, S.; Yokota, S.N. Particles Fluidized Bed Receiver/Reactor Tests with Quartz Sand Particles Using a 100-KWth Beam-down Solar Concentrating System at Miyazaki. In Proceedings of the AIP Conference Proceedings; American Institute of Physics Inc.: College Park, MD, USA, 2017; Volume 1850. [Google Scholar]
  88. Gokon, N.; Kumaki, S.; Miyaguchi, Y.; Bellan, S.; Kodama, T.; Cho, H. Development of a 5kWth Internally Circulating Fluidized Bed Reactor Containing Quartz Sand for Continuously-Fed Coal-Coke Gasification and a Beam-down Solar Concentrating System. Energy 2019, 166, 1–16. [Google Scholar] [CrossRef]
  89. Kiwan, S.; Soud, Q.R. Numerical Investigation of Sand-Basalt Heat Storage System for Beam-down Solar Concentrators. Case Stud. Therm. Eng. 2019, 13, 100372. [Google Scholar] [CrossRef]
  90. Yang, S.; Wang, J.; Lund, P.D.; Jiang, C.; Li, X. High Performance Integrated Receiver-Storage System for Concentrating Solar Power Beam-down System. Sol. Energy 2019, 187, 85–94. [Google Scholar] [CrossRef]
  91. Yang, S.; Wang, J.; Lund, P.D.; Jiang, C.; Li, X. Modelling and Performance Evaluation of an Integrated Receiver-Storage for Concentrating Solar Power Beam-down System under Heterogeneous Radiative Conditions. Sol. Energy 2019, 188, 1264–1273. [Google Scholar] [CrossRef]
  92. Díaz-Heras, M.; Barreneche, C.; Belmonte, J.F.; Calderón, A.; Fernández, A.I.; Almendros-Ibáñez, J.A. Experimental Study of Different Materials in Fluidized Beds with a Beam-down Solar Reflector for CSP Applications. Sol. Energy 2020, 211, 683–699. [Google Scholar] [CrossRef]
  93. Fresno, F.; Yoshida, T.; Gokon, N.; Fernández-Saavedra, R.; Kodama, T. Comparative Study of the Activity of Nickel Ferrites for Solar Hydrogen Production by Two-Step Thermochemical Cycles. Int. J. Hydrogen Energy 2010, 35, 8503–8510. [Google Scholar] [CrossRef]
  94. Taylor, R.W.; Berjoan, R.; Coutures, P. Solar gasification of carbonaceous materials. Sol. Energy 1983, 30, 513–525. [Google Scholar] [CrossRef]
  95. Flamant, G.; Olalde, G. High temperature solar gas heating comparison between packed and fluidized bed receivers. Sol. Energy 1983, 31, 463–471. [Google Scholar] [CrossRef]
  96. Flamant, G.; Gauthier, D.; Boudhari, C.; Flitris, Y. A 50 KW Fluidized Bed High Temperature Solar Receiver: Heat Transfer Analysis. J. Sol. Energy Eng. 1988, 110, 313–320. [Google Scholar] [CrossRef]
  97. Steinfeld, A.; Frei, A.; Kuhn, P.; Wuillemin, D. solar thermal production of zinc and syngas via combined zno-reduction and CH4 reforming processes. J. Hydrogen Energy 1995, 20, 793–804. [Google Scholar] [CrossRef]
  98. Gokon, N.; Takahashi, S.; Yamamoto, H.; Kodama, T. Thermochemical Two-Step Water-Splitting Reactor with Internally Circulating Fluidized Bed for Thermal Reduction of Ferrite Particles. Int. J. Hydrogen Energy 2008, 33, 2189–2199. [Google Scholar] [CrossRef]
  99. Nikulshina, V.; Gebald, C.; Steinfeld, A. CO2 Capture from Atmospheric Air via Consecutive CaO-Carbonation and CaCO3-Calcination Cycles in a Fluidized-Bed Solar Reactor. Chem. Eng. J. 2009, 146, 244–248. [Google Scholar] [CrossRef]
  100. Kodama, T.; Gokon, N.; Cho, H.S.; Matsubara, K.; Etori, T.; Takeuchi, A.; Yokota, S.N.; Ito, S. Particles Fluidized Bed Receiver/Reactor with a Beam-down Solar Concentrating Optics: 30-KWth Performance Test Using a Big Sun-Simulator. In Proceedings of the AIP Conference Proceedings; American Institute of Physics Inc.: College Park, MD, USA, 2016; Volume 1734. [Google Scholar]
  101. Bellan, S.; Gokon, N.; Matsubara, K.; Cho, H.S.; Kodama, T. Numerical and Experimental Study on Granular Flow and Heat Transfer Characteristics of Directly-Irradiated Fluidized Bed Reactor for Solar Gasification. Int. J. Hydrogen Energy 2018, 43, 16443–16457. [Google Scholar] [CrossRef]
  102. Hoskins, A.L.; Millican, S.L.; Czernik, C.E.; Alshankiti, I.; Netter, J.C.; Wendelin, T.J.; Musgrave, C.B.; Weimer, A.W. Continuous On-Sun Solar Thermochemical Hydrogen Production via an Isothermal Redox Cycle. Appl. Energy 2019, 249, 368–376. [Google Scholar] [CrossRef]
  103. Jiang, K.; Kong, Y.; Xu, C.; Ge, Z.; Du, X. Experimental Performance of Gas-Solid Countercurrent Fluidized Bed Particle Solar Receiver with High-Density Suspension. Appl. Therm. Eng. 2022, 213, 118661. [Google Scholar] [CrossRef]
Figure 1. Number of publications per year with the keywords “thermal water splitting using iron and oxides at high temperatures” retrieved from Web of Science.
Figure 1. Number of publications per year with the keywords “thermal water splitting using iron and oxides at high temperatures” retrieved from Web of Science.
Applsci 14 07056 g001
Figure 2. Step diagram of the non-catalytic superficial heterogeneous reaction of the electrochemical dissolution of iron in water vapour.
Figure 2. Step diagram of the non-catalytic superficial heterogeneous reaction of the electrochemical dissolution of iron in water vapour.
Applsci 14 07056 g002
Figure 3. Diagram of the electrochemical dissolution of iron. The symbol “?” means that the question remains about the interaction between metal (Fe) and water vapor.
Figure 3. Diagram of the electrochemical dissolution of iron. The symbol “?” means that the question remains about the interaction between metal (Fe) and water vapor.
Applsci 14 07056 g003
Figure 4. The mechanism in O2-H2O vapour mixtures after long periods (adapted from [62]).
Figure 4. The mechanism in O2-H2O vapour mixtures after long periods (adapted from [62]).
Applsci 14 07056 g004
Figure 5. (a) Cross-sectional SEM image of oxides formed on iron after 1 h of oxidation at 650 °C and (b) cross-sectional SEM image of oxide scale formed on iron after 10 h of steam oxidation at 650 °C, where the outermost white layer is nickel-plated (from [52]).
Figure 5. (a) Cross-sectional SEM image of oxides formed on iron after 1 h of oxidation at 650 °C and (b) cross-sectional SEM image of oxide scale formed on iron after 10 h of steam oxidation at 650 °C, where the outermost white layer is nickel-plated (from [52]).
Applsci 14 07056 g005
Figure 6. Operating temperature of typical thermochemical water-splitting cycles in the reduction stage.
Figure 6. Operating temperature of typical thermochemical water-splitting cycles in the reduction stage.
Applsci 14 07056 g006
Table 1. Thermodynamic properties of water, oxides, and hydroxides.
Table 1. Thermodynamic properties of water, oxides, and hydroxides.
Chemical SpeciesGibbs Free Energy
ΔG (kJ/mol)
Enthalpy
ΔH (kJ/mol)
Entropy
S (kJ/mol)
Temp. K
(1 atm)
Refs.
H2O(l)−237.1−285.870.0298 [31,32]
H2O(g)−228.6−241.8188.8298 [31,32]
−214.1−244.7213.8600 [32]
−198.2−247.1228.6900
H2(g)00130.6298 [31,32]
00151.0600[32]
00163.1900
OH (aqueous ion)−157.3−230.0−10.7298 [31]
OH(g)34.738.9183.7298 [32]
29.538.9204.4600
24.938.4216.5900
Fe00027.09298 [31]
Fe2+−90.0−91.1−107.1298 [31]
Fe3+−16.7−49.9−280.0298 [31]
Fe.947O−244.9−266.356.6298 [31]
−245.1−266.357.5298[32]
−224.8−263.793.0600
−205.8−262.7115.1900
FeO−251.4−272.060.6298 [31,32]
−231.7−269.497.3600[32]
−213.1−268.4120.2900
F e 2 O 3 −774.4−826.287.4298 [31,32]
−742.3−824.287.4298[32]
−661.4−817.6173.3600
−585.3−808.4235.4900
F e 3 O 4 −731.4−811.693.0298 [33]
−1012.7−1115.7146.1298 [31]
−1015.2−1118.4146.1298[32]
−914.4−1107.4272.1600
−821.8−1088.6368.3900
FeO(OH)−491.8−562.660.4298 [31]
Fe(OH)2−486.9−568.987.9298[32]
−405.7−564.1160.4600
−327.6−558.9207.7900
Table 2. Thermodynamic properties of the reaction of iron or oxides in water Vapour.
Table 2. Thermodynamic properties of the reaction of iron or oxides in water Vapour.
ReactionGibbs Free Energy
ΔG (kJ/mol)
Enthalpy
ΔH (kJ/mol)
Temp. K
(pres. atm)
Refs.
0.75   F e + H 2 O   0.25 F e 3 O 4 + 4 H 2 −14.52−32.09600[34]
F e + 2 H 2 O     F e O + H 2 −199.3-1000[35]
F e + 2 H 2 O   F e O H 2 ( g ) + H 2 38.9 + 5 × 80T (cal)--[28]
2 F e + 3 H 2 O   F e 2 O 3 + 3 H 2 48.9-973–1123[36]
3 F e + 4 H 2 O   F e 3 O 4 + 4 H 2 11.0-973–1123[36,37]
3 F e O + H 2 O   F e 3 O 4 + H 2 -−33.6873[22,23,38,39]
F e 2 O 3 + H 2 O   2 F e O ( O H ) 10.39−47.54400[34]
F e 3 O 4 + 2 H 2 O   F e ( O H ) 2 + 2 F e O ( O H ) 41.02−79.23400[34]
F e 3 O 4 + 4 H 2 O   F e ( O H ) 2 + 2 F e ( O H ) 3 228.22−120.48600[34]
2 F e 3 O 4 + H 2 O   3 F e 2 O 3 + H 2 40.5--[40]
2 F e 3 O 4 + 4 H 2 O   6 F e O O H + H 2 44.4--[40]
Table 3. Reaction-rate equations applicable to high-temperature oxidation.
Table 3. Reaction-rate equations applicable to high-temperature oxidation.
EquationObservationRefs.
Logarithmic reaction rate:
X = K r e a c t i o n ln   t + t o + A It represents the initial oxidation states at low temperatures.[41]
X = K r e a c t i o n ln   B t + 1 -[41]
Linear reaction rate:
d x d t = K L i n e a l It represents a constant rate of oxide growth applicable at very high temperatures.[41]
X = K L i n e a l t + C L i n e a l -[41]
W A =   K L   t -[52]
Parabolic reaction rate:
d X d t = K P a r a b x It adjusts to the processes controlled by the diffusion of species.[41]
X 2 = K p a r a b t + C p a r a b -[41,53]
W A 2 = K p a r a b t   -[52]
d X d t = k   f ( X ) The function f(X) depends on the reaction mechanism for diffusion in one, two, or three dimensions.[54]
d X d t = k T f ( X )
k T = k o e E a R T
The reaction rate of a sample of iron slag in water vapour.[55]
Where X is the thickness of oxide consumed per surface unit or the weight gain per area unit; ΔW/A is the mass gain per unit area (g/cm2); t is the time; k(T) is the reaction rate constant; f(X) is the function that represents the reaction mechanism; ko, Kreaction, KLineal, and KParab are reaction constants; Ea is the activation energy; R is the gas constant; T is the absolute temperature; A and B are constants; CLineal and CParab are constants of integration.
Table 4. Standard reduction potentials.
Table 4. Standard reduction potentials.
Half ReactionStandard Reduction Potential, E° (V)Ref.
H 2   2 H + + 2 e E° = 0.000 (V)[69]
F e   F e 2 + + 2 e E° = −0.447 (V)[69]
Reaction
3 F e + 4 H 2 O   F e 3 O 4 + 4 H 2 -[55]
2 F e O + H 2 O   F e 2 O 3 + H 2 E° = 0.356 (V)[55]
2 F e 3 O 4 + H 2 O   3 F e 2 O 3 + H 2     E° = 0.210 (V)[40]
2 F e 3 O 4 + 4 H 2 O   6 F e O O H + H 2     E° = 0.230 (V)[40]
Table 5. Temperatures reached in a thermochemical solar reactor with a fluidized bed.
Table 5. Temperatures reached in a thermochemical solar reactor with a fluidized bed.
Thermal PowerTemperatures Reached or RequiredThermal FluidFluidized Bed MaterialParticle Size (µm)Radiation or Concentration RatioRefs.
1 kWth1300 °CWater/SteamCoal-coke140
(200–300)
(300–500)
(500–710)
477 W/cm2[74,80,81,82]
10 to 20 kWth2200 °CWater/SteamZnO reagent powder1–5-[83]
3 MWth1500 °CWater/SteamCerium oxide material<300
(100–300)
1.5 kW/m2[38,77,84,85]
110 kWth1400 °CGasesNon-stoichiometric cerium oxide particles10–210-[79,86]
100 kWth960–1100 °CAirQuartz sand particles100–500-[87]
250 kWth800–1000 °CAirA mixture of coal-coke and quartz sand100-300 (coal)
100–700 (sand)
-[88]
30 kWth560 °CAirA mixture of sand and basalt -[89]
450 kWth770 °CAirIsotropic materials 4.8 kWh/m2/900 suns[90,91]
140.63 Wth
(2 kWe)
231.32 Wth
(4 kWe)
250 °CAirSand, ceramic casting media (carbo Accucast ID50) y SiC 65 kW/m2
(2kWe)
115 kW/m2
(4 kWe)
[92]
Table 6. Iron Oxide Materials.
Table 6. Iron Oxide Materials.
Iron Oxides MaterialSample/Particle Size (µm)Temp. (°C)
(Time)
Partial Vapour PressureThermal Fluid (Flow)Refs.
MagnetiteFe3O430–50
100–125
575
673
--[23,74]
The partial substitution of iron in Fe3O4 by Ni, Co, and Zr, to form mixed metal oxides.(Fe(1−x) Mx)3O4
NiFe2O4
NiFe2O4/m-ZrO2
Sample in a quartz tube1000--[23,38,74]
Magnetite supported on zirconium, stabilized with cubic yttriaFe3O4/c-YSZCeramic foam1100
(80 min)
75% of the steam pressure at 90 °C, 1 barH2O/N2
(10 Ncm3/min)
[86]
Commercial nickel ferrite supported on zirconium oxideNiFe2O4/ ZrO2Sample on Pt cup1000
(60 min)
steam pressure at 80 °C, 1 barH2O/N2
(4 mL/min)
[93]
Unsupported commercial nickel ferriteNiFe2O4212–710 1.6–1.7 kWth
(10–92 min)
51% (0.51 atm) of the steam pressure at 82 °C, 1 atmH2O/N2
(0.24 Ndm3/min)
[39]
Monoclinic magnetite supported on zirconium substratesFe3O4/m-ZrO2Polished Fe bar without oxidation397–602
(180 min)
-Nitrogen
(100 cc/min)
Argon
(200 cc/min)
Liquid water
(12.5 µL/min)
[67]
Iron oxide with various support materials, ZrO2, CeO2, yttria-stabilized zirconia (YSZ), and gadolinia-doped
ceria (GDC)
Fe2O3/ZrO2
Fe2O3/CeO2
Fe2O3/YSZ
Fe2O4/GDC
150–300550-H2O/Ar
(prop. 5:95)
(300 mL/min)
[70]
Table 7. Technologies of thermochemical solar reactors with fluidized beds.
Table 7. Technologies of thermochemical solar reactors with fluidized beds.
Reactor GeometryBed/GasMaterialPower/RadiationReactorYear/Refs.
Diameter = 5 cm
Height = 32 cm
Charcoal/CO2Silice glass tube2 kW/
400 W/cm2
Applsci 14 07056 i0011983
[94]
ZrO2,
SiC/air,
N2,
CO2
Clear quartz tube6.5 kW/
1000 W/m2
Applsci 14 07056 i0021983
[95]
0.90 m high, 0.78 m wide, with a radius of curvature of 0.78 mAlumina particles/airRefractory stainless steel (AISI 310)30–45 kW/
1 kW/m2
Applsci 14 07056 i0031988
[96]
Diameter = 2 cmZnO + Al2O3/
CH4 + Ar
Quartz tube15 kW/
4000 suns
Applsci 14 07056 i0041995
[97]
Diameter = 45 mm, thickness 2.5 mmNiFe2O4—ZrO2/N2Quartz tube6 kWApplsci 14 07056 i0052008
[98]
25 mm in diameter, thickness 1.5 mm, height 25 cm.CaO or CaCO3/H2O, Ar, CO2Quartz tube75 kW/
4250 suns
Applsci 14 07056 i0062009
[99]
420 mm in length, 62.3 mm in inner diameter, and 7 mm in thicknessCoke/CO2Stainless steel with a quartz window6 kW/
477 W/cm2
Applsci 14 07056 i0072010
[80]
The inner diameter was 45 mm, and the thickness was 2.5 mm.NiFe2O4/vapourStainless-steel tube with a quartz cap3 × 6 kW/
637 W/cm2
Applsci 14 07056 i0082011
[39]
Length 420 mm, inner diameter 62.3 mm, thickness 7 mmCoke/Ar-vapour
CeO2/N2
stainless-steel tube (SUS310S) with a quartz window18 kW/
637 W/cm2
Applsci 14 07056 i0092015
[81,84]
42 cm diameter at the topQuartz sand particles/airInconel and stainless steel, with a quartz window133 kW/
950 kW/m2
Applsci 14 07056 i0102016
[100]
0.2 m long and 0.3 m internal diameter at the top and 0.007 m thickQuartz particles/N2Stainless steel tube with quartz window (5 mm thick)21 kW/
903 kW/m2
Applsci 14 07056 i0112018
[101]
35.6 cm long and 2.54 cm outer diameterIron aluminate (FeAl2O4)/N2Silicon carbide cylindrical tubes (SiC)10 kWApplsci 14 07056 i0122019
[102]
Inside diameter of 7.62 cm and a height of 8 cmSand, carbon Accucast ID50 y SiC/airStainless-steel tube
(304)
2kWe/
65 kW/m2
4 kWe/
115 kW/m2
Applsci 14 07056 i0132020
[92]
Tube with 40 mm inner diameter, 10 mm wall thickness, and 1780 mm lengthQuartz sand/airPure iron metal tube, hot particle container made of stainless steel AISI 30410–40 kWeApplsci 14 07056 i0142022
[103]
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

Fuentes, M.; Pulido, D.; Fuentealba, E.; Soliz, A.; Toro, N.; Sagade, A.; Galleguillos Madrid, F.M. Direct Solar Thermal Water-Splitting Using Iron and Iron Oxides at High Temperatures: A Review. Appl. Sci. 2024, 14, 7056. https://doi.org/10.3390/app14167056

AMA Style

Fuentes M, Pulido D, Fuentealba E, Soliz A, Toro N, Sagade A, Galleguillos Madrid FM. Direct Solar Thermal Water-Splitting Using Iron and Iron Oxides at High Temperatures: A Review. Applied Sciences. 2024; 14(16):7056. https://doi.org/10.3390/app14167056

Chicago/Turabian Style

Fuentes, Manuel, Diego Pulido, Edward Fuentealba, Alvaro Soliz, Norman Toro, Atul Sagade, and Felipe M. Galleguillos Madrid. 2024. "Direct Solar Thermal Water-Splitting Using Iron and Iron Oxides at High Temperatures: A Review" Applied Sciences 14, no. 16: 7056. https://doi.org/10.3390/app14167056

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