Next Article in Journal
Numerical Investigation of Flow Boiling in Interconnected Microchannels at Varying Mass Fluxes
Previous Article in Journal
Toward Prediction of Energy Consumption Peaks and Timestamping in Commercial Supermarkets Using Deep Learning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Reduction of Iron Oxides for CO2 Capture Materials

by
Antonio Fabozzi
*,
Francesca Cerciello
and
Osvalda Senneca
*
National Research Council, Institute of Sciences and Technologies for Sustainable Energy and Mobility (CNR-STEMS), P. le V. Tecchio 80, 80125 Napoli, Italy
*
Authors to whom correspondence should be addressed.
Energies 2024, 17(7), 1673; https://doi.org/10.3390/en17071673
Submission received: 28 February 2024 / Revised: 22 March 2024 / Accepted: 28 March 2024 / Published: 1 April 2024
(This article belongs to the Section A: Sustainable Energy)

Abstract

:
The iron industry is the largest energy-consuming manufacturing sector in the world, emitting 4–5% of the total carbon dioxide (CO2). The development of iron-based systems for CO2 capture and storage could effectively contribute to reducing CO2 emissions. A wide set of different iron oxides, such as hematite (Fe2O3), magnetite (Fe3O4), and wüstite (Fe(1−y)O) could in fact be employed for CO2 capture at room temperature and pressure upon an investigation of their capturing properties. In order to achieve the most functional iron oxide form for CO2 capture, starting from Fe2O3, a reducing agent such as hydrogen (H2) or carbon monoxide (CO) can be employed. In this review, we present the state-of-the-art and recent advances on the different iron oxide materials employed, as well as on their reduction reactions with H2 and CO.

1. Introduction

In the last few decades, CO2 emissions have significantly increased, attracting the attention of the scientific community because these emissions are the main cause of the greenhouse effect [1,2]. The rise of the CO2 concentration in the atmosphere is due to deforestation [3] and highly energetic industries, particularly in the sectors of chemicals [4], power, and steel, largely sustained by fossil fuels [5,6]. The steel industry alone is responsible for ~7% of the total CO2 emission because of the wide use of fossil fuels [7,8].
Despite the increasing share of renewables for power generation and the acceleration towards the utilization of H2 as a combustible in the steel industry [9,10], the International Energy Agency (IEA) reports that currently more than 80% of the world’s energy is still based on fossil fuel combustion [11,12], and in a mid-term scenario all large scale/industrial plants should employ some type of equipment in order to reduce CO2 emissions [13,14]. For this reason, a wide set of energy companies are developing and testing strategies of carbon capture and sequestration (CCS) [15,16]
CCS [17,18] includes the separation, liquefaction, and finally storage of CO2. The CO2 can be stored (i) in deep ocean masses, (ii) in deep geological formations, or (iii) in the form of carbonates [19,20]. The most promising sequestration sites are the geological formations due to the possibility of storing ~2000 Gigatons of CO2 [21,22]. In particular, the injection of CO2 in oil reservoirs, mixed with different types of surfactants [23,24,25,26,27], is a validated technology to enhance oil recovery (EOR) [28,29]. Indeed, the combination based on the capture of CO2 and its employment in EOR technology effectively reduces the CCS costs. Unfortunately, nowadays there are no power plants present that can operate with complete CCS methods because of the high cost and energy penalty of the capture stage, especially for hot and dilute flows derived from power plants [30,31]. Recent advances in research are primarily aimed at reducing costs and increasing the efficiency and selectivity of CO2 separation and capture [32,33].
CO2 capture technologies are based on either chemical [34] or physical methods [35]. In physical methods, CO2 is adsorbed by a solid and is further released by decompression and/or heating under different conditions of temperature and pressure [36,37]. Contrariwise, the chemical methods are based on chemical reactions between CO2 and liquids, such as amine solutions, or reactive solid sorbents, such as CaO [38,39]. The current benchmark for CO2 capture in coal, oil, or natural gas power plants and other heavy industrial processes is absorption with amine solutions [40,41,42,43,44,45], while calcium looping capture is one step back in the Technology Readiness Level (TRL) [46]. Unfortunately, capture with amine solutions is very energy-intensive and expensive [47,48]; indeed, amine systems are capable of capturing ~85% of the CO2 from the exhaust gas of a fossil-fuel power plant but with a penalty of ~25% of the plant electricity production [49,50].
CO2 capture with membranes is a more recent solution but does not yet represent a mature alternative to amine capture. Indeed, their convenience is hampered by the large amount of exhaust gas that has to be treated [51,52]. The scientific community is, in fact, focusing on the enhancement of CO2-nitrogen (N2) membrane selectivity by using the facilitated mechanism membranes [53,54].
Another approach which has been recently gaining a prominent role is the use of metal oxides, such as iron oxides, magnesium oxides (MgO), and calcium oxides (CaO), for CO2 capture, via the formation of carbonates [55,56]. In the second step, carbonates can be heated up to release back pure CO2 gas and regenerate metal oxides [57,58]. Overall a cyclic process can be set, which includes the exothermic carbonation reaction and the endothermic carbonate decomposition [59,60].
In this review, we discuss the prospects of iron oxides in the context of energy transition, with a particular focus on their application for CCS. The features of the different types of iron oxides and their suitability for CO2 capture will be surveyed. The thermodynamic and kinetic aspects of iron oxides reduction with H2 or CO are also addressed.

2. Iron Oxides in the Context of Energy Transition and CCS

2.1. Motivation for the Use of Iron Oxides in CCS

The use of iron in the context of CCS is particularly appealing thanks to its large availability across the globe [61,62,63] and relative economy.
Iron is present in the planet mainly as hematite (Fe2O3), magnetite (Fe3O4), and wüstite (FeO). Other iron-based compounds include sulfides, hydroxides, and carbonates [64,65,66].
Iron oxides have been proposed both for pre-capture processes, such as chemical-looping combustion (CLC) [67,68,69], and for post combustion CO2 capture. CLC processes have been developed in order to increase the concentration of CO2 in the exhaust gases, thus reducing the costs of post combustion CO2 capture processes. The high concentration of CO2 is attained by performing carbon oxidation in a Fuel reactor, where the airflow is replaced by recycled flue gas. Here the oxygen necessary for carbon oxidation is released from the metal oxygen carriers, in particular iron oxides, which reduce to iron. The latter is, then, re-oxidized in a reactor, named the Air reactor, and recycled back to the Fuel reactor.
As far as the post combustion capture applications are concerned, iron has been proposed for the preparation of enhanced sorbents and used in combination with carbonaceous supports. However, the simplest and probably most direct and economic way to exploit iron oxides, in particular Fe2O3 and Fe3O4, is for capturing CO2 from exhaust gases through the formation of carbonates (siderite FeCO3). The carbonation stage can be followed by a high temperature decarbonation step to regenerate iron and release gaseous CO2. Such a solution appears particularly appealing for capturing CO2 from steel plants, where iron powders (even waste) are available on site and can be profitably exploited. Iron carbonates can be regenerated through the decarbonation reaction either in the same facility [70], or at locations far from the CO2 emission/capture sites.
In this context, another important advantage can be envisaged: siderites can be safely transported via trucks, allowing CO2 transport without the need of pipelines. However, in the techno-economic analysis, it must be considered that for the transport of 44 ton of CO2, approximately 116 ton of material (in the form of siderite) must be handled (for a ratio of 44/156 = 0.38 g/g, which can be compared with the values of 0.1–0.3 g/g reported for CO2 capture over metal organic frameworks (MOFs) [71]. Another issue that has to be considered for the economy of the process is that, in order to capture CO2, the divalent or trivalent oxides of iron must be previously reduced in the metallic state [72,73] by a reduction agent, such as H2 or CO. This step is crucial and might end up being the bottleneck of the process. This step will, indeed, be addressed in detail in the next paragraphs. Finally, the possibility of the utilization of iron materials over repeated cycles needs to be further investigated.

2.2. Hematite

Fe2O3 is one of the most important ores for industrial applications and represents 60%, of the iron ore reserve worldwide [74]. Oxygen (O2) carriers using Fe2O3 as an active phase [75] or natural Fe2O3 have been widely investigated in the context of CLC. Notably, Al2O3 and SiO2 in combination with Fe2O3, can further enhance the properties of O2 carriers in CLC, their durability and reactivity [76].
Some studies investigated the use of Fe2O3 in natural ores and in waste materials as oxygen carriers in the CLC of fossil fuels, including coal and biomass with gas. Indeed, Zhang et al. studied the use of natural Fe2O3 in the CLC of coal under pressurized conditions and their results found no sintering on the Fe2O3 surface and good reaction performance [77]. Gu et al. used natural Fe2O3 [78] in the CLC of sawdust at 800 °C. Song et al. investigated the natural Fe2O3 CLC combustion of ShenHua bituminous coal and HuaiBei anthracite [79].
Furthermore, Fe3O4 and Fe2O3 can be employed as sorbents for CO2 capture. Indeed, Mendoza et al. have investigated CO2 capture and release employing Fe2O3 and Fe3O4. In particular, the chemical reactions among CO2 and iron oxides were explored versus both the CO2 pressure and ball mill process parameters. The experimental results indicated that the carbonation of iron oxides can be achieved at 10–30 bar and at room temperature. Furthermore, the complete calcination of FeCO3, was investigated under Ar and vacuum atmospheres and it was found that at 367 °C FeCO3 decomposed, yielding carbon and or ion and Fe3O4. In particular, it was highlighted that this mixture can reversibly capture CO2 in several carbonation–calcination cycles [36].

2.3. Magnetite

Fe3O4 is an iron oxide widely employed as a component of industrial heterogeneous catalysts for its availability, stability, and low cost [80]. It has a spinel structure, and the mixed valence state, +2/+3, of the Fe can catalyze both acid–base and oxidation–reduction reactions [81]. For instance, in the Haber–Bosch process, Fe3O4 is part of the catalysts employed for ammonia production, in which Fe3O4 is reduced by H2 to the active form α-Fe, allowing the adsorption and dissociation of the N2 molecules [82].
In the last decade, in the CO2 capture scenario, Fe3O4 has also been used to produce amine-functionalized nanofluids with higher CO2 absorption rates than pure solvents [83]. Fe3O4 nanoparticles attracted scientific attention thanks to their low toxicity and high stability [84]. Park et al. have developed novel nano absorbents composed of Fe3O4 nanoparticles functionalized with different organic and inorganic materials such as (3-Aminopropyl) triethoxysilane (ATS), tetraethyl orthosilicate (TeOS), and L (+)-ascorbic acid (A). In order to synthesize an Fe3O4 double-functionalization, a primary coating with A followed by a secondary SiO2 shell and subsequently, the amine functionalization by means of ATS was applied. The CO2 absorption capacity of newly functionalized material, with A, was 11% higher than that of the un-functionalized ones, Fe3O4-SiO2-NH2. However, in order to determine the absorption improvement of functionalized material at 25 °C, concentrations ranging from 0.1 to 0.4 wt% were compared with those in which water was used as a reference. The highest CO2 absorption improvement of 59.2% was achieved at a concentration of 0.3 wt% which is higher than blank Fe3O4, Fe3O4-A-SiO2 at 0.3 wt%. Furthermore, a demonstration of cyclic performances was established in order to test its potential employment for industrial applications [85].
The observed enhancement of the CO2 absorption of iron-amine-based nanofluids has been attributed to improved mass transfer compared to conventional amines solutions, thanks to a combination of effects [86]: the hydrodynamic effect [87,88], the shuttle effect [86,89], and the bubble-breaking effect [90,91]. However, these nanofluids still have limitations for industrial applications because they are susceptible to acid and oxidation conditions.
Furthermore, recently, a mechanochemical process aiming at capturing CO2 by using Fe2O3 and Fe3O4 has been employed in the CO2 capture field; this process led to the formation of FeCO3 via carbonation reactions (1) and (2) [36].
F e 3 O 4 ( s ) + F e ( s ) + 4 C O 2 4 F e C O 3 ( s )
F e 2 O 3 ( s ) + F e ( s ) + 3 C O 2 3 F e C O 3 ( s )
In particular, Mendoza et al. [36], have investigated the ball milling condition effects (revolution speed, pressure, and reaction time) for both reactions (1) and (2) and found that the capturing capacity of the Fe2O3 and Fe system was higher than that of Fe3O4 and Fe at the same conditions of pressure, temperature, and reaction. In particular, the results demonstrated a capture capacity of 7.98 mmol CO2 g−1sorbent in 3 h at 200 °C and 10 bar CO2 pressure. Furthermore, the oxides could be completely regenerated by applying an oxidation step at 350 °C under airflow [15].

2.4. Wüstite

Wüstite is not commonly found as a natural material but is an intermediate derived from the reduction of iron ores, Fe2O3 and Fe3O4. The chemical formula of wüstite is Fe(1−y)O, where y is the relative number of the ferrous iron ion vacancies. The O2 content can change at different temperatures from a 23.16 wt% to 25.60 wt% range in variation. Fe(1−y)O is thermodynamically unstable and prone to re-oxidize at ambient conditions; however, it can be also further reduced to metallic iron using agents such as carbon (C), CO, and H2.
Recently O2 carriers composed of Fe(1−y)O mixed with other oxides, such as those of manganese (Mn) have been used in the chemical looping combustion of coal. It has been found that a fraction of O2 was transferred via an uncoupling mechanism [92]. Perèz-Vega et al. prepared particles starting from Mn3O4 and Fe2O3 powders with an atomic ratio fixed at Mn:Fe, 77:23 by a ball mill. Subsequently, the solids mixture was calcined for 4 h at 1050 °C, and the calcined solid was milled for 300 min. Furthermore, the material behavior was evaluated during both the decomposition–regeneration of the [(Mn0.77Fe0.23)2O3] phase, bixbyite, in chemical looping with O2 uncoupling, and the reduction–oxidation of [(Mn0.77Fe0.23)3O4], spinel phase, with CO, H2, and CH4. In particular, the chemical looping combustion redox cycle, using gaseous fuels, highlighted high reactivity with CO and H2 and high O2 transport capacity due to the reduction of [(Mn0.77Fe0.23)O], mangano-wüstite phase. Therefore, this material is considered a promising candidate to be employed in CLC with both coal and gas [93].
Furthermore, even if the wüstite is not stable, it has been used in combination with MgO in order to increase the CO2 capture capability of MgO sorbents. Chen et al. investigated the ferric salt effect on the adsorption of strong solid bases in order to determine the optimal FeO amount to increase the CO2 capture in flue-gas at T > 150 °C. They synthesized the FeO–MgO using magnesium acetate and ferrous acetate. In particular, after grinding and carbonization treatments of the two salts at 550 °C, a composite FeO-MgO is obtained in which the ferric species are surrounded by magnesia particles with a surface area higher than 160 m2 g−1. Furthermore, with a 3% mass ratio of FeO the basic sorbent can capture 35 mg g−1 CO2 at 150°. The CO2 capture capacity and the ratio of strong basic sites results are higher than those of other MgO sorbents, resulting in an optimal situation for warm CO2 capture [94].

3. Thermodynamics of Carbon Capture with Iron Oxides

In order to obtain low-cost iron products with high performances from natural iron ores, the reactions with reducing agents such as H2 and CO need to be investigated in detail. The kinetics and thermodynamics of these reduction reactions will be discussed in the following chapter.
The reduction of Fe2O3 to the metallic iron (Fe) occurs stepwise. At temperatures > 570 °C, reduction is composed of three steps (Fe2O3 → Fe3O4 → Fe(1−y)O → Fe) according to the Equation (3a–c). At temperatures < 570 °C, Fe2O3 reduces firstly to Fe3O4 and then to Fe (Fe2O3 → Fe3O4 → Fe), according to the two-step reaction reported in Equation (4a,b), while the Fe(1−y)O intermediate is unstable.
3 F e 2 O 3 + H 2 2 F e 3 O 4 + H 2 O T > 570   ° C
( 1 y ) F e 3 O 4 + ( 1 4 y ) H 2 3 F e ( 1 y ) O + ( 1 4 y ) H 2 O
F e ( 1 y ) O + H 2 ( 1 y ) F e + H 2 O
3 F e 2 O 3 + H 2 2 F e 3 O 4 + H 2 O , T < 570   ° C
F e 3 O 4 + 4 H 2 3 F e + 4 H 2 O ,
The stability of different iron oxide forms can be analyzed by the diagram composed of temperature versus O (wt%/100) reported in Figure 1.
The oxygen-to-iron content in the Fe-based systems scales in the following order: Fe2O3 > Fe3O4 > FeO. As previously mentioned, FeO is stable at T > 570 °C while at T < 570 °C Fe2O3 decomposes to Fe3O4 and Fe. However, the FeO stability area expands by increasing the temperature, because not all places in the lattice are occupied by iron ions. For this reason, the FeO formula has to be considered as Fe(1−y)O in which 1−y represents vacancies in the iron lattice [96].
The Baur–Glässner diagram based on the thermodynamics of iron ore reduction is shown in Figure 2.
In particular, the diagram describes the stability areas of the different iron oxide phases during the reaction with H2 (continuous three-phase line) and with CO (dashed three-phase line) as a function of gas oxidation degree (GOD) and temperature [95,97]. The GOD value is determined as the ratio of oxidized gas components over the sum of oxidized and oxidizable gas components (e.g., for a CO-CO2 mixture, GOD = CO/(CO + CO2)) and is an indicator of the reduction force of a gas mixture (the lower the GOD, the higher the reduction force of the gas mixture). It is possible to note from the diagram that when reduction is carried out with H2, the stability region of Fe increases with temperature and GOD. Contrariwise, when reduction is carried out with CO, Fe is never stable for GOD > 0.5 and the iron stability region decreases with temperature for GOD < 0.5.
Therefore, in order to achieve Fe with H2 the temperature of reduction must be lowered as the GOD decreases. Instead, in order to achieve Fe with CO, GOD must be kept above 0.5 and the temperature of reduction must be increased as the GOD decreases. For the best possible gas utilization, it is therefore recommended to operate at high temperatures with H2 and at low temperatures with CO.
For the Fe-C-O system, the Boudouard reaction, see Equation (5), must also be considered. For this reason, the corresponding equilibrium line (calculated assuming a carbon activity of 1 and a pressure of 1 bar) is also reported in the diagram.
2 C O C O 2 + C
The Boudouard equilibrium line delimitates the Baur–Glässner diagram into two sections. For a CO-CO2 gas mixture with composition and temperature below the equilibrium line, carbon precipitation can occur and hinder the reduction.
Furthermore, it is necessary to highlight that the reduction with CO is exothermic, while the reduction with H2 is endothermic; therefore, in this case energy must be added to the system in order to maintain a constant reaction temperature. From an industrial point of view, an adequate process design, that involves the pre-heating of the materials, could aid in overcoming issues due to the supply of energy to the system [95].

4. Morphology and Microstructural Properties of Iron Oxides

Iron oxides involved in industrial processes and research activities have different features such as particle size, porosity, and purity degree. The true density of iron oxides decreases in the order Fe > Fe(1−y)O > Fe2O3 > Fe3O4 (7.86 > 5.46 > 5.20 > 5.16 kg/m3 [98]. However, bulk or apparent density depends on porosity. In several industrial applications cm-size pellets or briquettes are used, compacting finer grains, from tens of micrometers to a few millimeters in size. The interstices between the grains generate large macroporosity from micro to even millimetric scale, while pores inside the fine grains contribute with smaller (micro and mesopore) range porosity.
Hakim et al. investigated the microstructure of fine powders of different iron oxides by SEM, shown in Figure 3. In particular, a honeycomb structure can be observed for FeO, a rhombohedral lattice for Fe2O3, and a clear cubical structure for Fe3O4. In terms of porosity, the following average pore diameters (dpore) and pore volumes (Vpore) are reported: dpore= 62.5 nm and Vpore = 0.0003 cm3/g for FeO; dpore = 17.5 nm and Vpore = 0.029 cm3/g for Fe2O3; and dpore = 15.5 nm and Vpore = 0.018 cm3/g for Fe3O4 [99].

4.1. Porosity Changes in Reduction of Iron Ores

The porosity and density of the iron ore materials are key parameters controlling the reduction kinetics of the iron ore materials [100]. An analysis of the cross sections of large and hard particles after partial reduction reveals that reduction follows a topochemical pattern and concentric layers of different oxides are established, composed of Fe2O3, Fe3O4, Fe(1−y)O and Fe, respectively. Concentric layers are not observed for porous particles, yet a topochemical type of reduction could still be possible at a smaller scale, within the micro-grains which constitute larger pellets, Figure 4.
In several papers, the mutual relationship between porosity and reduction rate or progress was addressed [96,101,102].
Ghadi et al. compared several pellets of iron ores and highlighted the porosity, grain size, and particle size effect and found that the rate of reduction increased linearly with the initial porosity degree of the samples, Figure 5 [103,104].
Sarkar et al. [100] reported, instead, for twenty different samples of Fe2O3 pellets (diameter 12 mm), a progressive decrease in porosity and increase in density upon reduction at 900 °C with CO and H2.
El-Geassy et al. studied the effects of temperature on the structure of Fe2O3 pellets upon reduction with H2 and reported different trends in the evolution of porosity, according to the compaction degree and grains particles size. Indeed, for the less porous/more compact pellets and larger grains size, high temperature and prolonged reduction treatments induce sintering, which reduces the reactivity of the samples [101]. On the other hand, for highly porous pellets and small grain sizes [105], sintering effects are less important, while the densification of the unreduced oxide grains results in an overall increase in porosity and reduction reactivity.

4.2. Effect of Gangue on Reduction Kinetics and Porosity

Iron ores also contain additional gangue, such as Al2O3, SiO2, CaO, and MgO, in different amounts, depending on the ore and the enrichment process. In general, gangue can have a prominent effect on the reduction rate of iron oxides and on sintering [101].
Teplov et al. investigated the rate of reduction with H2 of natural Fe3O4 concentrates in the range 300–570 °C and observed that stronger oxides, such as Al2O3, Ti2O3, V2O3, Mn3O4, and MgO, that form solid solutions with Fe3O4, decrease reduction rate, because they decrease the Gibbs thermodynamic potential and the equilibrium oxygen pressure over Fe3O4. Among these oxides, Al2O3 strongly decreases the rate of hydrogen reduction of Fe3O4, while MgO showed a weak effect [106].
Kapelyushin et al. investigated the effect of Al2O3 on the reactivity of iron ore reduction. In particular, they compared the reduction rate of undoped and doped Fe3O4 with CO-CO2 gas mixtures at different temperatures. As shown in Figure 5 of [107], an amount of alumina, 3%, enhances the reduction reactivity of iron ore, but for higher amounts of alumina the effect is reversed and reduction rate decreases [107,108].
Paananen et al. also investigated the effect of alumina on Fe3O4 during reduction in a CO/CO2 gas mixture at 850 °C and found that an Al-doped Fe3O4 sample reduces more rapidly than the undoped Fe3O4 sample [109]. They attributed this to swelling and crack formation, suggesting that, upon doping, Al cations diffuse into the Fe3O4 structure forming hercynite. Upon reduction, high tension forces in the Fe3O4-hercynite phase close to the interface with the FeO would cause the disruption of the structure and generation of creaks.
As far as Ca is concerned, different papers addressed the effects of lime as well as of calcium carbonate on the reducibility of iron ores, and showed that their effect is beneficial, resulting in an enhancement of porosity upon reduction [101,110]. On the contrary, MgO would have a detrimental effect on iron reducibility [111]. As far as SiO2 is concerned, controversial results have been reported in the literature; however, it seems that the effect of silica alone on the reduction reactivity of iron ores [112,113] can be considered of lower importance than that of other gangues, however SiO2 can be involved in the formation of several phases when it is combined with Fe and Ca, such as in fayalite and calcio-ferrite.

4.3. Microstructural Changes in CO2 Capture by Iron Oxides

The relation between microstructure and CO2 captures by Fe2O3, Fe3O4, and FeO has been widely investigated in the literature [96,99,114].
Hakim et al. [99] investigated the microstructure and morphology of different iron oxides upon reaction with CO2. In particular, they compared materials before and after CO2 exposure for 4, 24, and 48 h and observed differences in carbonate growth on the surface, as shown in Figure 6.
In particular, the formation of bicarbonate, monodentate carbonate, and bidentate carbonates, as well as the morphological changes of the iron particles were observed. For FeO, SEM micrographs highlighted the honeycomb structure that became smooth and groove-like upon exposure with CO2, suggesting the chemical interaction of the iron structure with several carbonate species. For Fe3O4 some sharp nanoparticles, indicative of carbonate formation, were clearly identified after exposure to CO2; however, the cubic lattice of the iron oxide was preserved. A similar, but more remarkable effect was observed for Fe3O4, where nanoparticles form larger aggregates. Among all iron oxides, the uptake of CO2 was consistently highest with Fe2O3 (at 3.01 mgCO2/g).

5. Kinetics of Iron Ore Reduction

Several papers reported kinetic studies on iron ores reduction in different atmospheres. The majority, in fact, addressed the kinetics of reduction with H2 and used thermogravimetric methods, either isothermal or non-isothermal. The results obtained are very scattered, as can be observed from the values of activation energies reported in Table 1, which indeed span between 57.1 (isothermal test) and 246 (non-isothermal test) kJ/mol. Kinetic models also spanned from simple first-order reactions to more complex models.
Notably, the reliability of kinetic expressions obtained by conventional (thermogravimetric) methods for any heterogenous (gas–solid) reaction relies on the fulfilment of some conditions:
1. Reactions result in a single stage of mass loss or in well-resolved and defined sequential stages.
2. The structure of the solid reactant is relatively constant throughout the experiment.
3. Inter- and intra-particle mass and temperature diffusion resistances are negligible.
It is evident that, due to the complexity of its reaction scheme, the reduction of iron by hydrogen does not fulfil the first condition.
The second condition is not fulfilled either, because the reduction of iron oxides leads to different metallic iron formations (Fe2O3, Fe3O4, and FeO).
As for the third condition, the mass transfer limitation (of the gaseous reactants) can be avoided by performing experiments at low temperature and with fine particles; however, in the reduction of iron ores, even for small particles size, an additional mechanism must be considered, that is, the transfer of electrons and iron ions generated by changes in iron oxides.
The complex interactions between physical and chemical phenomena and their impacts on the rate of the reaction of iron oxides with H2/CO will be better explained in the following paragraph.

6. Effect of Physical and Chemical Phenomena on the Reaction Rate

The rate of reduction of iron ore particles depends on a complex series of chemical and transport phenomena, as reported in Scheme 1.
The first step is the mass transfer of the reducing gas from the bulk of the gas phase to the particles’ outer surface. This is followed by diffusion within pores and chemical reaction (including adsorption on the iron oxide internal surface and phase boundary reaction). Chemical reactions inside the particles produce gaseous products (CO2 or H2O), which have to diffuse towards the particle’s outer surface, through particle pores. Electrons and iron ions generated by changes in iron oxides, instead, diffuse towards the particle’s inner core by means of a solid-state diffusion mechanism. Reaction fronts and even concentrical layers of solid reduction products can therefore be established inside the particle as already shown in Figure 4 [125].
Notably, solid-state diffusion, which involves interstitial sites and defects, is much slower than gas diffusion and strongly dependent on the form of iron oxide. Gas transport in porous particles, instead, occurs through molecular diffusion mechanisms in pores with diameters larger than the mean free path, and according to a Knudsen diffusion mechanism in pores with diameters smaller than the mean free path [126].
As reported in previous paragraphs, reduction and heat treatment can progressively alter the porosity and microstructure of the material and modify the controlling reaction pattern. Two opposite phenomena are possible: on one side, the formation of creaks upon reduction can generate new porosity, which facilitates gas diffusion, and on the other side, sintering and the formation of dense iron layers reduces porosity and slows down gas diffusion.
Generally, in the literature, at least for fine particles [127,128,129,130], the first stage of reduction, Fe2O3 → Fe3O4, is relatively fast and limited by gas diffusion, while the second stage of the reaction, whereby Fe3O4 is reduced to Fe(1−y)O or Fe, is slower and parameters such as temperature, concentration, particle size, and porosity all contribute to determine the rate-limiting step [123,124]. In particular, Table 2 reports a selection of experimental works that highlighted the rate-limiting process in the reduction of Fe3O4 to Fe(1−y)O or Fe for different particles typology (grains or pellets), size, morphology, and porosity, and for different process parameters such as temperature, pressure, and gas composition. Results are also summarized in Figure 7a, for the reaction with H2 and H2/CO mixtures, and in Figure 7b, for CO and CO/H2 mixtures.
At high temperatures and for large particles size or large porosity, the rate-limiting step is generally external mass transfer. On the contrary, for low temperatures and fine particles size, intrinsic kinetics/phase boundary reaction controls the overall rate [140,141,142]. In intermediate cases, the controlling rate mechanism could be either gas or solid-state diffusion and even mixed regimes are possible where, for example, transport and phase boundary reaction have comparable relevance [106,132].
Table 2. The effect of different parameters on iron oxide reduction.
Table 2. The effect of different parameters on iron oxide reduction.
Reducing AgentParticle SizePorosity (ε)T (°C)P (bar)Ea (kJ/mol)Limitation Steps
(2nd Reaction Step)
Reference
H24 mm × 4 mm × 8 mm-900–1100-99.2Solid-state diffusion[138]
H2, CO125 to 500 μm0.15350–6001091Phase boundary reaction[125]
CO100–150 μm-700–850-80Nucleation of FeO and then internal diffusion[136]
H2<74 µm0.54500–11006851 for T < 650 °C;
84 for 650–900 °C;
176 for T > 900 °C
T < 650 °C chemical reaction,
650–900 °C chemical reaction + solid state diffusion,
T > 900 °C solid state diffusion
[105]
CO-CO2
H2
50.8 µm0.0115240–417-71Phase boundary reaction[134]
H2/CO/mixtures1.07–1.24 cm0.06850--Mixed (chemical reaction + pore diffusion) at the beginning, pore diffusion at the end[143]
H2/CO--800–1000-48.64 with CO
63.19 with H2
First phase boundary reaction, then pore diffusion and mixed regime (at 800 °C and 900 °C).[144]
H275–180 µm0.27700–10000.25–133Pore diffusion[123]
H2/CO--150–900--Solid-state diffusion[135]
H2/CO1.5–4.4 cm0.31700–12001–2-Gas diffusion
At 700 °C in the late stages solid state diffusion
[133]
H2
CO/CO2
14 mm-900--Pore diffusion[101]
H2/CO12 mm-800–1000--Chemical reaction at the beginning, pore diffusion at the end[137]
H250–160 µm0.54001-Chemical reaction[106]
H2100–160 µm0.55001-External gas transport + chemical reaction[106]
H2pellets0.5300–5001-Transport in the external H2O layer[106]
CO-CO2~150 µm0.69590–10001-T < 700 °C, mixed control in the pores
T > 700 °C external gas transport
[130]
CO-CO2150–500 µm0.42590–10001-Chemical reaction[130]
CO-CO2<75 µm0.23590–10001-Chemical reaction[130]
H20.249 µm-400–570 131.5 Fe3O4 → FeO
76.0 FeO → Fe
Chemical reaction[131]
H2
CO
<100 µm-800–1100-41.15 with H2
54.19 with CO
Phase boundary reaction[145]
Notably it has also been reported that the rate-controlling step can change throughout reduction because of changes in the microstructure of the particles [133,136,143,144]. Kawasaki et al. performed the reduction of porous Fe2O3 spherical pellets of a few centimeters in size in streams of pure H2 or CO at temperatures between 700° and 1200 °C. At the higher temperature, the reduction rate was entirely diffusion controlled and a shell of reduction was clearly visible in sections of partially reduced specimens, which moved concentrically into the core of the samples. However at the lower reduction temperatures, Fe(1−y)O was trapped inside dense shells of iron impervious to the reducing gas and the reaction proceeded by solid-state diffusion [133].
Variations in the controlling regime were also observed for finer particles. Chen et al. studied the reduction kinetics of 100–150 μm Fe2O3 by CO between 700–850 °C and concluded that the initial stage the reduction process was controlled by the gas–solid reaction occurring at the Fe/FeO interface, but subsequently it was controlled by the nucleation of FeO, and finally by diffusion. Moreover, the reactions of Fe2O3 → FeO and FeO → Fe occurred simultaneously but with different time dependences [136].
The changes in reaction regimes throughout the reduction of iron ores can be explained considering that the true density of the different oxides decreases in the order Fe > Fe(1−y)O > Fe2O3 > Fe3O4 and that denser iron oxide forms imply slower diffusion [146,147,148,149].
Thermodynamic considerations, summarized in the Baur–Glässner diagram, shown in Figure 2, indicate that transition from Fe3O4 to the denser Fe and Fe(1−y)O occurs above a critical temperature, whose value, however, varies according to the reduction potential of the gaseous atmosphere (which is the inverse of the GOD reported in Figure 2). In CO/CO2 mixtures, a transition occurs at a temperature of ~570 °C for GOD above 50%, while it occurs at 1000 °C above 30% of GOD transition. In H2/H2O mixtures for GOD above 20%, the transition occurs above ~570 °C. As far as the differences between H2 and CO as reducing agents are concerned, it can be observed that H2 and CO have the same reduction potential at ~810 °C, while at higher temperatures the reduction potential of H2 is higher than that of CO, while the CO’s reduction potential is higher than H2′s at low temperatures.
Several authors investigated different gas composition effects on the reduction rate of iron ores [104,150]. Turkdogan et al. highlighted that the reduction rate of iron ore pellets in H2/CO mixtures increased with H2 concentration, likely due to its better gas diffusivity [101]. Also, El-Geassy et al. found that for the reduction of FeO micropellets between 900 and 1100 °C, the rate increased with the percentage of H2 in the gas mixture, and they observed a neat formation of Fe nuclei. On the contrary, when reduction was carried out with pure CO, they detected a delay in reduction resulting in an incubation time [151]. These results confirm that the thermodynamically driven changes of iron forms end up also having an impact on the reaction rate, because of the different densities of the iron forms that determine different diffusional resistances which might ultimately control the overall reaction rate.
In addition, the diffusivity of the reducing agent molecule should be considered.
Indeed, Zuo et al. investigated the reduction rate in a H2/CO atmosphere at 800 °C, where the thermodynamic transition between more or less dense iron forms is comparable for H2 and CO, see Figure 3 of [137]. Even under these conditions, they observed that the reduction rate is higher with H2, thanks to the better diffusivity of the H2 molecule. They also conclude that increasing the CO amount, even by a small amount, in a gas mixture hinders the diffusion path of H2 leading to an inhibition of the chemical reaction [137]. The decrease in the rate when CO is present in the reducing gas was also observed by El Geassy et al. who attributed it to a poisoning effect of CO molecules on the Fe(1−y)O surface [152].

7. Conclusions, Challenges and Future Perspectives

The use of iron-based materials for CO2 capture is still under exploration; however, research executed in the last five years has shown that iron oxide materials could facilitate CO2 capture thanks to their utilization directly in industrial plants. An interesting route for CO2 capture, especially in the context of steel industry, is the reaction of Fe2O3 or Fe3O4 with reduced iron and with CO2 to form FeCO3. The CO2, captured in the form of FeCO3, can be easily transported by truck, and on a decarbonation stage, can free the CO2 and regenerate iron.
While the main advantages of this route are clearly the large availability and relative cheapness of iron across the globe, a successful application of natural iron ores requires a careful analysis and comprehension of the relationships between their structural properties (density, porosity, etc.) and process conditions (temperature, gas composition). The pre-reduction step, necessary to reduce iron ores to metallic iron, may, in this respect, play a crucial role.
Luckily, a great deal of information on the microstructural properties of the different iron oxide forms and on the kinetics and thermodynamics of their reduction can be obtained from the large amount of literature which has been produced over the last decades in the context of steel making. The following general indications can be pointed out:
  • H2, for thermodynamic and kinetic reasons, is a better reduction agent than CO thanks to better diffusion behavior. Indeed, the viscosity and the molecule size of H2 are lower than CO, affecting the diffusion behavior.
  • The first step of reduction, leading from Fe2O3 to Fe3O4, below 570 °C, is relatively fast compared to further reduction from Fe3O4 to Fe(1−y)O or Fe; however, it is important because it opens up porosity.
The rate of reduction of Fe3O4 to Fe(1−y)O or Fe can be limited by structural features of the starting material and by microstructural changes occurring upon reduction and heat treatment. The rate-controlling step at high temperatures and for large particle size or large porosity is generally external gas transport. On the contrary, for low temperatures and finer particles size, solid-state diffusion or phase boundary reaction can hamper the overall rate.
It is clear from the current state-of-the-art that the best conditions for the production of iron-based materials, which are not only efficient in capturing CO2, but that can also be re-used over several cycles, cannot be superficially selected and require a careful analysis of the material properties, like grain size, porosity, mineralogy, and the presence of gangue. Further research activity in this field is warmly encouraged to provide guidelines for the utilization of natural iron ores in CO2 capture, as well as to unlock the potential of advances/engineered iron-based materials for CCS.

Author Contributions

A.F.: Conceptualization, Writing—original draft, Writing—review & editing. F.C.: Writing—original draft. O.S.: Conceptualization, Writing—original draft, Writing—review and editing, Supervision, and Funding Acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

Antonio Fabozzi acknowledges funding from the European Union—NextGenerationEU under the National Recovery and Resilience Plan (PNRR), Mission 04 Component 2 Investment 3.1, and Project “ECCSELLENT—Development of ECCSEL-R.I. ItaLian facilities: usEr access, services and loNg-Term sustainability” Code: IR0000020—CUP F53C22000560006.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Abbreviations

AbbreviationName
CO2Carbon dioxide
Fe2O3Hematite
Fe3O4Magnetite
Fe(1−y)OWüstite
IEAInternational Energy Agency
H2Hydrogen
COCarbon oxide
CCSCarbon Capture and Sequestration
EOREnhance Oil Recovery
CaOCalcium oxides
TRLTechnology Readiness Level
N2Nitrogen
MgOMagnesium oxides
CLCChemical Looping Combustion
FeCO3Siderite
MOFsMetal organic frameworks
O2Oxygen
Al2O3Alumina
SiO2Silica
AlAluminium
FeIron
ATS(3-Aminopropyl) triethoxysilane
TeOSTetraethyl orthosilicate
AL (+)-ascorbic acid
CCarbon
MnManganese
Mn3O4Hausmannite
CH4Methane
GODGas Oxidation Degree
SEMScanning electron microscopy
dporePore diameter
VporePore volume
Ti2O3Tistarite
V2O3Karelianite
H2OWater

References

  1. Yoro, K.O.; Daramola, M.O. CO2 emission sources, greenhouse gases, and the global warming effect. In Advances in Carbon Capture; Elsevier: Amsterdam, The Netherlands, 2020; pp. 3–28. [Google Scholar]
  2. Jeffry, L.; Ong, M.Y.; Nomanbhay, S.; Mofijur, M.; Mubashir, M.; Show, P.L. Greenhouse gases utilization: A review. Fuel 2021, 301, 121017. [Google Scholar] [CrossRef]
  3. Joshi, R.; Singh, H. Carbon sequestration potential of disturbed and non-disturbed forest ecosystem: A tool for mitigating climate change. Afr. J. Environ. Sci. Technol. 2020, 14, 385–393. [Google Scholar]
  4. Nica, A.; Popescu, A.; Ibanescu, D.-C. Human influence on the climate system. Curr. Trends Nat. Sci. 2019, 8, 209–215. [Google Scholar]
  5. Saleh, T.A. Nanomaterials and hybrid nanocomposites for CO2 capture and utilization: Environmental and energy sustainability. RSC Adv. 2022, 12, 23869–23888. [Google Scholar] [CrossRef] [PubMed]
  6. Wani, O.A.; Kumar, S.S.; Hussain, N.; Wani, A.I.A.; Subhash, B.; Parvej, A.; Rashid, M.; Popescu, S.M.; Mansoor, S. Multi-scale processes influencing global carbon storage and land-carbon-climate nexus: A critical review. Pedosphere 2023, 33, 250–267. [Google Scholar] [CrossRef]
  7. Gajdzik, B.; Sroka, W.; Vveinhardt, J. Energy Intensity of Steel Manufactured Utilising EAF Technology as a Function of Investments Made: The Case of the Steel Industry in Poland. Energies 2021, 14, 5152. [Google Scholar] [CrossRef]
  8. Kim, J.; Sovacool, B.K.; Bazilian, M.; Griffiths, S.; Lee, J.; Yang, M.; Lee, J. Decarbonizing the iron and steel industry: A systematic review of sociotechnical systems, technological innovations, and policy options. Energy Res. Soc. Sci. 2022, 89, 102565. [Google Scholar] [CrossRef]
  9. Pareek, A.; Dom, R.; Gupta, J.; Chandran, J.; Adepu, V.; Borse, P.H. Insights into renewable hydrogen energy: Recent advances and prospects. Mater. Sci. Energy Technol. 2020, 3, 319–327. [Google Scholar] [CrossRef]
  10. Amin, M.; Shah, H.H.; Fareed, A.G.; Khan, W.U.; Chung, E.; Zia, A.; Farooqi, Z.U.R.; Lee, C. Hydrogen production through renewable and non-renewable energy processes and their impact on climate change. Int. J. Hydrogen Energy 2022, 47, 33112–33134. [Google Scholar] [CrossRef]
  11. Berdysheva, S.; Ikonnikova, S. The energy transition and shifts in fossil fuel use: The study of international energy trade and energy security dynamics. Energies 2021, 14, 5396. [Google Scholar] [CrossRef]
  12. Kober, T.; Schiffer, H.-W.; Densing, M.; Panos, E. Global energy perspectives to 2060–WEC’s World Energy Scenarios 2019. Energy Strategy Rev. 2020, 31, 100523. [Google Scholar] [CrossRef]
  13. Griffin, P.W.; Hammond, G.P. Industrial energy use and carbon emissions reduction in the iron and steel sector: A UK perspective. Appl. Energy 2019, 249, 109–125. [Google Scholar] [CrossRef]
  14. Holappa, L. A general vision for reduction of energy consumption and CO2 emissions from the steel industry. Metals 2020, 10, 1117. [Google Scholar] [CrossRef]
  15. Budinis, S.; Krevor, S.; Mac Dowell, N.; Brandon, N.; Hawkes, A. An assessment of CCS costs, barriers and potential. Energy Strategy Rev. 2018, 22, 61–81. [Google Scholar] [CrossRef]
  16. Van der Spek, M.; Roussanaly, S.; Rubin, E.S. Best practices and recent advances in CCS cost engineering and economic analysis. Int. J. Greenh. Gas Control. 2019, 83, 91–104. [Google Scholar] [CrossRef]
  17. Jiang, K.; Ashworth, P. The development of Carbon Capture Utilization and Storage (CCUS) research in China: A bibliometric perspective. Renew. Sustain. Energy Rev. 2021, 138, 110521. [Google Scholar] [CrossRef]
  18. Hussin, F.; Aroua, M.K. Recent trends in the development of adsorption technologies for carbon dioxide capture: A brief literature and patent reviews (2014–2018). J. Clean. Prod. 2020, 253, 119707. [Google Scholar] [CrossRef]
  19. Snæbjörnsdóttir, S.Ó.; Sigfússon, B.; Marieni, C.; Goldberg, D.; Gislason, S.R.; Oelkers, E.H. Carbon dioxide storage through mineral carbonation. Nat. Rev. Earth Environ. 2020, 1, 90–102. [Google Scholar] [CrossRef]
  20. Kelemen, P.; Benson, S.M.; Pilorgé, H.; Psarras, P.; Wilcox, J. An overview of the status and challenges of CO2 storage in minerals and geological formations. Front. Clim. 2019, 1, 9. [Google Scholar] [CrossRef]
  21. Ajayi, T.; Gomes, J.S.; Bera, A. A review of CO2 storage in geological formations emphasizing modeling, monitoring and capacity estimation approaches. Pet. Sci. 2019, 16, 1028–1063. [Google Scholar] [CrossRef]
  22. Herzog, H.J. Scaling up carbon dioxide capture and storage: From megatons to gigatons. Energy Econ. 2011, 33, 597–604. [Google Scholar] [CrossRef]
  23. Fabozzi, A.; Della Sala, F.; di Gennaro, M.; Solimando, N.; Pagliuca, M.; Borzacchiello, A. Polymer based nanoparticles for biomedical applications by microfluidic techniques: From design to biological evaluation. Polym. Chem. 2021, 12, 6667–6687. [Google Scholar] [CrossRef]
  24. Fabozzi, A.; Vitiello, R.; Krauss, I.R.; Iuliano, M.; De Tommaso, G.; Amoresano, A.; Pinto, G.; Paduano, L.; Jones, C.; Di Serio, M.; et al. Synthesis, Surface Properties, and Self-Aggregation Behavior of a Branched N,N-Dimethylalkylamine Oxide Surfactant. J. Surfactants Deterg. 2019, 22, 115–124. [Google Scholar] [CrossRef]
  25. Fabozzi, A.; Krauss, I.R.; Vitiello, R.; Fornasier, M.; Sicignano, L.; King, S.; Guido, S.; Jones, C.; Paduano, L.; Murgia, S.; et al. Branched alkyldimethylamine oxide surfactants: An effective strategy for the design of high concentration/low viscosity surfactant formulations. J. Colloid Interface Sci. 2019, 552, 448–463. [Google Scholar] [CrossRef]
  26. Scermino, L.; Fabozzi, A.; De Tommaso, G.; Valente, A.J.M.; Iuliano, M.; Paduano, L.; D’Errico, G. pH-responsive micellization of an amine oxide surfactant with branched hydrophobic tail. J. Mol. Liq. 2020, 316, 113799. [Google Scholar] [CrossRef]
  27. Savignano, L.; Fabozzi, A.; Vitiello, R.; Fornasier, M.; Murgia, S.; Guido, S.; Guida, V.; Paduano, L.; D’Errico, G. Effect of tail branching on the phase behavior and the rheological properties of amine oxide/ethoxysulfate surfactant mixtures. Colloids Surf. A-Physicochem. Eng. Asp. 2021, 613, 126091. [Google Scholar]
  28. Massarweh, O.; Abushaikha, A.S. A review of recent developments in CO2 mobility control in enhanced oil recovery. Petroleum 2022, 8, 291–317. [Google Scholar] [CrossRef]
  29. Wang, L.; Tian, Y.; Yu, X.; Wang, C.; Yao, B.; Wang, S.; Winterfeld, P.H.; Wang, X.; Yang, Z.; Wang, Y. Advances in improved/enhanced oil recovery technologies for tight and shale reservoirs. Fuel 2017, 210, 425–445. [Google Scholar] [CrossRef]
  30. Vega, F.; Baena-Moreno, F.; Fernández, L.M.G.; Portillo, E.; Navarrete, B.; Zhang, Z. Current status of CO2 chemical absorption research applied to CCS: Towards full deployment at industrial scale. Appl. Energy 2020, 260, 114313. [Google Scholar] [CrossRef]
  31. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A. Carbon capture and storage (CCS): The way forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar]
  32. Oschatz, M.; Antonietti, M. A search for selectivity to enable CO2 capture with porous adsorbents. Energy Environ. Sci. 2018, 11, 57–70. [Google Scholar] [CrossRef]
  33. Odunlami, O.; Vershima, D.; Oladimeji, T.; Nkongho, S.; Ogunlade, S.; Fakinle, B. Advanced techniques for the capturing and separation of CO2—A review. Results Eng. 2022, 15, 100512. [Google Scholar] [CrossRef]
  34. Salvi, B.L.; Jindal, S. Recent developments and challenges ahead in carbon capture and sequestration technologies. SN Appl. Sci. 2019, 1, 885. [Google Scholar] [CrossRef]
  35. Madejski, P.; Chmiel, K.; Subramanian, N.; Kuś, T. Methods and techniques for CO2 capture: Review of potential solutions and applications in modern energy technologies. Energies 2022, 15, 887. [Google Scholar] [CrossRef]
  36. Mendoza, E.Y.M.; Santos, A.S.; López, E.V.; Drozd, V.; Durygin, A.; Chen, J.; Saxena, S.K. Iron oxides as efficient sorbents for CO2 capture. J. Mater. Res. Technol. 2019, 8, 2944–2956. [Google Scholar] [CrossRef]
  37. Ramirez-Vidal, P.; Sdanghi, G.; Celzard, A.; Fierro, V. High hydrogen release by cryo-adsorption and compression on porous materials. Int. J. Hydrogen Energy 2022, 47, 8892–8915. [Google Scholar] [CrossRef]
  38. Nicotera, I.; Policicchio, A.; Conte, G.; Agostino, R.G.; Rehman, M.H.U.; Lufrano, E.; Simari, C. Quaternized polyepichlorohydrin-based membrane as high-selective CO2 sorbent for cost-effective carbon capture. J. CO2 Util. 2022, 63, 102135. [Google Scholar] [CrossRef]
  39. Tao, H.; Qian, X.; Zhou, Y.; Cheng, H. Research progress of clay minerals in carbon dioxide capture. Renew. Sustain. Energy Rev. 2022, 164, 112536. [Google Scholar] [CrossRef]
  40. Powell, C.E.; Qiao, G.G. Polymeric CO2/N2 gas separation membranes for the capture of carbon dioxide from power plant flue gases. J. Membr. Sci. 2006, 279, 1–49. [Google Scholar] [CrossRef]
  41. Barbera, E.; Mio, A.; Pavan, A.M.; Bertucco, A.; Fermeglia, M. Fuelling power plants by natural gas: An analysis of energy efficiency, economical aspects and environmental footprint based on detailed process simulation of the whole carbon capture and storage system. Energy Convers. Manag. 2022, 252, 115072. [Google Scholar] [CrossRef]
  42. Sanchez Moore, C.C.; Kulay, L. Effect of the implementation of carbon capture systems on the environmental, energy and economic performance of the Brazilian electricity matrix. Energies 2019, 12, 331. [Google Scholar] [CrossRef]
  43. Dutcher, B.; Fan, M.; Russell, A.G. Amine-based CO2 capture technology development from the beginning of 2013 A Review. ACS Appl. Mater. Interfaces 2015, 7, 2137–2148. [Google Scholar] [CrossRef] [PubMed]
  44. Osman, A.I.; Hefny, M.; Abdel Maksoud, M.; Elgarahy, A.M.; Rooney, D.W. Recent advances in carbon capture storage and utilisation technologies: A review. Environ. Chem. Lett. 2021, 19, 797–849. [Google Scholar] [CrossRef]
  45. Yamada, H. Amine-based capture of CO2 for utilization and storage. Polym. J. 2021, 53, 93–102. [Google Scholar] [CrossRef]
  46. Haaf, M.; Anantharaman, R.; Roussanaly, S.; Ströhle, J.; Epple, B. CO2 capture from waste-to-energy plants: Techno-economic assessment of novel integration concepts of calcium looping technology. Resour. Conserv. Recycl. 2020, 162, 104973. [Google Scholar] [CrossRef]
  47. Aghel, B.; Janati, S.; Wongwises, S.; Shadloo, M.S. Review on CO2 capture by blended amine solutions. Int. J. Greenh. Gas Control 2022, 119, 103715. [Google Scholar] [CrossRef]
  48. Ochedi, F.O.; Yu, J.; Yu, H.; Liu, Y.; Hussain, A. Carbon dioxide capture using liquid absorption methods: A review. Environ. Chem. Lett. 2021, 19, 77–109. [Google Scholar] [CrossRef]
  49. Gür, T.M. Carbon dioxide emissions, capture, storage and utilization: Review of materials, processes and technologies. Prog. Energy Combust. Sci. 2022, 89, 100965. [Google Scholar] [CrossRef]
  50. Siegelman, R.L.; Milner, P.J.; Kim, E.J.; Weston, S.C.; Long, J.R. Challenges and opportunities for adsorption-based CO2 capture from natural gas combined cycle emissions. Energy Environ. Sci. 2019, 12, 2161–2173. [Google Scholar] [CrossRef]
  51. Raganati, F.; Miccio, F.; Ammendola, P. Adsorption of carbon dioxide for post-combustion capture: A review. Energy Fuels 2021, 35, 12845–12868. [Google Scholar] [CrossRef]
  52. Sifat, N.S.; Haseli, Y. A critical review of CO2 capture technologies and prospects for clean power generation. Energies 2019, 12, 4143. [Google Scholar] [CrossRef]
  53. Ramli, N.A.; Hashim, N.A.; Aroua, M.K. Prediction of CO2/O2 absorption selectivity using supported ionic liquid membranes (SILMs) for gas–liquid membrane contactor. Chem. Eng. Commun. 2018, 205, 295–310. [Google Scholar] [CrossRef]
  54. da Silva Freitas, W.; Mecheri, B.; Vecchio, C.L.; Gatto, I.; Baglio, V.; Ficca, V.C.; Patra, A.; Placidi, E.; D’Epifanio, A. Metal-organic-framework-derived electrocatalysts for alkaline polymer electrolyte fuel cells. J. Power Sources 2022, 550, 232135. [Google Scholar] [CrossRef]
  55. Ruhaimi, A.; Aziz, M.; Jalil, A. Magnesium oxide-based adsorbents for carbon dioxide capture: Current progress and future opportunities. J. CO2 Util. 2021, 43, 101357. [Google Scholar] [CrossRef]
  56. Dunstan, M.T.; Donat, F.; Bork, A.H.; Grey, C.P.; Müller, C.R. CO2 capture at medium to high temperature using solid oxide-based sorbents: Fundamental aspects, mechanistic insights, and recent advances. Chem. Rev. 2021, 121, 12681–12745. [Google Scholar] [CrossRef] [PubMed]
  57. Sun, H.; Wu, C.; Shen, B.; Zhang, X.; Zhang, Y.; Huang, J. Progress in the development and application of CaO-based adsorbents for CO2 capture—A review. Mater. Today Sustain. 2018, 1, 1–27. [Google Scholar]
  58. Abd, A.A.; Othman, M.R.; Kim, J. A review on application of activated carbons for carbon dioxide capture: Present performance, preparation, and surface modification for further improvement. Environ. Sci. Pollut. Res. 2021, 28, 43329–43364. [Google Scholar] [CrossRef] [PubMed]
  59. André, L.; Abanades, S. Recent advances in thermochemical energy storage via solid–gas reversible reactions at high temperature. Energies 2020, 13, 5859. [Google Scholar] [CrossRef]
  60. Jin, B.; Wei, K.; Ouyang, T.; Fan, Y.; Zhao, H.; Zhang, H.; Liang, Z. Chemical looping CO2 capture and in-situ conversion: Fundamentals, process configurations, bifunctional materials, and reaction mechanisms. Appl. Energy Combust. Sci. 2023, 16, 100218. [Google Scholar] [CrossRef]
  61. Goyal, P.; Tiwary, C.S.; Misra, S.K. Ion exchange based approach for rapid and selective Pb (II) removal using iron oxide decorated metal organic framework hybrid. J. Environ. Manag. 2021, 277, 111469. [Google Scholar] [CrossRef]
  62. Zhang, X.; Jiao, K.; Zhang, J.; Guo, Z. A review on low carbon emissions projects of steel industry in the World. J. Clean. Prod. 2021, 306, 127259. [Google Scholar] [CrossRef]
  63. Chai, W.S.; Cheun, J.Y.; Kumar, P.S.; Mubashir, M.; Majeed, Z.; Banat, F.; Ho, S.-H.; Show, P.L. A review on conventional and novel materials towards heavy metal adsorption in wastewater treatment application. J. Clean. Prod. 2021, 296, 126589. [Google Scholar] [CrossRef]
  64. Taghizadeh, S.-M.; Berenjian, A.; Zare, M.; Ebrahiminezhad, A. New perspectives on iron-based nanostructures. Processes 2020, 8, 1128. [Google Scholar] [CrossRef]
  65. Hamza, A.; Hussein, I.A.; Jalab, R.; Saad, M.; Mahmoud, M. Review of iron sulfide scale removal and inhibition in oil and gas wells: Current status and perspectives. Energy Fuels 2021, 35, 14401–14421. [Google Scholar] [CrossRef]
  66. Garibello, C.F.; Eldridge, D.S.; Malherbe, F.; Hocking, R.K. Abiotic transformations of nitrogen mediated by iron sulfides and related species from early Earth to catalyst design. Inorg. Chem. Front. 2023, 10, 6792–6811. [Google Scholar] [CrossRef]
  67. Ubando, A.T.; Chen, W.H.; Show, P.L.; Ong, H.C. Kinetic and thermodynamic analysis of iron oxide reduction by graphite for CO2 mitigation in chemical-looping combustion. Int. J. Energy Res. 2020, 44, 3865–3882. [Google Scholar] [CrossRef]
  68. Abuelgasim, S.; Wang, W.; Abdalazeez, A. A brief review for chemical looping combustion as a promising CO2 capture technology: Fundamentals and progress. Sci. Total Environ. 2021, 764, 142892. [Google Scholar] [CrossRef] [PubMed]
  69. Raganati, F.; Alfe, M.; Gargiulo, V.; Chirone, R.; Ammendola, P. Kinetic study and breakthrough analysis of the hybrid physical/chemical CO2 adsorption/desorption behavior of a magnetite-based sorbent. Chem. Eng. J. 2019, 372, 526–535. [Google Scholar] [CrossRef]
  70. Manchisi, J.; Matinde, E.; Rowson, N.A.; Simmons, M.J.; Simate, G.S.; Ndlovu, S.; Mwewa, B. Ironmaking and steelmaking slags as sustainable adsorbents for industrial effluents and wastewater treatment: A critical review of properties, performance, challenges and opportunities. Sustainability 2020, 12, 2118. [Google Scholar] [CrossRef]
  71. Gaikwad, S.; Kim, Y.; Gaikwad, R.; Han, S. Enhanced CO2 capture capacity of amine-functionalized MOF-177 metal organic framework. J. Environ. Chem. Eng. 2021, 9, 105523. [Google Scholar] [CrossRef]
  72. Hussein, A.; Burra, K.; Bassioni, G.; Hammouda, R.; Gupta, A. Production of CO from CO2 over mixed-metal oxides derived from layered-double-hydroxides. Appl. Energy 2019, 235, 1183–1191. [Google Scholar] [CrossRef]
  73. Chandrasekaran, S.; Bowen, C.; Zhang, P.; Li, Z.; Yuan, Q.; Ren, X.; Deng, L. Spinel photocatalysts for environmental remediation, hydrogen generation, CO2 reduction and photoelectrochemical water splitting. J. Mater. Chem. A 2018, 6, 11078–11104. [Google Scholar] [CrossRef]
  74. Nikolaeva, N.V.; Aleksandrova, T.N.; Chanturiya, E.L.; Afanasova, A. Mineral and technological features of magnetite–hematite ores and their influence on the choice of processing technology. ACS Omega 2021, 6, 9077–9085. [Google Scholar] [CrossRef] [PubMed]
  75. Cabello, A.; Abad, A.; García-Labiano, F.; Gayán, P.; De Diego, L.; Adánez, J. Kinetic determination of a highly reactive impregnated Fe2O3/Al2O3 oxygen carrier for use in gas-fueled Chemical Looping Combustion. Chem. Eng. J. 2014, 258, 265–280. [Google Scholar] [CrossRef]
  76. Enoch, K.; Sundaram, A.; Ponraj, S.S.; Sathya, P.; George, S.D.B.; Kumar, M.R. Enhancement of MXene optical properties towards medical applications via metal oxide incorporation. Nanoscale 2023, 15, 16874–16889. [Google Scholar] [CrossRef]
  77. Zhang, S.; Saha, C.; Yang, Y.; Bhattacharya, S.; Xiao, R. Use of Fe2O3-containing industrial wastes as the oxygen carrier for chemical-looping combustion of coal: Effects of pressure and cycles. Energy Fuels 2011, 25, 4357–4366. [Google Scholar] [CrossRef]
  78. Gu, H.; Shen, L.; Xiao, J.; Zhang, S.; Song, T. Chemical looping combustion of biomass/coal with natural iron ore as oxygen carrier in a continuous reactor. Energy Fuels 2011, 25, 446–455. [Google Scholar] [CrossRef]
  79. Song, T.; Shen, T.; Shen, L.; Xiao, J.; Gu, H.; Zhang, S. Evaluation of hematite oxygen carrier in chemical-looping combustion of coal. Fuel 2013, 104, 244–252. [Google Scholar] [CrossRef]
  80. Pawar, A.A.; Bandal, H.A.; Kim, H. Spinel type Fe3O4 polyhedron supported on nickel foam as an electrocatalyst for water oxidation reaction. J. Alloys Compd. 2021, 863, 158742. [Google Scholar] [CrossRef]
  81. Santos-Carballal, D.; Roldan, A.; Dzade, N.Y.; De Leeuw, N.H. Reactivity of CO2 on the surfaces of magnetite (Fe3O4), greigite (Fe3S4) and mackinawite (FeS). Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2018, 376, 20170065. [Google Scholar] [CrossRef]
  82. Zhu, H.; Ren, X.; Yang, X.; Liang, X.; Liu, A.; Wu, G. Fe-based catalysts for nitrogen reduction toward ammonia electrosynthesis under ambient conditions. SusMat 2022, 2, 214–242. [Google Scholar] [CrossRef]
  83. Raza, S.; Orooji, Y.; Ghasali, E.; Hayat, A.; Karimi-Maleh, H.; Lin, H. Engineering approaches for CO2 converting to biomass coupled with nanobiomaterials as biomediated towards circular bioeconomy. J. CO2 Util. 2023, 67, 102295. [Google Scholar] [CrossRef]
  84. Shan, J.; Wang, L.; Yu, H.; Ji, J.; Amer, W.; Chen, Y.; Jing, G.; Khalid, H.; Akram, M.; Abbasi, N. Recent progress in Fe3O4 based magnetic nanoparticles: From synthesis to application. Mater. Sci. Technol. 2016, 32, 602–614. [Google Scholar] [CrossRef]
  85. Park, J.H.; Lee, J.W.; Ahn, H.; Kang, Y.T. Development of novel nanoabsorbents by amine functionalization of Fe3O4 with intermediate ascorbic acid coating for CO2 capture enhancement. J. CO2 Util. 2022, 65, 102228. [Google Scholar] [CrossRef]
  86. Yu, W.; Wang, T.; Park, A.-H.A.; Fang, M. Review of liquid nano-absorbents for enhanced CO2 capture. Nanoscale 2019, 11, 17137–17156. [Google Scholar] [CrossRef] [PubMed]
  87. Kluytmans, J.; Van Wachem, B.; Kuster, B.; Schouten, J. Mass transfer in sparged and stirred reactors: Influence of carbon particles and electrolyte. Chem. Eng. Sci. 2003, 58, 4719–4728. [Google Scholar] [CrossRef]
  88. Li, L.; Kang, Y.T. Enhancement mechanisms of mass transfer performance by nanoabsorbents during CO2 absorption process. Int. J. Heat Mass Transf. 2021, 164, 120444. [Google Scholar] [CrossRef]
  89. Kars, R.; Best, R.; Drinkenburg, A. The sorption of propane in slurries of active carbon in water. Chem. Eng. J. 1979, 17, 201–210. [Google Scholar] [CrossRef]
  90. Kim, W.-g.; Kang, H.U.; Jung, K.-m.; Kim, S.H. Synthesis of silica nanofluid and application to CO2 absorption. Sep. Sci. Technol. 2008, 43, 3036–3055. [Google Scholar] [CrossRef]
  91. Li, L.; Kang, Y.T. Effects of bubble coalescence and breakup on CO2 absorption performance in nanoabsorbents. J. CO2 Util. 2020, 39, 101170. [Google Scholar] [CrossRef]
  92. Abad, A.; de Las Obras-Loscertales, M.; García-Labiano, F.; de Diego, L.; Gayán, P.; Adánez, J. In situ gasification Chemical-Looping Combustion of coal using limestone as oxygen carrier precursor and sulphur sorbent. Chem. Eng. J. 2017, 310, 226–239. [Google Scholar] [CrossRef]
  93. Pérez-Vega, R.; Abad, A.; Gayán, P.; de Diego, L.; García-Labiano, F.; Adánez, J. Development of (Mn0.77Fe0.23)2O3 particles as an oxygen carrier for coal combustion with CO2 capture via in-situ gasification chemical looping combustion (iG-CLC) aided by oxygen uncoupling (CLOU). Fuel Process. Technol. 2017, 164, 69–79. [Google Scholar] [CrossRef]
  94. Chen, J.L.; Dong, X.Y.M.; Shi, C.L.; Li, S.H.; Wang, Y.; Zhu, J.H. Fabrication of strong solid base FeO–MgO for warm CO2 capture. CLEAN–Soil Air Water 2019, 47, 1800447. [Google Scholar] [CrossRef]
  95. Spreitzer, D.; Schenk, J. Reduction of iron oxides with hydrogen—A review. Steel Res. Int. 2019, 90, 1900108. [Google Scholar] [CrossRef]
  96. Kim, S.-H.; Zhang, X.; Ma, Y.; Souza Filho, I.R.; Schweinar, K.; Angenendt, K.; Vogel, D.; Stephenson, L.T.; El-Zoka, A.A.; Mianroodi, J.R. Influence of microstructure and atomic-scale chemistry on the direct reduction of iron ore with hydrogen at 700 C. Acta Mater. 2021, 212, 116933. [Google Scholar] [CrossRef]
  97. Cavaliere, P. Hydrogen Ironmaking. In Hydrogen Assisted Direct Reduction of Iron Oxides; Springer: Berlin/Heidelberg, Germany, 2022; pp. 131–183. [Google Scholar]
  98. Akiyama, T.; Ohta, H.; Takahashi, R.; Waseda, Y.; Yagi, J.-I. Measurement and modeling of thermal conductivity for dense iron oxide and porous iron ore agglomerates in stepwise reduction. ISIJ Int. 1992, 32, 829–837. [Google Scholar] [CrossRef]
  99. Hakim, A.; Marliza, T.S.; Abu Tahari, N.M.; Wan Isahak, R.W.; Yusop, R.M.; Mohamed Hisham, W.M.; Yarmo, A.M. Studies on CO2 adsorption and desorption properties from various types of iron oxides (FeO, Fe2O3, and Fe3O4). Ind. Eng. Chem. Res. 2016, 55, 7888–7897. [Google Scholar] [CrossRef]
  100. Sarkar, A.; Chavan, V.; Pai, N.; Prakash, A.; Hazra, B.; Raut, P.; Sunilkumar, D.; Sivananda, C.; Kundu, S.; Nag, S. Reduction of Iron Ore Pellets: A Microstructural Perspective? Metall. Mater. Trans. A 2023, 55, 537–549. [Google Scholar] [CrossRef]
  101. Forsmo, S.; Forsmo, S.-E.; Samskog, P.-O.; Björkman, B. Mechanisms in oxidation and sintering of magnetite iron ore green pellets. Powder Technol. 2008, 183, 247–259. [Google Scholar] [CrossRef]
  102. Zare Ghadi, A.; Valipour, M.S.; Vahedi, S.M.; Sohn, H.Y. A review on the modeling of gaseous reduction of iron oxide pellets. Steel Res. Int. 2020, 91, 1900270. [Google Scholar] [CrossRef]
  103. Turkdogan, E.; Vinters, J. Reducibility of iron ore pellets and effect of additions. Can. Metall. Q. 1973, 12, 9–21. [Google Scholar] [CrossRef]
  104. Hayes, P. The kinetics of formation of H2O and CO2 during iron oxide reduction. Metall. Trans. B 1979, 10, 211–217. [Google Scholar] [CrossRef]
  105. Joseph, T. Porosity, reducibility and size preparation of iron ores. Trans. AIME 1936, 120, 72–89. [Google Scholar]
  106. El-Geassy, A.; Nasr, M. Influence of the original structure on the kinetics of hydrogen reduction of hematite compacts. Trans. Iron Steel Inst. Jpn. 1988, 28, 650–658. [Google Scholar] [CrossRef]
  107. Teplov, O. Kinetics of the low-temperature hydrogen reduction of magnetite concentrates. Russ. Metall. (Met.) 2012, 2012, 8–21. [Google Scholar] [CrossRef]
  108. Kapelyushin, Y.; Xing, X.; Zhang, J.; Jeong, S.; Sasaki, Y.; Ostrovski, O. Effect of alumina on the gaseous reduction of magnetite in CO/CO2 gas mixtures. Metall. Mater. Trans. B 2015, 46, 1175–1185. [Google Scholar] [CrossRef]
  109. Kapelyushin, Y.; Sasaki, Y.; Zhang, J.; Jeong, S.; Ostrovski, O. Formation of a network structure in the gaseous reduction of magnetite doped with alumina. Metall. Mater. Trans. B 2017, 48, 889–899. [Google Scholar] [CrossRef]
  110. Paananen, T.; Heinänen, K.; Härkki, J. Degradation of iron oxide caused by alumina during reduction from magnetite. ISIJ Int. 2003, 43, 597–605. [Google Scholar] [CrossRef]
  111. Pal, J. Innovative development on agglomeration of iron ore fines and iron oxide wastes. Miner. Process. Extr. Metall. Rev. 2018, 40, 248–264. [Google Scholar] [CrossRef]
  112. Cores, A.; Babich, A.; Muñiz, M.; Ferreira, S.; Mochon, J. The influence of different iron ores mixtures composition on the quality of sinter. ISIJ Int. 2010, 50, 1089–1098. [Google Scholar] [CrossRef]
  113. Hsieh, L.-H.; JA, W. Effect of oxygen potential on mineral formation in lime-fluxed iron ore sinter. ISIJ Int. 1989, 29, 625–634. [Google Scholar] [CrossRef]
  114. Mazanek, E.; Wyderko, M. Zur Optimierung der Eigenschaften von Eisenerzsintern. Arch. Für Das Eisenhüttenwesen 1976, 47, 457–463. [Google Scholar] [CrossRef]
  115. Li, J.; Zhang, H.; Gao, Z.; Fu, J.; Ao, W.; Dai, J. CO2 capture with chemical looping combustion of gaseous fuels: An overview. Energy Fuels 2017, 31, 3475–3524. [Google Scholar] [CrossRef]
  116. Munteanu, G.; Ilieva, L.; Andreeva, D. TPR data regarding the effect of sulfur on the reducibility of α-Fe2O3. Thermochim. Acta 1999, 329, 157–162. [Google Scholar] [CrossRef]
  117. Munteanu, G.; Ilieva, L.; Andreeva, D. Kinetic parameters obtained from TPR data for α-Fe2O3 and Auα-Fe2O3 systems. Thermochim. Acta 1997, 291, 171–177. [Google Scholar] [CrossRef]
  118. Lin, H.-Y.; Chen, Y.-W.; Li, C. The mechanism of reduction of iron oxide by hydrogen. Thermochim. Acta 2003, 400, 61–67. [Google Scholar] [CrossRef]
  119. Jozwiak, W.; Kaczmarek, E.; Maniecki, T.; Ignaczak, W.; Maniukiewicz, W. Reduction behavior of iron oxides in hydrogen and carbon monoxide atmospheres. Appl. Catal. A Gen. 2007, 326, 17–27. [Google Scholar] [CrossRef]
  120. Sastri, M.; Viswanath, R.; Viswanathan, B. Studies on the reduction of iron oxide with hydrogen. Int. J. Hydrogen Energy 1982, 7, 951–955. [Google Scholar] [CrossRef]
  121. Abd Elhamid, M.; Khader, M.; Mahgoub, A.; El Anadouli, B.; Ateya, B. Autocatalytic reduction of hematite with hydrogen under conditions of surface control: A vacancy-based mechanism. J. Solid State Chem. 1996, 123, 249–254. [Google Scholar] [CrossRef]
  122. Barde, A.A.; Klausner, J.F.; Mei, R. Solid state reaction kinetics of iron oxide reduction using hydrogen as a reducing agent. Int. J. Hydrogen Energy 2016, 41, 10103–10119. [Google Scholar] [CrossRef]
  123. Lee, G.-Y.; Choi, J.-P.; Song, J.-I.; Jung, S.-S.; Lee, J.-S. The kinetics of isothermal hydrogen reduction of nanocrystalline Fe2O3 powder. Mater. Trans. 2014, 55, 1611–1617. [Google Scholar] [CrossRef]
  124. Kuila, S.K.; Chaudhuri, S.; Chatterjee, R.; Ghosh, D. In Reduction of magnetite ore fines with hydrogen. In Proceedings of the 4th International Conference on Chemical Engineering, Dhaka, Bangladesh, 29–30 December 2014; Chemical Engineering Department, BUET: Dhaka, Bangladesh, 2014; p. 81. [Google Scholar]
  125. Kuila, S.K.; Chatterjee, R.; Ghosh, D. Kinetics of hydrogen reduction of magnetite ore fines. Int. J. Hydrogen Energy 2016, 41, 9256–9266. [Google Scholar] [CrossRef]
  126. Feilmayr, C.; Thurnhofer, A.; Winter, F.; Mali, H.; Schenk, J. Reduction behavior of hematite to magnetite under fluidized bed conditions. ISIJ Int. 2004, 44, 1125–1133. [Google Scholar] [CrossRef]
  127. Swimm, K.; Reichenauer, G.; Vidi, S.; Ebert, H.-P. Gas pressure dependence of the heat transport in porous solids with pores smaller than 10 μm. Int. J. Thermophys. 2009, 30, 1329–1342. [Google Scholar] [CrossRef]
  128. Du, Z.; Liu, J.; Liu, F.; Pan, F. Relationship of particle size, reaction and sticking behavior of iron ore fines toward efficient fluidized bed reduction. Chem. Eng. J. 2022, 447, 137588. [Google Scholar] [CrossRef]
  129. Komatina, M.; Gudenau, H.W. The sticking problem during direct reduction of fine iron ore in the fluidized bed. Metall. Mater. Eng. 2018. [Google Scholar] [CrossRef] [PubMed]
  130. Chung, U.-C.; Lee, I.-O.; Kim, H.-G.; Sahajwalla, V.; Chung, W.-B. Degradation characteristics of iron ore fines of a wide size distribution in fluidized-bed reduction. ISIJ Int. 1998, 38, 943–952. [Google Scholar] [CrossRef]
  131. Corbari, R.; Fruehan, R. Reduction of iron oxide fines to wustite with CO/CO2 gas of low reducing potential. Metall. Mater. Trans. B 2010, 41, 318–329. [Google Scholar] [CrossRef]
  132. Gudenau, H. Materialsammlung Zum Praktikum Metallurgie; Trans-Aix-Press: Aachen, Germany, 2002. [Google Scholar]
  133. Qu, Y.; Xing, L.; Shao, L.; Luo, Y.; Zou, Z. Microstructural characterization and gas-solid reduction kinetics of iron ore fines at high temperature. Powder Technol. 2019, 355, 26–36. [Google Scholar] [CrossRef]
  134. Wolfinger, T.; Spreitzer, D.; Zheng, H.; Schenk, J. Influence of a prior oxidation on the reduction behavior of magnetite iron ore ultra-fines using hydrogen. Metall. Mater. Trans. B 2022, 53, 14–28. [Google Scholar] [CrossRef]
  135. Sato, K.; Ueda, Y.; Nishikawa, Y.; Goto, T. Effect of Pressure on Reduction Rate of Iron Ore with High Pressure Fluidized Bed. Trans. Iron Steel Inst. Jpn. 1986, 26, 697–703. [Google Scholar] [CrossRef]
  136. Bahgat, M.; Khedr, M. Reduction kinetics, magnetic behavior and morphological changes during reduction of magnetite single crystal. Mater. Sci. Eng. B 2007, 138, 251–258. [Google Scholar] [CrossRef]
  137. Chen, H.; Zheng, Z.; Chen, Z.; Yu, W.; Yue, J. Multistep reduction kinetics of fine iron ore with carbon monoxide in a micro fluidized bed reaction analyzer. Metall. Mater. Trans. B 2017, 48, 841–852. [Google Scholar] [CrossRef]
  138. El-Rahaiby, S.; Rao, Y. The kinetics of reduction of iron oxides at moderate temperatures. Metall. Trans. B 1979, 10, 257–269. [Google Scholar] [CrossRef]
  139. Bonalde, A.; Henriquez, A.; Manrique, M. Kinetic analysis of the iron oxide reduction using hydrogen-carbon monoxide mixtures as reducing agent. ISIJ Int. 2005, 45, 1255–1260. [Google Scholar] [CrossRef]
  140. Liu, D.; Wang, X.; Zhang, J.; Liu, Z.; Jiao, K.; Liu, X.; Wang, R. Study on the controlling steps and reduction kinetics of iron oxide briquettes with CO-H2 mixtures. Metall. Res. Technol. 2017, 114, 611. [Google Scholar] [CrossRef]
  141. Tahari, M.N.A.; Salleh, F.; Saharuddin, T.S.T.; Samsuri, A.; Samidin, S.; Yarmo, M.A. Influence of hydrogen and carbon monoxide on reduction behavior of iron oxide at high temperature: Effect on reduction gas concentrations. Int. J. Hydrogen Energy 2021, 46, 24791–24805. [Google Scholar] [CrossRef]
  142. Kawasaki, E.; Sanscrainte, J.; Walsh, T.J. Kinetics of reduction of iron oxide with carbon monoxide and hydrogen. AIChE J. 1962, 8, 48–52. [Google Scholar] [CrossRef]
  143. Zuo, H.-B.; Wang, C.; Dong, J.-J.; Jiao, K.-X.; Xu, R.-S. Reduction kinetics of iron oxide pellets with H2 and CO mixtures. Int. J. Miner. Metall. Mater. 2015, 22, 688–696. [Google Scholar] [CrossRef]
  144. Baolin, H.; Zhang, H.; Hongzhong, L.; Qingshan, Z. Study on kinetics of iron oxide reduction by hydrogen. Chin. J. Chem. Eng. 2012, 20, 10–17. [Google Scholar]
  145. Hammam, A.; Li, Y.; Nie, H.; Zan, L.; Ding, W.; Ge, Y.; Li, M.; Omran, M.; Yu, Y. Isothermal and non-isothermal reduction behaviors of iron ore compacts in pure hydrogen atmosphere and kinetic analysis. Min. Metall. Explor. 2021, 38, 81–93. [Google Scholar] [CrossRef]
  146. Moradmand, S.; Allen, J. Magnetic carbon formation via in-situ CO2 capture and electrolysis in a molten carbonate system. Mater. Today Sustain. 2024, 25, 100645. [Google Scholar] [CrossRef]
  147. Ouyang, T.; Jin, B.; Mao, Y.; Wei, D.; Liang, Z. Control of strong electronic oxide-support interaction in iron-based redox catalysts for highly efficient chemical looping CO2 conversion. Appl. Catal. B Environ. 2024, 343, 123531. [Google Scholar] [CrossRef]
  148. Rao, Q.; Zhang, J.; Yang, T.; Li, Y.; Gai, Z.; Li, P.; Wang, X.; Pan, Y.; Jin, H. A nickel-modified perovskite-supported iron oxide oxygen carrier for chemical looping dry reforming of methane for syngas production. Chem. Eng. J. 2024, 485, 150033. [Google Scholar] [CrossRef]
  149. Saqline, S.; Wang, H.; Fan, Q.; Donat, F.; Müller, C.; Liu, W. Investigation of barium iron oxides for CO2 capture and chemical looping oxygen uncoupling. Appl. Energy Combust. Sci. 2024, 17, 100238. [Google Scholar] [CrossRef]
  150. Guo, D.; Hu, M.; Pu, C.; Xiao, B.; Hu, Z.; Liu, S.; Wang, X.; Zhu, X. Kinetics and mechanisms of direct reduction of iron ore-biomass composite pellets with hydrogen gas. Int. J. Hydrogen Energy 2015, 40, 4733–4740. [Google Scholar] [CrossRef]
  151. El-Geassy, A.A.; Rajakumar, V. Gaseous reduction of wustite with H2, CO and H2-CO mixtures. Trans. Iron Steel Inst. Jpn. 1985, 25, 449–458. [Google Scholar] [CrossRef]
  152. El-Geassy, A.-H.A. In Rate Controlling Step in the Reduction of Iron Oxides; Kinetics and Mechanism of Wüstite-Iron Step in H2, CO and H2/CO Gas Mixtures; IOP Conference Series: Materials Science and Engineering, 2017; IOP Publishing: Bristol, UK, 2017; p. 012002. [Google Scholar]
Figure 1. Binary system of Fe-O in which the effect of temperature versus O wt%/100 is reported. Reprinted from [95].
Figure 1. Binary system of Fe-O in which the effect of temperature versus O wt%/100 is reported. Reprinted from [95].
Energies 17 01673 g001
Figure 2. Baur–Glässner diagram for the Fe-O-H2 and Fe-O-C system including the Boudouard equilibrium for 1 bar and a carbon activity of 1. Reprinted from [95].
Figure 2. Baur–Glässner diagram for the Fe-O-H2 and Fe-O-C system including the Boudouard equilibrium for 1 bar and a carbon activity of 1. Reprinted from [95].
Energies 17 01673 g002
Figure 3. SEM micrographs of (i) FeO, (ii) Fe2O3, and (iii) Fe3O4. Reprinted from [99].
Figure 3. SEM micrographs of (i) FeO, (ii) Fe2O3, and (iii) Fe3O4. Reprinted from [99].
Energies 17 01673 g003
Figure 4. Representation of topochemical reduction of an iron ore particle.
Figure 4. Representation of topochemical reduction of an iron ore particle.
Energies 17 01673 g004
Figure 5. Pellet porosity effect on the rate of reduction for different iron ore samples. Reprinted from [104].
Figure 5. Pellet porosity effect on the rate of reduction for different iron ore samples. Reprinted from [104].
Energies 17 01673 g005
Figure 6. SEM micrographs for iron oxides treated with CO2: (a) for 4 h; (b) for 24 h; and (c) for 48 h of Fe2O3, and Fe3O4. Reprinted from [99].
Figure 6. SEM micrographs for iron oxides treated with CO2: (a) for 4 h; (b) for 24 h; and (c) for 48 h of Fe2O3, and Fe3O4. Reprinted from [99].
Energies 17 01673 g006
Scheme 1. Physical and chemical phenomena contributing to the reduction rate of porous iron ore particles.
Scheme 1. Physical and chemical phenomena contributing to the reduction rate of porous iron ore particles.
Energies 17 01673 sch001
Figure 7. Experimental works on rate-limiting step in reduction of Fe3O4 to Fe(1−y)O or Fe for different parameters: (a) reaction with H2 and H2/CO mixtures [101,105,106,123,125,131,132,133,134,135,136,137] and (b) reaction with CO and CO/H2 mixtures [101,125,130,131,133,134,135,138,139].
Figure 7. Experimental works on rate-limiting step in reduction of Fe3O4 to Fe(1−y)O or Fe for different parameters: (a) reaction with H2 and H2/CO mixtures [101,105,106,123,125,131,132,133,134,135,136,137] and (b) reaction with CO and CO/H2 mixtures [101,125,130,131,133,134,135,138,139].
Energies 17 01673 g007aEnergies 17 01673 g007b
Table 1. Reduction of iron oxides by hydrogen with their amounts of apparent activation energy [95].
Table 1. Reduction of iron oxides by hydrogen with their amounts of apparent activation energy [95].
Operative ConditionsReduction ReactionEa (kJ/mol)Reference
Non-Isothermal with H2Fe2O3 → Fe3O4246[115]
Fe3O4 → Fe93.2
Fe2O3 → Fe3O4162.1
Fe3O4 → Fe103.6
Fe2O3 → Fe3O4139.2[116]
Fe3O4 → FeO77.3
FeO → Fe85.7
Fe2O3 → Fe3O489.1[117]
Fe3O4 → Fe70.4
FeO → Fe104.0[118]
Isothermal with H2Fe2O3 → Fe57.1[119]
Fe2O3 → Fe 72.7
Fe2O3 → Fe 89.9
Fe2O3 → Fe3O430.1[120]
Fe2O3 → FeO47.0[121]
FeO → Fe30.0
Fe2O3 → Fe47.2[122]
Fe2O3 → Fe51.5
Fe2O3 → FeO42.0[123]
FeO → Fe55.0
Fe2O3 → FeO33.0[124]
FeO → Fe11.0
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

Fabozzi, A.; Cerciello, F.; Senneca, O. Reduction of Iron Oxides for CO2 Capture Materials. Energies 2024, 17, 1673. https://doi.org/10.3390/en17071673

AMA Style

Fabozzi A, Cerciello F, Senneca O. Reduction of Iron Oxides for CO2 Capture Materials. Energies. 2024; 17(7):1673. https://doi.org/10.3390/en17071673

Chicago/Turabian Style

Fabozzi, Antonio, Francesca Cerciello, and Osvalda Senneca. 2024. "Reduction of Iron Oxides for CO2 Capture Materials" Energies 17, no. 7: 1673. https://doi.org/10.3390/en17071673

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