Next Article in Journal
Dynamic Mechanical Properties and Mesoscopic Characteristics of Cemented Tailings Backfill Under Cyclic Dynamic Loading
Previous Article in Journal
Comprehensive Utilization Beneficiation Process of Lithium Pegmatite Ore: A Pilot-Scale Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unified Phase Diagram and Competition-Coupling Mechanism for Pyrite Thermal Transformation

by
Mingrui Liu
,
Guangyuan Xie
and
Jie Sha
*
Key Laboratory of Coal Processing and Efficient Utilization of Ministry of Education, School of Chemical Engineering and Technology, China University of Mining and Technology, Xuzhou 221116, China
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(11), 1139; https://doi.org/10.3390/min15111139
Submission received: 25 September 2025 / Revised: 21 October 2025 / Accepted: 28 October 2025 / Published: 30 October 2025

Abstract

The thermal transformation mechanism of pyrite in coal, which governs sulfur emissions and ash deposition, remains highly controversial. There are significant discrepancies in reported activation energies (Ea) (60–310 kJ/mol) and conflicting reaction pathways. To resolve these long-standing controversies, this study proposes a competition-coupling mechanism: pyrolysis and oxidation compete under local O2 and temperature gradients, while coupling through microstructural evolution. Specifically, pyrolysis generates a porous Fe1−XS that facilitates oxidation, which in turn can form a passivating oxide/sulfate layer that promotes further pyrolysis. This mechanism reconciles longstanding kinetic controversies by showing that the apparent activation energy is not a fixed value but instead a dynamic parameter, shifting along a continuous curve that bridges pyrolysis and oxidation-dominated regimes. Furthermore, we construct a unified phase diagram by incorporating the competition-coupling mechanism into classical thermodynamic equilibria. This diagram uses the molar ratio FeS2/(FeS2 + O2) and temperature to categorize the transformation process into four distinct regions—pyrolysis-dominated, competition-coupling, oxidation-dominated, and melt-dominated. The key contribution of this work lies in the diagram which offers a practical framework for optimizing combustion and roasting systems, allowing for improved control over sulfur emissions and ash-related issues such as slagging and fouling.

1. Introduction

Pyrite (FeS2), the most abundant sulfide mineral in coal, governs the formation of sulfur and iron oxides during thermal transformation. These products are primary contributors to air pollution, boiler corrosion, as well as ash slagging and deposition [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15]. Therefore, a fundamental and predictive understanding of the FeS2 transformation mechanism is essential for advancing clean coal utilization technologies.
Although considerable research has been conducted on pyrolysis and oxidation over the past several decades [7,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33], most studies have traditionally analyzed them as isolated processes, failing to reveal their complex interactions under non-equilibrium conditions. Consequently, the dominant reaction pathways and kinetic parameters remain contentious, as reflected in the wide range of reported apparent activation energies (60–310 kJ/mol).
To resolve these controversies fundamentally, we isolate the behavior of pyrite from the complexities of the coal matrix and then (1) reevaluate the thermodynamic boundaries of FeS2 transformation, (2) reconcile the discrepant kinetic data by identifying reaction mechanisms, (3) propose a unified competition-coupling mechanism, and (4) construct an original phase diagram.
By mapping the competition-coupling mechanism onto a modified thermodynamic diagram, this work yields a unified and predictive framework for FeS2 transformation, which has direct implications for controlling sulfur emissions and ash issues in coal utilization technologies.

2. Re-Evaluating the Thermodynamic Boundaries

2.1. Analysis of Reported Phase Equilibria

Reported thermodynamic phase diagrams for pyrite transformation exhibit considerable diversity, as they are derived from different subsystems (e.g., Fe-O, Fe-S, or Fe-S-O) and rely on different state variables (e.g., temperature, composition, or gas partial pressures). Through an integrative analysis, we establish three governing principles for phase stability in the Fe-S-O system:
1.
Collaborative Control of Temperature and Oxygen Partial Pressure (pO2):
Within the Fe-O subsystem, the work of Darken and Gurry (1946) [34] established the sequence of stable phases transitioning from Fe2O3→Fe3O4→FeO with increasing temperature and decreasing pO2.
In the more complex Fe-S-O system, the works of Jorgensen and Moyle (1982) [17] and Zhang et al. (2019) [35] further demonstrate that sulfates (FeSO4 and Fe2(SO4)3) emerge as key metastable phases at lower temperatures and under high pSO2 conditions.
Furthermore, within the FeS-FeO and FeO-Fe2O3 subsystems, McLennan et al. [28] demonstrated the following pathway: FeS2→Fe1−XS→FeS-FeO eutectic melt under reducing conditions and FeS2→Fe1−XS→FeO→Fe3O4→Fe2O3 under oxidation conditions.
2.
Non-Stoichiometry of Pyrrhotite (Fe1−XS):
In the Fe-S subsystem, the phase diagrams from Waldner and Pelton (2005) [36] and Lambert et al. (1998) [37] clarify that pyrrhotite (Fe1−XS) is an non-stoichiometric phase. Its composition (the value of x) is a function of both temperature and sulfur vapor pressure (pS2). For instance, FeS2 pyrolysis to FeSX occurs at 988 °C in a system composed of Fe and FeSX, while the pyrolysis temperature decreases to 743 °C in a system composed of FeS2 and FeSX.
3.
Existence of Key Phase Transition Temperatures:
In the Fe-S system, FeS2 pyrolysis initiates below 400 °C, producing non-stoichiometric Fe1−XS and sulfur vapor (S2). While within the (Fe + FeS) phase, Fe1−XS transforms via a peritectic reaction into liquid phase (liquid) and fcc-Fe at 910–988 °C [36,38,39,40,41,42,43]. Conversely, within (Fe1−XS + FeS2) phase, FeS2 directly pyrolyzes into Fe1−XS and sulfur vapor (S2) at 743 °C. Upon further heating to 1180 °C, this Fe1−XS melts into a sulfur-rich liquid, releasing additional sulfur vapor (S2) [36,38,39,40,41,42,43].
In the Fe-S-O system, a sequence of oxide formations occurs: Fe2O3 forms directly below 404 °C, with Fe3O4 as an intermediate phase stable between 404 °C and 552 °C. Sulfates, including FeSO4 and Fe2(SO4)3, are stable below 583 °C and 644 °C, respectively. At elevated temperatures, FeO melts at 1370–1424 °C, Fe3O4 melts at 1424–1597 °C, and Fe2O3 melts at around 1565 °C [28,34]. Notably, FeO is only stable under specific high-temperature, low-oxygen conditions.
Despite the delineation of these “thermodynamically feasible” pathways, significant discrepancies persist between these equilibrium predictions and experimental observations. This gap arises because equilibrium assumptions neglect several critical kinetic factors:
  • Local Atmospheric Conditions:
The gas composition (pO2 and pS2) at the particle surface can deviate drastically from the particle pores or at the reaction interface, which is influenced by gas flow rate, particle packing density, and heating rate.
  • Influence of Mineral Occurrence Mode:
For instance, the oxidation of coal-matrix-encapsulated FeS2 is significantly delayed, as oxygen must first diffuse through the coke layer.
  • Intra-Particle Temperature Gradients:
During rapid heating, a substantial temperature difference of 50–100 °C [16] exists between the particle exterior and interior, causing the surface and the core to follow different transformation pathways.

2.2. Thermodynamic Driving Force: Calculation ΔrGθ

We evaluated the Gibbs free energy change (ΔrGθ) for 24 major FeS2 transformation reactions, using data from Zhang et al. [35] to compare the thermodynamic feasibility of pyrolysis and oxidation pathways.
The data in Table 1 lead to two key conclusions.
First, at 500 K, sulfate formation is the most thermodynamically favored pathway, as indicated by the highest negative ΔrGθ value for Reactions 9 and 10. This is followed by direct oxidation to Fe3O4 and Fe2O3 (Reaction 4–6). The oxidation of Fe1−XS to Fe2(SO4)3 is also more favorable than its oxidation to iron oxides, as seen by comparing the ΔrGθ of Reactions 11–12, 14–15, and 17–18. In contrast, the pyrolysis of FeS2 to Fe1−XS or FeS (Reactions 1–2 and 7–8) is the least favored, exhibiting the highest ΔrGθ values at 500 K and 1000 K, confirming its non-spontaneous nature. However, at 1000 K, the thermodynamic favorability of sulfate formation decreases, becoming comparable to that of oxide formation.
Second, FeS2 pyrolysis is endothermic (ΔrH > 0), imposing a significant energy barrier. In contrast, oxidation reactions are exothermic (ΔrH < 0), which promotes self-heating and consequently accelerates reaction kinetics. This establishes that oxidation is both thermodynamically and kinetically preferable over pyrolysis.
However, this thermodynamic preference for oxidation relies on the assumption of sufficient oxygen supply to the reaction front. In the case that oxygen diffusion is limited (e.g., within large particles or under low bulk O2 concentrations), the oxidation process becomes kinetically hindered, potentially leading to pyrolysis instead.

2.3. Calculation of Thermodynamic Phase Diagram

The thermodynamic equilibrium calculations were performed using FactSage software (v8.4) with its associated databases (FToxid, FTsulf, and FactPS). These databases were selected for their comprehensive coverage of oxides, sulfides, and gaseous species, ensuring their suitability for simulating the high-temperature Fe-S-O system. The corresponding results are presented in Figure 1. The Y-axis represents temperature. The X-axis is defined by the initial molar ratio of FeS2/(FeS2 + O2), which quantitatively represents the reaction front redox environment.
1.
For FeS2/(FeS2 + O2) < 28.6% area (i.e., N (FeS2)/N (O2) < 0.4:1):
Below 1456 °C, the dominant phase is Fe2O3.
Between 1400 and 1580 °C, Fe3O4 is the primary product.
Above 1580 °C, liquid phase is the major phase.
2.
For 28.6% < FeS2/(FeS2 + O2) < 47.1% area (i.e., 0.4:1 < N (FeS2)/N (O2) < 0.89:1):
Below 480 °C, the dominant phase is Fe2O3.
Between 480 and 1069 °C, Fe3O4 is the primary product.
Above 1069 °C, liquid phase is the major phase.
3.
For 47.1% < FeS2/(FeS2 + O2) < 60% area (i.e., 0.89:1 < N (FeS2)/N (O2) < 1.5:1):
Below 1069 °C, the dominant phase is Fe3O4.
Above 1069 °C, the primary product is liquid phase.
4.
For 60% < FeS2/(FeS2 + O2) < 100% area (i.e., 1.5:1 < N (FeS2)/N (O2) < ∞):
Below 563 °C, the dominant phase is Fe3O4.
Between 563 and 947 °C, pyrolysis becomes dominant, and FeS is the primary product.
Above 947 °C, the primary product is liquid phase.
5
Sulfate area:
Fe2(SO4)3 is identified as an oxidation product at temperatures below approximately 600–700 °C for FeS2/(FeS2 + O2) < 26.7% (i.e., N (FeS2)/N (O2) < 0.36:1).
FeSO4 is observed as an oxidation product at temperatures below 395 °C for FeS2/(FeS2 + O2) = 26.5%–47.1% (i.e., N (FeS2)/N (O2) = 0.36:1 to 0.89:1).
Although thermodynamic analysis accurately predicts the stable equilibrium phases, the actual transformation pathways are kinetically controlled and exhibit more complex behavior. The following section reanalyzes the kinetic data to explain the discrepancies between these thermodynamic predictions and the experimental observations.

3. Reconciling Discrepant Kinetic Data

3.1. Reinterpreting Pyrolysis Mechanisms and Ea

The kinetics of the pyrolysis process are governed by two primary rate-controlling steps: (1) intrinsic chemical reaction control and (2) product layer diffusion control (apparent kinetics). Additionally, two-stage or transitional state models have also been proposed to describe more complex behavior.
1.
Chemical Reaction Controlled Model (Intrinsic)
The chemical reaction control mechanism dominates when sulfur vapor (S2) is rapidly and completely removed from the reaction interface, as occurs with coarse-grained or plate-like FeS2 particles in a high-flow inert atmosphere. This S-S bond-breaking process involves a high intrinsic activation energy barrier.
Morphologically, this process produces an “unreacted core-porous shell” structure, where the porous Fe1−XS shell results from S2 vapor release, continually exposing fresh FeS2 surface at the advancing reaction front [13,25,41,44,45,46,47,48,49,50,51,52,53,54,55,56,57].
2.
Diffusion Controlled Model (Apparent)
In experiments employing fine particles or under low-flow/static inert atmospheres (e.g., conventional Thermogravimetric Analysis, also called TGA), the rapidly released S2 accumulates due to inefficient convective removal. Consequently, the rate-limiting step shifts to the diffusion of accumulated S2 through the porous Fe1−XS product layer. This diffusion-controlled process exhibits a lower apparent activation energy, Ea, typically falling within the range of 90–150 kJ/mol.
3.
Two Steps and Transitional Model
Several studies have reported a two-stage or transitional process in the rate-controlling step during pyrolysis, leading to models that incorporate both chemical and diffusion control. For instance: Hoare et al. (1988) [53] identified a two-stage mechanism during non-isothermal pyrolysis (Figure 2). The initial stage was chemically controlled, following a three-dimensional shrinking core model (Ea = 286 kJ/mol), while the subsequent stage was diffusion-controlled, consistent with the Ginstling–Brounshtein model (Ea = 190 kJ/mol). Similarly, Lambert et al. (1998) [37] described the process as two consecutive reactions: FeS2→FeSX→FeS. The first stage, controlled by S2 vapor desorption, had an activation energy of 222 kJ/mol; the second stage, controlled by solid-state diffusion of iron, exhibited an apparent Ea of approximately 139 kJ/mol.
It should be noted that the assertion by Pemsler et al. [58] of chemical reaction control, based on the insensitivity of Ea to particle size under He flow, likely misinterprets “mixed control” within chemical control in conventional TGA as intrinsic kinetics.
The compiled data demonstrate that pyrite pyrolysis is characterized by two distinct Ea ranges, each corresponding to a specific rate-controlling step and set of experimental conditions.
As summarized in Table 2, a high Ea range (220–310 kJ/mol) characterizes intrinsic chemical reaction control (S–S bond cleavage), which dominates under conditions that ensure efficient S2 removal.
Conversely, Table 3 shows that a low Ea range (90–150 kJ/mol) signifies diffusion control, specifically the diffusion of S2 through the porous Fe1−XS layer. This mechanism prevails in systems where S2 accumulates due to rapid pyrolysis with high specific surface area, such as those employing fine powders or low-flow/static atmospheres.
Furthermore, the transitional models in Table 4 demonstrate a shift in the rate-limiting step, confirming that the apparent Ea is a dynamic indicator of the dominant mechanism. Similarly, the kinetics of mineral flotation also demonstrate that the rate-limiting steps undergo dynamic transitions with conditions such as particle size [66,67]. This understanding provides direct kinetic evidence for the competition-coupling mechanism.
The progressive inward advancement of the reaction front during pyrite pyrolysis is physically described by the shrinking core model. The kinetics of this process are mathematically defined by the rate-controlling step: the McKewan model applies to the initial, chemical reaction-controlled stage (high Ea), whereas the Ginstling–Brounshtein model describes the subsequent product layer diffusion-controlled stage (low Ea). For the very early surface reactions, a first-order kinetic equation often provides a valid approximation.
Critically, the dynamic transition between these controlling steps can be captured by the unifying Farrar–Smith model. Regarding kinetic analysis, the Coats–Redfern method is effective only when the correct reaction model is specified. In contrast, the model-free Friedman method directly reveals the continuous evolution of the apparent activation energy (Ea) with conversion. The initial slope method, by isolating the very beginning of the reaction, yields an Ea value closest to that of the intrinsic chemical control.

3.2. Clarifying the Complex Oxidation Mechanisms and Ea

The oxidation behavior is highly sensitive to the local environment and mineral occurrence mode, as evidenced by the significant variations in reported apparent (Ea) and the observed shifts in reaction mechanisms.

3.2.1. Predominant Oxidation Pathways

Synthesizing the extensive literature, the oxidation of FeS2 can be categorized into four predominant pathways, each with distinct characteristics:
1.
Direct Oxidation to Oxides:
Under low-temperature, high-oxygen conditions, FeS2 undergoes direct oxidation to Fe2O3 or Fe3O4 predominantly [16,46,67,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86], often forming a dense oxide shell that can passivate the particle and hinder further oxidation. The apparent Ea for this pathway is typically low (80–120 kJ/mol, Table 5), consistent with a rate-limiting step of oxygen diffusion through the product layer. Studies by McCarty et al. (1989) [77] and Schorr (1969) [83], which observed direct hematite formation without a pyrrhotite intermediate support this mechanism. The formation of a dense product layer is also responsible for a sharp increase in Ea observed under certain conditions.
2.
Indirect Oxidation Via Fe1−XS:
The sequential transformation of FeS2→Fe1−XS→Fe3O4/Fe2O3 predominates under low-oxygen partial pressures or in larger particles [18,19,22,32,45,71,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114]. In this pathway, pyrolysis precedes oxidation, and the porous intermediate Fe1−XS layer facilitates inward oxygen diffusion and outward SO2 release. This exemplifies a classical competition-coupling mechanism between pyrolysis and oxidation.
Pyrolysis and oxidation occur simultaneously in distinct zones within a particle, forming a characteristic “unreacted core-porous intermediate layer-oxide shell” morphology (e.g., Figure 3). The apparent Ea for this route spans a wide range (60–280 kJ/mol, Table 6), reflecting the dynamic balance and potential shift in the rate-determining step between chemical reaction (pyrolysis) and diffusion (oxidation). This mechanism is supported by studies from Dunn et al. (1989) [80], Hansen et al. (2003) [114], Zhou et al. (2018) [113], and Hu et al. (2006) [16], which documented morphology evolution and pathway dependence on O2 concentration and heating rate.
3.
Sulfate-Mediated Oxidation:
Under S2-rich atmospheres, FeS2 forms metastable FeSO4 or Fe2(SO4)3, which subsequently decompose to Fe2O3 at elevated temperatures [9,16,33,52,65,69,71,76,78,79,80,81,97,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131].
This process exhibits a dual Ea signature: a low Ea (~100 kJ/mol) for the exothermic sulfate formation, followed by a high Ea (190–220 kJ/mol) for the endothermic sulfate decomposition (Table 7). The significant volume expansion associated with sulfate formation can passivate the particle, further complicating the kinetics [16].
4.
Molten Phase-Dominated Transformation:
When the transformation temperature exceeds the melting points of iron-sulfur or iron-oxide phases (typically >1083 °C), the mechanism shifts from solid-state processes to melt-dominated kinetics [13,14,18,19,20,94,132,133,134,135,136,137,138,139,140,141]. This transition is characterized by particle spheroidization, accelerated coalescence, and dense sintered deposit formation [132,133,142].
The formation of a low-melting-point FeO-FeS eutectic melt (e.g., ~910 °C) can encapsulate unreacted or partially reacted cores, inhibiting sulfur release and facilitating deposit formation on the matrix. Studies by Huang et al. (2015, 2021) [143,144] and Yu et al. [145] confirm that such molten encapsulation becomes a dominant mechanism at temperatures around 950 °C.
Table 6. Indirect oxidation of pyrite via a porous Fe1−XS intermediate.
Table 6. Indirect oxidation of pyrite via a porous Fe1−XS intermediate.
No.AuthorParticle SizeEquipmentAtmosphereHeating RateModelStep1 Ea
kJ/mol
Step2 Ea
kJ/mol
Pathway
3Gomes [146]7.39 μmTGAAir
10 mL/min
2.5–20 °C/min.Second-order470–580 °C:145.6 kJ/molFeS2 chemical decomposition
15–20 °C/min.2.5–7.5 °C/min.First-order, Area-contracting580–1000
°C:
33.2
580–1000
°C:
281.4
Diffusion control in porousThe Fe2O3 layer inhabit reaction
4Yang [147] Theoretical calculations Density Functional Theory (DFT)197.96/175.83 Surface oxidation Ea of 197.96 kJ/mol,
Bulk sulfur migration Ea of 175.83 kJ/mol
Table 7. Sulfate-mediated oxidation and decomposition.
Table 7. Sulfate-mediated oxidation and decomposition.
No.AuthorParticle SizeEquipmentAtmosphereHeating RateModelStep1 Ea
kJ/mol
Step2 Ea
kJ/mol
Pathway
5Concer [61] TGAAir
50 mL/min
2.5, 5.0, 7.5, 10.0 K/minFriedman~650 K
924 kJ/mol
~770 K
451 kJ/mol
FeS2→Fe2(SO4)3→Fe2O3
6Coombs [148,149]0.1–10 μmTGAAir
430 mL/min
648–923 K
10.0 K/min
Friedman192 kJ/molSulfate decomposition 219 kJ/molFeS2→Fe3O4 (low T), FeS2→Fe2(SO4)3→Fe2O3(mid T), and direct Fe2O3 formation (high T).
7Ferrow [130]5–40 μmTGAAir200–380 °CWeibull + Arrhenius100 kJ/molFeS2→FeSO4 below 380 °C. Fe2O3 occurs over 380 °C.
8Tian [129]74 μmTGAAir 50 mL/min10 °C/minAvrami–Erofeev equation
Inverse Jander equation
194.81 kJ/molFeS2→FeSO4→Fe2O3
The data from Concer (2017) [61] are under questionable due to significant deviation from other existing studies.
The occurrence mode of pyrite (excluded vs. included) determines the final fate of the iron-containing melt, as shown in Figure 4. For excluded pyrite under reducing conditions, the pathway FeS2→Fe1−XS→FeO-FeS eutectic melt is well-established. Under oxidizing conditions, excluded particles can form FeO melt. In contrast, for included pyrite, the resulting Fe-S melt interacts with surrounding aluminosilicates to form FeS and Fe-bearing glass two phases (Figure 5). This indicates that pyrrhotite must first oxidize to FeO before it can be incorporated into the silicate glass phase. This important finding is well documented by McLennan et al. (2000) [28,29].
Under extreme conditions involving very high heating rates and oxygen concentrations (e.g., O2 ≥ 21%), rapid exothermic oxidation can induce complete particle melting. This results in the formation of spherical droplets with complex surface features (honeycomb-like, tortoise-shell patterns or dendritic) structures as shown in Figure 6, which are predominantly composed of Fe3O4 [72]. upon cooling, the formation of a dense Fe3O4 shell further passivates the particle. Kinetically, the melting process transitions from solid-state reaction or gas diffusion control to liquid-phase mass transfer.

3.2.2. Kinetic Model of Oxidation

Asaki et al. (1989, 1985) [138,139] were the first to propose an oxidation model for Fe1−XS particles in a vertical tube reactor, applying shrinking core model that integrated mass transfer, heat transfer, and interfacial reaction kinetics. Their model assumed that Fe1−XS oxidizes to solid Fe3O4 below 1468 K but melts at 1468 K. The simulations demonstrated that melting commenced between 950 and 1200 K at pO2 > 2.02 × 104 Pa, driven by exothermic reaction heat elevating the temperature beyond its melting point. The model also indicated that while smaller particles (e.g., 51 μm) initially oxidized faster than 88 μm particles due to their larger specific surface area, they may experience delayed further oxidation from earlier formation of denser oxide layers. A key limitation of this model is its presumption of an indirect oxidation pathway from pyrrhotite (Fe1−XS) to magnetite (Fe3O4), which fails to describe the formation of sulfate intermediates and their significant impact on reaction kinetics. Srinivasachar et al. (1989) [23] developed an alternative kinetic model base on thermal and mass balance equations, which suggest that the particles spend approximately 80% of the total transformation time in a molten state. Building on this work, Jassim et al. (2011) [141] improved the model by incorporating particle fragmentation, with simulations demonstrating that fragmentation shortens the heating time required to reach the melting point.

3.2.3. Synthesis of Oxidation Activation Energy

The complex and variable pathways of FeS2 oxidation have limited studies on its activation energy, as summarized in Table 5, Table 6 and Table 7.
The sudden shift in Ea from 82 kJ/mol to 293 kJ/mol around 460 °C in O2-CO2 observed by Hong et al. (1997) [25] is not only a temperature effect but results from a morphological transition in the product layer: a porous layer at low temperature allows for rapid diffusion (low Ea), while a dense layer at high temperatures impedes oxygen diffusion, raising the apparent Ea to near-pyrolysis levels.
Gomes et al. (2022) [146] found that during the late oxidation stage (580–1000 °C), the apparent Ea divided into two values (~33.2 kJ/mol and ~281.4 kJ/mol) depending on the heating rate, revealing two competing mechanisms: rapid heating maintains a porous structure, leading to reaction control (low Ea), whereas slow heating causes product layer densification, shifting the control to diffusion through the layer (high Ea). This study thus confirms that no single “oxidation mechanism” exists, as the apparent Ea is highly dependent on the intermediate phase microstructure dictated by the thermal history.
Notably, the implausibly high activation energy of 924 kJ/mol reported by Concer et al. [60] for the oxidation cannot represent an intrinsic kinetic barrier but rather signals severe diffusion limitations from layer passivation.
Table 5 shows that direct oxidation to oxides, governed by oxygen diffusion, is characterized by low Ea values (80–120 kJ/mol). In contrast, the indirect route via a Fe1−XS intermediate (Table 6) exhibits a wide Ea range (60–280 kJ/mol), reflecting the dynamic balance between pyrolysis and oxidation. Table 7 further highlights the dual role of sulfates, with low Ea formation and high Ea decomposition.
This pathway classification, however, does not fully capture the dynamic nature of the oxidation process. Vázquez et al. (2019) [150] observed a progressive decrease in Ea from ~450 kJ/mol to ~80 kJ/mol during a single oxidation process, indicating a clear shift in the rate-determining step. Similarly, Charpentier et al. (2010) [62] demonstrated the Ea value for pyrite pyrolysis varies with the reaction progress. Together, these observations demonstrate that the apparent Ea is not a fixed value but a dynamic indicator.
Kinetic models for pyrite oxidation are highly diverse, ranging from geometrical contraction models (surface chemical control) to Jander-type models (product layer diffusion control), reflecting the complexity and multi-mechanistic nature of the process. Furthermore, semi-empirical approaches like the Weibull + Arrhenius model are particularly suited for such complex systems with parallel pathways, as they characterize the distribution of reaction conversion and then determine the corresponding activation energy.

3.3. Evidence for the Interplay Between Pyrolysis and Oxidation Kinetics

To synthesize the extensive kinetic data from Section 3.1 and Section 3.2.3, Figure 7 illustrates the continuous variation in reported Ea for both pyrolysis and oxidation processes.
As shown in Figure 7, the Ea values are distributed across distinct intervals. Pyrolysis-dominated reactions exhibit activation energies of 220–310 kJ/mol (corresponding to S–S bond cleavage) and 90–150 kJ/mol (corresponding to S2 vapor diffusion). In contrast, direct oxidation-dominated reactions, governed by oxygen diffusion kinetics, have lower activation energy values (80–120 kJ/mol). However, indirect oxidation exhibits the broadest ranges (60 to 280 kJ/mol), where pyrolysis and oxidation occur simultaneously.
The empirical trend line in Figure 7 qualitatively maps the evolution of the apparent activation energy (Ea). More importantly, broad, continuous distribution of Ea values is a characteristic kinetic signature of the competition-coupling mechanism, in which Ea measured at any point represents the dynamic balance. The approach of resolving complex transformations into distinct mechanistic zones is consistent with previous studies that found that the dominant flotation mechanism shifts with particle size changes and energy input [66,151]. Consequently, this mechanism reconciled the discrepancy among Ea values reported in the literature.

4. Proposing the Competition-Coupling Mechanism and Constructing the Unified Phase Diagram

4.1. The Principle of Competition-Coupling Mechanism

Competition arises from intrinsic gradients of oxygen partial pressure and temperature within a particle. The surface, which is directly exposed to the bulk gas, constitutes an oxidation-dominated zone. In contrast, the particle core, where oxygen is depleted by the surface reaction and inward diffusion is hindered, becomes a pyrolysis-dominated zone.
Coupling occurs in two key stages:
  • Pyrolysis couples forward to oxidation: The initial pyrolysis of FeS2 generates a porous Fe1−XS intermediate, creating a “core-pore” structure. This porosity enhances the inward diffusion of oxygen, thereby facilitating the subsequent oxidation of the intermediate.
  • Oxidation couples backward to pyrolysis: The oxidation of Fe1−XS can lead to the formation of a dense oxide or sulfate shell, resulting in a “core-pore-shell” morphology. This shell limits oxygen access to the core, which in turn promotes the continuation of pyrolysis internally.

4.2. A Predictive Phase Diagram: Bridging Thermodynamics and Kinetics

Although Figure 1 offers a thermodynamic “map” of equilibrium possibilities, the actual transformation is governed by dominant reaction mechanisms under non-equilibrium conditions. To address this divergence, we propose a theoretical phase diagram (Figure 8) that incorporates competition-coupling mechanism. This diagram is categorized into four distinct transformation regions.
1.
Pyrolysis-Dominated Region
Location: High molar ratio regime where FeS2/(FeS2 + O2) ≈ 60%–100%, typically between 563 °C and 947 °C, or under fully inert conditions.
Kinetic Mechanism and Signature: This region is characterized by the endothermic pyrolysis of FeS2. The kinetics exhibit high apparent activation energies (220–310 kJ/mol) corresponding to the cleavage of S-S bonds in the initial stage (FeS2→Fe1−XS + S2). Morphologically, the reaction produces a characteristic “unreacted core-porous shell” structure.
Final Products: Troilite (FeS) is the primary product, which may form a molten phase above ~947 °C (or ~1180 °C if silicate interactions occur).
2.
Competition-Coupling Region
Location: This region covers a broad domain defined by three primary conditions: high molar ratio regime (FeS2/(FeS2 + O2) ≈ 100%–60%) at temperatures below 563 °C, intermediate molar ratio regime (FeS2/(FeS2 + O2) ≈ 60%–47.1%) at temperatures below 1069 °C, and low molar ratio regime (FeS2/(FeS2 + O2) ≈ 47.1%–28.6%) at temperatures ranging from 480 °C to 1069 °C.
Kinetic Mechanism and Signature: This is the most complex region, where pyrolysis and oxidation reactions compete and couple dynamically. The mechanism typically involves two pyrolysis stages generating a porous Fe1−XS intermediate. This enhanced porosity facilitates inward oxygen diffusion for further oxidation to Fe2O3, Fe3O4, and even sulfate. This coupling results in a multi-layered “unreacted core-porous layer-shell” morphology (e.g., FeS2 core/Fe1−XS layer/Fe3O4 or Fe2O3 or sulfate shell). Consequently, the apparent Ea in this region is not fixed [66] but forms a continuous curve (Figure 7).
3.
Oxidation-Dominated Region
Location: Low molar ratio regime where FeS2/(FeS2 + O2) ≈ 28.6%–0%, and regime where FeS2/(FeS2 + O2) ≈ 47.1%–28.6% is below 480 °C.
Kinetic Mechanism and Signature: The transformation is dominated by direct exothermic oxidation, characterized by lower apparent activation energies (80–120 kJ/mol) typical of diffusion of oxygen through a product layer. At lower temperatures, sulfates (FeSO4 and Fe2(SO4)3) form as metastable intermediates, leading to initial mass gain and particle passivation with Ea ≈ 100 kJ/mol. The subsequent decomposition of these sulfates has a higher Ea (190–220 kJ/mol). Morphologically, dense oxide/sulfate shells often form, which can hinder further reaction progress. Critically, the formation of such dense shells can shift the dominant mechanism from direct to indirect oxidation, resulting in a sharp increase in the apparent Ea.
4.
Melt-Dominated Region
Location: High-temperature domain across all molar ratio regimes.
Kinetic Mechanism and Signature: At elevated temperatures, solid products melt, shifting the transformation mechanism from solid-state reaction/diffusion to liquid phase mass transfer. Morphologically, particles start to spheroidize, which possibly leads to the formation of dense deposits upon cooling. The melting temperature is system-dependent: for example, it is ~947 °C for FeS, ~1180 °C for FeS-FeO, and 1069 °C for Fe3O4 under oxygen-lean and moderate-oxygen regimes, and ~1565 °C for Fe3O4 in an oxygen-rich regime.
Unlike conventional thermodynamic diagrams that describe equilibrium states (“what should happen”), our diagram’s key innovation is its ability to predict feasible pathways under non-equilibrium conditions (“what can happen”). It achieves this by incorporating competition-coupling mechanisms into its thermodynamic framework, thereby bridging the gap between thermodynamic prediction and kinetic reality.

4.3. Practical Modulators: Particle Size, Heating Rate, and Occurrence Mode

Additionally, practical combustion involves critical factors that indirectly modulate temperature and atmospheric composition.
1.
Particle Fragmentation Effects:
FeS2 pyrolysis releases sulfurous gases, generating internal pressure. When this pressure exceeds the yield strength at FeS2/Fe1−XS interfaces, it causes particle fragmentation. This process critically enhances oxidation accessibility [152], shifting the competitive balance toward oxidation.
Boyabat et al. (2004) [64] confirmed enhanced pyrolysis rates with decreasing particle size, while Srinivasachar et al. (1989) [23] established that FeS2 oxidation time is proportional to d2 (53–75 μm particles exhibit >30% slower S-release than 38–45 μm particles). Hu et al. (2006) [16] observed that fine particles (<0.045 mm) undergo direct oxidation to Fe2O3, whereas coarser fractions (0.09–0.12 mm) favor a sequential decomposition-oxidation pathway. This phenomenon, where particle size governs the reaction pathway, has also been observed in the kinetics of coal slime flotation [66,153]. Hawkins et al. (2014) [154], Baxter et al. (1992) [155,156], and Elder et al. (1965) [106] attributed accelerated oxidation to increased reactive surface area, surface roughness, and microcracking resulting from fragmentation.
In summary, fragmentation is a critical positive feedback mechanism that promotes oxidation by drastically improving O2 accessibility.
2.
Heating Rate Dependence:
Slow heating facilitates direct FeS2 oxidation but also leads to the formation of a sulfate/Fe2O3 layer that inhibits reactions. Conversely, rapid heating suppresses direct oxidation, allowing pyrolysis to occur preferentially. As demonstrated by Yani et al. (2010) [157], complete decomposition is achieved by 1200 K at 10 K/min, but remains incomplete even at 1250 K when heating rate is increased to 50 K/min.
Furthermore, high heating rates and large particle sizes yield underestimated Ea values due to internal temperature gradients, while isothermal or slow-heating conditions provide closer estimations of intrinsic Ea.
3.
Excluded vs. Included FeS2:
Excluded particles: Direct exposure enables rapid oxidation [158,159].
Included particles: Coke-derived reducing environments delay oxidation [87,160]. When in contact with aluminosilicates, it forms FeS/Fe-glass biphasic particles. Crucially, full incorporation of FeS into the glass phase requires its oxidation to FeO first. This pathway is consistent with the work of Zhao et al. (2015) [161] and Bool et al. (1995) (1996) [21,22], who summarized the distinct route as follows: Excluded FeS2→Fe1−XS→molten Fe-O-S→Fe3O4. Included FeS2→Fe1−XS→silicate-glass phase.

4.4. Validation: Mapping Pathways in the Literature onto the Phase Diagram

An attempt to validate the proposed unified phase diagram encountered a fundamental challenge: oxygen concentrations reported in the literature are typically given as volume percentages, which are influenced by gas flow rates, reactor geometry, and pressure. These values are not convertible to the molar ratio FeS2/(FeS2 + O2) used herein. Despite this limitation, the dominant transformation pathways identified in previous studies were systematically mapped onto the phase diagram in Figure 9.
1.
Direct and Sulfate-Mediated Oxidation (red arrowed line in the diagram):
FeS2Fe2O3: (3) Bhargava [102], (20) Hu [16], (23) Jorgensen [17], and (33) Almeida [119].
FeS2FeSO4 or Fe2(SO4)3Fe2O3: (1) M.Vázquez [150], (12) Earnest [78], (13) Thomas [118], (14) Komraus [162], (33) Almeida [119].
FeS2FeSO4Fe2O3: (17) Ferrow [130,131], (37) Concer [61].
FeS2FeSO4Fe2(SO4)3Fe3O4: (17) Ferrow, (39) Banerjee [120].
2.
Pyrrhotite (Fe1−XS)-Mediated Pathways (green arrowed line in the diagram):
FeS2Fe1−XSFeSO4Fe2(SO4)3Fe2O3: (1) M.Vázquez [150], (9) Paulik [122], (13) Thomas [118], (15) Zhang [35], (32) Coombs [148], (37) Concer [61], (39) Banerjee [120].
FeS2Fe1−XSFe2O3: (11) Music [46].
FeS2Fe1−XSFe3O4Fe2(SO4)3 or FeSO4Fe2O3: (23) Jorgensen [17], (25) PeterHansen [114].
FeS2Fe1−XS(Fe3O4)Fe2O3: (2) Hong [25] and Fegely [27], (10) Thrope, (19) Aracena [71,76], (20) Hu [16], (21) Bool [21,22], (27) Liu [73,74,163], (41) Schoenlaub [97].
FeS2Fe1−XSFe3O4 and Fe1−XSFe2O3: (4) Wang [103], (18) Tomeczek, (28) Wang [72,164], (30) McLennan [28], (31) Tao [165], (36) Yu [112], (38) Wang [98].
FeS2Fe1−XSFe3O4meltFe3O4Fe2O3: (24) Jassim [141].
3.
Melt-Involving Pathways (blue arrowed line in the diagram):
FeS2Fe1−XSFe3O4 (or Fe2O3)melt: (22) Asaki [138,139], (26) Sheng [159], (28) Wang [72,164], (29) Huffman [18,19,94], (34) Helbe [93], (35) Huffman [19], (36) Yu [112].
FeS2Fe1−XS→melt: (6) Huang [143,144].
FeS2Fe1−XS→FeS + FeO: (5) McLennan [29].
FeS2Fe1−XS→FeS + FeOFe3O4Fe2O3: (16) Srinivasachar [23].
4.
Other Pathways (yellow arrowed line in the diagram):
FeS2Fe3O4Fe1−XS: (8) Li [88].
FeS2Fe1−XSFe3O4: (3) Bhargava [102], (7) Wu [30,32].
The successful mapping of 16 distinct literature-derived pathways onto the four regions of our phase diagram (Figure 9) validates the framework’s consistency and predictive power. Previously anomalous pathways, such as (8) Li [88], (3) Bhargava [102], and (7) Wu [30,32], are logically explained within the competition-coupling region. This reinterpretation of disparate pathways emphasizes the validity of the proposed diagram and its advantage over intertwined pathways.

5. Conclusions and Future Perspectives

5.1. Key Findings and Conclusions

This work establishes a unified framework for pyrite thermal transformation, with the following key findings:
1.
Proposal of the Competition-Coupling Mechanism:
We proposed a novel competition-coupling mechanism wherein pyrolysis and oxidation compete under local O2 and temperature gradients, while simultaneously coupling through microstructural evolution. Specifically, pyrolysis generates a porous Fe1−XS intermediate that facilitates oxidation, which in turn forms a passivating layer that promotes further pyrolysis.
2.
Reconciliation of Kinetic Data: Ea as a Dynamic Signature of Mixed Reaction.
The apparent Ea is identified as an indicator of the rate-determining step. It varies continuously from pyrolysis-dominated regimes (high Ea for S–S bond cleavage) to oxidation-dominated regimes (low Ea for O2 diffusion). This principle provides a unifying interpretation for reconciling disparate kinetic data in the literature.
3.
Construction of Phase Diagram: Bridging Theory and Practice.
We constructed an original, unified phase diagram that categorizes the transformation process into four distinct regions—pyrolysis-dominated, competition-coupling, oxidation-dominated, and melt-dominated—based on the molar ratio FeS2/(FeS2 + O2) and temperature. This kinetically modified diagram predicts feasible pathways under non-equilibrium conditions.
This framework provides a fundamental understanding for controlling sulfur emissions and mitigating ash-related issues (e.g., slagging and fouling) in coal combustion, roasting, and gasification processes, thereby contributing to the advancement of clean coal utilization technologies.

5.2. Future Challenges and Research Directions

The proposed framework highlights several key challenges for future work:
  • Quantifying Kinetic Phase Boundaries: Future research should employ techniques like HT-SEM, HT-XRD, and in situ XPS to quantitatively map how boundaries shift with particle size, heating rate, and gas flow, ultimately enabling the construction of a precise predictive diagram.
  • Defining the Role of Sulfates: The critical switch between sulfate present as a temporary transitional intermediate or a final passivating layer requires clarification through surface analysis (XPS) and coupled TG-MS.
  • Quantifying the Competition-Coupling Mechanism: Advanced kinetic analysis is required to distinguish whether pyrolysis and oxidation reactions are independent or strongly coupled.

5.3. Theoretical Limitations

The proposed framework is subject to the following constraints:
  • Pathways, Not End States: The phase diagram predicts the dominant kinetic pathway, rather than the final thermodynamic equilibrium. This explains why pyrolysis can prevail over thermodynamically favored oxidation in realistic.
  • Apparent Ea as a Composite Metric: In the competition-coupling region, the apparent activation energy represents a blend of overlapping processes. It reflects the dynamic balance between pyrolysis and oxidation rather than a single elementary step.
  • Static Situation Assumption: The phase diagram does not account for dynamic flow or particle fragmentation, which can instantaneously expose fresh surfaces and drastically alter the transformation pathway.

Author Contributions

Conceptualization, G.X. and J.S.; Methodology, J.S. and M.L.; Software, J.S. and M.L.; Validation, G.X. and J.S.; Formal analysis, J.S. and M.L.; Investigation, M.L.; Resources, G.X. and J.S.; Data curation, M.L.; Writing—original draft preparation, M.L.; Writing—review and editing, J.S.; visualization, M.L.; supervision, G.X.; project administration, G.X.; funding acquisition, G.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (NSFC), “Study on the Mechanism and Regulation of Coarse Slime Desorption in the Slurry/Foam Interface Zone of a Flotation Column”, Grant No. 52074288. The APC was funded by Mingrui Liu.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors wish to sincerely thank Guangyuan Xie for his invaluable guidance and unwavering support throughout this study. His expertise and thoughtful feedback were essential in defining the research direction and enhancing the quality of the manuscript. We are also profoundly grateful to Yaoli Peng, Long Liang, Jiakun Tan, Wencheng Xia, and Weiguang Zhou for their insightful suggestions, scholarly input, and practical assistance during the literature review and methodological development. Particularly, we sincerely appreciate Jie Sha and Xiangning Bu for their exceptional dedication and patient guidance throughout the writing and revision process. This review would not have been possible without their combined wisdom and steadfast encouragement.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had a role in the collection, analysis, and interpretation of literature.

References

  1. Nie, Z.; Luo, J.; Tang, J.; Li, B.; Chen, B.; Gao, M.; Wang, X.; Guo, L. Pollution sources, characteristics and environmental risk assessment of heavy metals in surface water and sediments of typical pyrite mine in Southwest China. J. Environ. Sci. 2025, 157, 742–755. [Google Scholar] [CrossRef]
  2. Alhamed, M.; Wohnlich, S. Environmental impact of the abandoned coal mines on the surface water and the groundwater quality in the south of Bochum, Germany. Environ. Earth Sci. 2014, 72, 3251–3267. [Google Scholar] [CrossRef]
  3. Tabelin, C.B.; Corpuz, R.D.; Igarashi, T.; Villacorte-Tabelin, M.; Alorro, R.D.; Yoo, K.; Raval, S.; Ito, M.; Hiroyoshi, N. Acid mine drainage formation and arsenic mobility under strongly acidic conditions: Importance of soluble phases, iron oxyhydroxides/oxides and nature of oxidation layer on pyrite. J. Hazard. Mater. 2020, 399, 122844. [Google Scholar] [CrossRef]
  4. Machado de Oliveira, C.; Gesser Müller, T.; Patricio Ferreira, L.; Prado Cechinel, M.A.; Peterson, M.; Raupp-Pereira, F. Valorization of iron pyrite from coal mining in southern Brazil. J. Environ. Chem. Eng. 2019, 7, 102931. [Google Scholar] [CrossRef]
  5. Oliveira, C.M.; Machado, C.M.; Duarte, G.W.; Peterson, M. Beneficiation of pyrite from coal mining. J. Clean. Prod. 2016, 139, 821–827. [Google Scholar] [CrossRef]
  6. Park, I.; Tabelin, C.B.; Jeon, S.; Li, X.L.; Seno, K.; Ito, M.; Hiroyoshi, N. A review of recent strategies for acid mine drainage prevention and mine tailings recycling. Chemosphere 2019, 219, 588–606. [Google Scholar] [CrossRef]
  7. Zhang, K.; Liu, L.; Liu, L.; Yang, Y.; Liu, Q. Iron removal from kaolinitic coal gangue via magnetic separation after oxidizing calcination with the crystal structure of kaolinite protected. Mater. Today Commun. 2023, 37, 107175. [Google Scholar] [CrossRef]
  8. Gao, J.; Sui, H.; Wu, S.; Zhang, R.; Zhang, M.; Cui, B.; Chu, H. Interaction study of oxygen and iron-sulfur clusters based on the density functional theory. Int. J. Chem. Eng. 2022, 2022, 9812188. [Google Scholar] [CrossRef]
  9. Luo, B.; Peng, T.; Sun, H. Innovative Methodology for Sulfur Release from Copper Tailings by the Oxidation Roasting Process. J. Chem. 2020, 2020, 8090846. [Google Scholar] [CrossRef]
  10. Yadollahi, A.; Abdollahi, H.; Ardejani, F.D.; Mirmohammadi, M.; Magdouli, S. Bio-oxidation behavior of pyrite, marcasite, pyrrhotite, and arsenopyrite by sulfur- and iron-oxidizing acidophiles. Bioresour. Technol. Rep. 2021, 15, 100699. [Google Scholar] [CrossRef]
  11. Sokić, M.; Marković, B.; Stanković, S.; Kamberović, Ž.; Štrbac, N.; Manojlović, V.; Petronijević, N. Kinetics of chalcopyrite leaching by hydrogen peroxide in sulfuric acid. Metals 2019, 9, 1173. [Google Scholar] [CrossRef]
  12. Leiva, E.; Cayazzo, M.; Dávila, L.; Torres, M.; Ledezma, C. Acid mine drainage dynamics from a paste tailing deposit: Effect of sulfate content on the consistency and chemical stability after storage. Metals 2021, 11, 860. [Google Scholar] [CrossRef]
  13. Mayoral, M.C.; Andrés, J.M.; Izquierdo, M.T.; Rubio, B. Pyrrhotite deposition through thermal projection to simulate iron sulphide slagging in oxyfuel combustion. Fuel 2012, 101, 197–204. [Google Scholar] [CrossRef]
  14. Mayoral, M.C.; Izquierdo, M.T.; Andrés, J.M.; Rubio, B. Mechanism of interaction of pyrite with hematite as simulation of slagging and fireside tube wastage in coal combustion. Thermochim. Acta 2002, 390, 103–111. [Google Scholar] [CrossRef]
  15. Bryers, R. Physical and chemical characteristics of pyrites and their influence on fireside problems in steam generators. J. Eng. Power 1976, 98, 517–527. [Google Scholar] [CrossRef]
  16. Hu, G.; Dam-Johansen, K.; Wedel, S.; Hansen, J.P. Decomposition and oxidation of pyrite. Prog. Energy Combust. Sci. 2006, 32, 295–314. [Google Scholar] [CrossRef]
  17. Jorgensen, F.R.A.; Moyle, F.J. Phases formed during the thermal analysis of pyrite in air. J. Therm. Anal. 1982, 25, 473–485. [Google Scholar] [CrossRef]
  18. Huffman, G.P.; Huggins, F.E.; Levasseur, A.A.; Chow, O.; Srinivasachar, S.; Mehta, A.K. Investigation of the transformations of pyrite in a drop-tube furnace. Fuel 1989, 68, 485–490. [Google Scholar] [CrossRef]
  19. Huffman, G.P.; Huggins, F.E.; Dunmyre, G.R. Investigation of the high-temperature behaviour of coal ash in reducing and oxidizing atmospheres. Fuel 1981, 60, 585–597. [Google Scholar] [CrossRef]
  20. Groves, S.; Williamson, J.; Sanyal, A. Decomposition of pyrite during pulverized coal combustion. Fuel 1987, 66, 461–466. [Google Scholar] [CrossRef]
  21. Bool, L.E.; Helble, J.J. Iron oxidation state and its effect on ash particle stickiness. In Applications of Advanced Technology to Ash-Related Problems in Boilers; Baxter, L., DeSollar, R., Eds.; Springer: Boston, MA, USA, 1996; pp. 281–292. [Google Scholar]
  22. Bool, L.E.; Peterson, T.W.; Wendt, J.O.L. The partitioning of iron during the combustion of pulverized coal. Combust. Flame 1995, 100, 262–270. [Google Scholar] [CrossRef]
  23. Srinivasachar, S.; Boni, A.A. A kinetic model for pyrite transformations in a combustion environment. Fuel 1989, 68, 829–836. [Google Scholar] [CrossRef]
  24. Srinivasachar, S.; Helble, J.J.; Boni, A.A. Mineral behavior during coal combustion 1. Pyrite transformations. Prog. Energy Combust. Sci. 1990, 16, 281–292. [Google Scholar] [CrossRef]
  25. Hong, Y.; Fegley, B. The kinetics and mechanism of pyrite thermal decomposition. Berichte Bunsenges. Phys. Chem. 1997, 101, 1870–1881. [Google Scholar] [CrossRef]
  26. Fegley, B. Why pyrite is unstable on the surface of venus. Icarus 1997, 128, 474–479. [Google Scholar] [CrossRef]
  27. Fegley, B.; Lodders, K.; Treiman, A.H.; Klingelhöfer, G. The rate of pyrite decomposition on the surface of venus. Icarus 1995, 115, 159–180. [Google Scholar] [CrossRef]
  28. McLennan, A.R.; Bryant, G.W.; Bailey, C.W.; Stanmore, B.R.; Wall, T.F. An experimental comparison of the ash formed from coals containing Pyrite and Siderite mineral in oxidizing and reducing conditions. Energy Fuels 2000, 14, 308–315. [Google Scholar] [CrossRef]
  29. McLennan, A.R.; Bryant, G.W.; Stanmore, B.R.; Wall, T.F. Ash formation mechanisms during pf combustion in reducing conditions. Energy Fuels 2000, 14, 150–159. [Google Scholar] [CrossRef]
  30. Wu, J.; Yu, D.; Zeng, X.; Yu, X.; Yu, G.; Han, J.; Liu, F.; Xu, M. Impacts of CO2 on the pyrite–kaolinite interaction and the product sintering strength. Proc. Combust. Inst. 2019, 37, 4479–4486. [Google Scholar] [CrossRef]
  31. Cheng, H.; Liu, Q.; Zhang, S.; Wang, S.; Frost, R.L. Evolved gas analysis of coal-derived pyrite/marcasite. J. Therm. Anal. Calorim. 2014, 116, 887–894. [Google Scholar] [CrossRef]
  32. Wu, J.; Yu, D.; Yu, X.; Liu, F.; Chen, S.; Xu, M. High-temperature transformation of pyrite in CO2: Effects of residence time and the presence of O2. Proc. Combust. Inst. 2021, 38, 5493–5500. [Google Scholar] [CrossRef]
  33. Evangelou, V.P. Pyrite Oxidation and Its Control; CRC Press: Boca Raton, FL, USA, 1995; 293p. [Google Scholar]
  34. Darken, L.S.; Gurry, R.W. The system iron-oxygen. II. equilibrium and thermodynamics of liquid oxide and other phases. J. Am. Chem. Soc. 1946, 68, 798–816. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Li, Q.; Liu, X.L.; Xu, B.; Yang, Y.B.; Jiang, T. A Thermodynamic analysis on the roasting of pyrite. Minerals 2019, 9, 220. [Google Scholar] [CrossRef]
  36. Waldner, P.; Pelton, A.D. Thermodynamic modeling of the Fe-S system. J. Phase Equilibria Diffus. 2004, 26, 23–38. [Google Scholar] [CrossRef]
  37. Lambert, J.M.; Simkovich, G.; Walker, P.L. The kinetics and mechanism of the pyrite-to-pyrrhotite transformation. Metall. Mater. Trans. B 1998, 29, 385–396. [Google Scholar] [CrossRef]
  38. Kullerud, G.; Yoder, H.S. Pyrite stability relations in the Fe-S system. Econ. Geol. 1959, 54, 533–572. [Google Scholar] [CrossRef]
  39. Jensen, E. Pyrrhotite: Melting relation and composition. Am. J. Sci. 1942, 240, 695–709. [Google Scholar] [CrossRef]
  40. Hillert, M.; Staffansson, L.-I. An analysis of the phase equilibria in the Fe−FeS system. Metall. Mater. Trans. B 1975, 6, 37–41. [Google Scholar] [CrossRef]
  41. Morimoto, N.; Gyobu, A.; Mukaiyama, H.; Izawa, E. Crystallography and stability of pyrrhotites. Econ. Geol. 1975, 70, 824–833. [Google Scholar] [CrossRef]
  42. Moreau, J.G.; Jõeleht, A.; Aruväli, J.; Heikkilä, M.J.; Stojic, A.N.; Thomberg, T.; Plado, J.; Hietala, S. Bulk synthesis of stoichiometric/meteoritic troilite (FeS) by high-temperature pyrite decomposition and pyrrhotite melting. Meteorit. Planet. Sci. 2022, 57, 588–602. [Google Scholar] [CrossRef]
  43. Nickel, E.H. Mineral Chemistry of Metal Sulfides. Mineral. Mag. 1979, 43, 186–187. [Google Scholar] [CrossRef]
  44. Jagadeesh, M.S.; Seehra, M.S. Thermomagnetic studies of conversion of pyrite and marcasite in different atmospheres (vacuum, H2, He and CO). J. Phys. D Appl. Phys. 1981, 14, 2153–2167. [Google Scholar] [CrossRef]
  45. Monteiro, J.L.F. Thermal decomposition of pyrite in a fluidized bed. Can. J. Chem. Eng. 1981, 59, 511–516. [Google Scholar] [CrossRef]
  46. Music, S.; Popovi, S.; Risti, M. Thermal decomposition of pyrite. J. Radioanal. Nucl. Chem. 1992, 162, 217–226. [Google Scholar] [CrossRef]
  47. Nakazawa, H. Phase relations and modulated structures of pyrrhotite Fe1−XS. Nihon Kessho Gakkaishi 1980, 22, 251–262. [Google Scholar] [CrossRef]
  48. Kissin, S.A.; Scott, S.D. Phase relations involving pyrrhotite below 350 degrees C. Econ. Geol. 1982, 77, 1739–1754. [Google Scholar] [CrossRef]
  49. Stull, D.R.; Prophet, H. JANAF Thermochemical Tables; National Institute of Standards and Technology: Gaithersburg, MD, USA, 1971. [Google Scholar]
  50. Cheng, H.; Liu, Q.; Huang, M.; Zhang, S.; Frost, R.L. Application of TG-FTIR to study SO2 evolved during the thermal decomposition of coal-derived pyrite. Thermochim. Acta 2013, 555, 1–6. [Google Scholar] [CrossRef]
  51. Zhang, X.; Kou, J.; Sun, C. A comparative study of the thermal decomposition of pyrite under microwave and conventional heating with different temperatures. J. Anal. Appl. Pyrolysis 2019, 138, 41–53. [Google Scholar] [CrossRef]
  52. Tian, C.; Rao, Y.; Su, G.; Huang, T.; Xiang, C.; Kordulis, C. The Thermal Decomposition Behavior of Pyrite-Pyrrhotite Mixtures in Nitrogen Atmosphere. J. Chem. 2022, 2022, 8160007. [Google Scholar] [CrossRef]
  53. Hoare, I.C.; Hurst, H.J.; Stuart, W.I.; White, T.J. Thermal decomposition of pyrite. Kinetic analysis of thermogravimetric data by predictor–corrector numerical methods. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1988, 84, 3071–3077. [Google Scholar] [CrossRef]
  54. Yan, J.; Xu, L.; Yang, J. A study on the thermal decomposition of coal-derived pyrite. J. Anal. Appl. Pyrolysis 2008, 82, 229–234. [Google Scholar] [CrossRef]
  55. Zhao, H.L.; Bai, Z.Q.; Guo, Z.X.; Kong, L.X.; Wei, Y.C.; Li, H.Z.; Bai, J.; Li, W. In situ study of the decomposition of pyrite in coal during hydropyrolysis. J. Anal. Appl. Pyrolysis 2021, 154, 105024. [Google Scholar] [CrossRef]
  56. Rau, H. Energetics of defect formation and interaction in pyrrhotite Fe1−xS and its homogeneity range. J. Phys. Chem. Solids 1976, 37, 425–429. [Google Scholar] [CrossRef]
  57. Ruan, S.; Wang, C.; Jie, X.; Yin, F.; Zhang, Y.; Yao, Z.; Chen, Y. Kinetics of pyrite multi-step thermal decomposition in refractory gold sulphide concentrates. J. Therm. Anal. Calorim. 2021, 147, 3689–3702. [Google Scholar] [CrossRef]
  58. Pemsler, J.; Lam, R.; Litchfield, J.; Dallek, S.; Larrick, B.; Beard, B. Discharge behavior and thermal stability of synthetic FeS2 cathode material. J. Electrochem. Soc. 1990, 137, 1. [Google Scholar] [CrossRef]
  59. Coats, A.W.; Bright, N.F.H. The kinetics of the thermal decomposition of pyrite. Can. J. Chem. 1966, 44, 1191–1195. [Google Scholar] [CrossRef]
  60. Hu, H.P.; Chen, Q.Y.; Yin, Z.L.; Zhang, P.M.; Zou, J.P.; Che, H.S. Study on the kinetics of thermal decomposition of mechanically activated pyrites. Thermochim. Acta 2002, 389, 79–83. [Google Scholar] [CrossRef]
  61. Concer, P.H.; Oliveira, C.M.D.; Montedo, O.R.K.; Angioletto, E.; Peterson, M.; Fiori, M.A.; Moreira, R.D.F.P. Kinetics of the oxidation reactions and decomposition of pyrite. Cerâmica 2017, 63, 39–43. [Google Scholar] [CrossRef]
  62. Charpentier, L.; Masset, P.J. Thermal decomposition of pyrite FeS2 under reducing conditions. In Proceedings of the 7th Pacific Rim International Conference on Advanced Materials and Processing, Cairns, Australia, 2–6 August 2010; pp. 2398–2401. [Google Scholar]
  63. Lv, W.; Yu, D.; Wu, J.; Zhang, L.; Xu, M. The chemical role of CO2 in pyrite thermal decomposition. Proc. Combust. Inst. 2015, 35, 3637–3644. [Google Scholar] [CrossRef]
  64. Boyabat, N.; Özer, A.K.; Bayrakçeken, S.; Gülaboğlu, M.Ş. Thermal decomposition of pyrite in the nitrogen atmosphere. Fuel Process. Technol. 2004, 85, 179–188. [Google Scholar] [CrossRef]
  65. Schwab, G.-M.; Philinis, J. Reactions of iron pyrite-its thermal decomposition, reduction by Hydrogen and air oxidation. J. Am. Chem. Soc. 1947, 69, 2588–2596. [Google Scholar] [CrossRef]
  66. Bu, X.N.; Xie, G.Y.; Chen, Y.R.; Ni, C. The Order of Kinetic Models in Coal Fines Flotation. Int. J. Coal Prep. Util. 2016, 37, 113–123. [Google Scholar] [CrossRef]
  67. Ni, C.; Bu, X.N.; Xia, W.C.; Peng, Y.L.; Xie, G.Y. Effect of slimes on the flotation recovery and kinetics of coal particles. Fuel 2018, 220, 159–166. [Google Scholar] [CrossRef]
  68. Jovanović, D. Kinetics of thermal decomposition of pyrite in an inert atmosphere. J. Therm. Anal. 1989, 35, 1483–1492. [Google Scholar] [CrossRef]
  69. Luganov, V.A.; Shabalin, V.I. Behaviour of Pyrite During Heating. Can. Metall. Q. 2013, 21, 157–162. [Google Scholar] [CrossRef]
  70. Toro, C.; Torres, S.; Parra, V.; Fuentes, R.; Castillo, R.; Díaz, W.; Reyes, G.; Balladares, E.; Parra, R. On the detection of spectral emissions of iron oxides in combustion experiments of pyrite concentrates. Sensors 2020, 20, 1284. [Google Scholar] [CrossRef]
  71. Aracena, A.; Jerez, O. Mechanism and kinetics of pyrite transformation at elevated temperatures. Physicochem. Probl. Miner. Process. 2021, 57, 127–139. [Google Scholar] [CrossRef]
  72. Wang, T.; Zhang, H.; Yang, H.R.; Lv, J.F. Oxidation mechanism of pyrite concentrates (PCs) under typical circulating fluidized bed (CFB) roasting conditions and design principles of PCs’ CFB roaster. Chem. Eng. Process. Process Intensif. 2020, 153, 107944. [Google Scholar] [CrossRef]
  73. Lihui, L.; Qinfu, L.; Kenan, Z.; Shuai, Z.; Kuo, L.; Jintao, L.; Gaoyu, P. Thermal decomposition and oxidation of pyrite with different morphologies in the coal gangue of north China. J. Therm. Anal. Calorim. 2023, 148, 2039–2040. [Google Scholar] [CrossRef]
  74. Liu, L.; Liu, Q.; Zhang, S.; Li, Y.; Yang, L. The thermal transformation behavior and products of pyrite during coal gangue combustion. Fuel 2022, 324, 124803. [Google Scholar] [CrossRef]
  75. Pelovski, Y.; Petkova, V. Investigation on thermal decomposition of pyrites part 1. J. Therm. Anal. Calorim. 1999, 56, 95–99. [Google Scholar] [CrossRef]
  76. Aracena, A.; Jerez, O.; Ortiz, R.; Morales, J. Pyrite oxidation kinetics in an oxygen-nitrogen atmosphere at temperatures from 400 to 500 °C. Can. Metall. Q. 2016, 55, 195–201. [Google Scholar] [CrossRef]
  77. McCarty, K.F.; Hamilton, J.C.; Boehme, D.R.; Nagelberg, A.S. In situ Raman spectroscopy of high temperature pyrite reactions related to deposit formation from coal. J. Electrochem. Soc. 1989, 136, 1223–1229. [Google Scholar] [CrossRef]
  78. Earnest, C. Descriptive oxidative profiles for pyrite in the low temperature ash component of coals by differential thermal analysis. Thermochim. Acta 1984, 75, 219–232. [Google Scholar] [CrossRef]
  79. Dunn, J.G. The oxidation of sulphide minerals. Thermochim. Acta 1997, 300, 127–139. [Google Scholar] [CrossRef]
  80. Dunn, J.G.; De, G.C.; O’Connor, B.H. The effect of experimental variables on the mechanism of the oxidation of pyrite: Part 2. oxidation of particles of size 90–125 μm. Thermochim. Acta 1989, 155, 135–149. [Google Scholar] [CrossRef]
  81. Dunn, J.G.; Mackey, L.C. The measurement of the ignition temperatures of commercially important sulfide minerals. J. Therm. Anal. 1992, 38, 487–494. [Google Scholar] [CrossRef]
  82. Labus, M. Pyrite thermal decomposition in source rocks. Fuel 2021, 287, 119529. [Google Scholar] [CrossRef]
  83. Schorr, J.R.; Everhart, J.O. Thermal behavior of pyrite and its relation to carbon and sulfur oxidation in clays. J. Am. Ceram. Soc. 1969, 52, 351–354. [Google Scholar] [CrossRef]
  84. Chandra, A.P.; Gerson, A.R. Pyrite (FeS2) oxidation: A sub-micron synchrotron investigation of the initial steps. Geochim. Cosmochim. Acta 2011, 75, 6239–6254. [Google Scholar] [CrossRef]
  85. Özdemir, Ö.; Banerjee, S.K. High temperature stability of maghemite (γ-Fe2O3). Geophys. Res. Lett. 1984, 11, 161–164. [Google Scholar] [CrossRef]
  86. Feitknecht, W.; Gallagher, K.J. Mechanisms for the oxidation of Fe3O4. Nature 1970, 228, 548–549. [Google Scholar] [CrossRef]
  87. Chen, H.K.; Li, B.Q.; Zhang, B.J. Decomposition of pyrite and the interaction of pyrite with coal organic matrix in pyrolysis and hydropyrolysis. Fuel 2000, 79, 1627–1631. [Google Scholar] [CrossRef]
  88. Li, H.-Y.; Zhang, S.-H. Detection of mineralogical changes in pyrite using measurements of temperature-dependence susceptibilities. Chin. J. Geophys. 2005, 48, 1454–1461. [Google Scholar] [CrossRef]
  89. Prasad, A.; Singru, R.M.; Biswas, A.K. Study of the roasting of pyrite minerals by Mössbauer spectroscopy. Phys. Status Solidi (A) 1985, 87, 267–271. [Google Scholar] [CrossRef]
  90. Jerzy Tomeczek, H.P. Kinetics of mineral matter transformation during coal combustion. Fuel 2002, 81, 1251–1258. [Google Scholar] [CrossRef]
  91. Helble, J.; Neville, M.; Sarofim, A.F. Aggregate formation from vaporized ash during pulverized coal combustion. Symp. (Int.) Combust. 1988, 21, 411–417. [Google Scholar] [CrossRef]
  92. Helble, J.J.; Sarofim, A.F. Influence of char fragmentation on ash particle size distributions. Combust. Flame 1989, 76, 183–196. [Google Scholar] [CrossRef]
  93. Helble, J.J.; Srinivasachar, S.; Boni, A.A. Factors influencing the transformation of minerals during pulverized coal combustion. Prog. Energy Combust. Sci. 1990, 16, 267–279. [Google Scholar] [CrossRef]
  94. Huffman, G.P.; Huggins, F.E.; Shah, N.; Shah, A. Behavior of basic elements during coal combustion. Prog. Energy Combust. Sci. 1990, 16, 243–251. [Google Scholar] [CrossRef]
  95. Wang, X.S.; Ma, T.D.; Tang, Y.G.; Gupta, R.; Schobert, H.H.; Zhang, J.Y. Thermal transformation of coal pyrite with different structural types during heat treatment in air at 573–1473 K. Fuel 2022, 327, 124918. [Google Scholar] [CrossRef]
  96. Blachere, J. Desulfurization of pyrite. J. Am. Ceram. Soc. 1966, 49, 590–593. [Google Scholar] [CrossRef]
  97. Schoenlaub, R.A. Oxidation of pyrite. J. Am. Ceram. Soc. 1969, 52, 40–43. [Google Scholar] [CrossRef]
  98. Wang, L.; Pan, Y.X.; Li, J.H.; Qin, H.F. Magnetic properties related to thermal treatment of pyrite. Sci. China Ser. D-Earth Sci. 2008, 51, 1144–1153. [Google Scholar] [CrossRef]
  99. Thorpe, A.N.; Senftle, F.E.; Alexander, C.C.; Dulong, F.T. Oxidation of pyrite in coal to magnetite. Fuel 1984, 63, 662–668. [Google Scholar] [CrossRef]
  100. Nishihara, K. Studies on the oxidation of pyrite:(part 1) thermal decomposition of pyrite. Mem. Fac. Eng. 1958, 20, 285–306. [Google Scholar]
  101. Nishihara, K.; Kondo, Y. Studies on the oxidation of pyrite III: The intermediate products in the oxidation of pyrite. Fac. Eng. Kyoto Univ. 1959, 21, 214–228. [Google Scholar]
  102. Bhargava, S.K.; Garg, A.; Subasinghe, N.D. In situ high-temperature phase transformation studies on pyrite. Fuel 2009, 88, 988–993. [Google Scholar] [CrossRef]
  103. Wang, L.; Fan, B.W.; He, Y.T.; Li, P.; Yin, D.Q.; Hu, Y.H. Characteristics of minerals and their associations of transformation processes in pyrite at elevated temperatures: An X-ray diffraction study. Ironmak. Steelmak. 2013, 41, 147–152. [Google Scholar] [CrossRef]
  104. Colombo, U.; Gazzarrini, F.; Lanzavecchia, G.; Sironi, G. Magnetite oxidation: A proposed mechanism. Science 1965, 147, 1033. [Google Scholar] [CrossRef]
  105. Davis, B.L.; Rapp, G.; Walawender, M.J. Fabric and structural characteristics of the martitization process. Am. J. Sci. 1968, 266, 482–496. [Google Scholar] [CrossRef]
  106. Elder, T. Particle-size effect in oxidation of natural magnetite. J. Appl. Phys. 1965, 36, 1012–1013. [Google Scholar] [CrossRef]
  107. Feitknecht, W.; Mannweiler, U. Der Mechanismus der Umwandlung von γ-zu α-Eisensesquioxid [1]. Helv. Chim. Acta 1967, 50, 570–581. [Google Scholar] [CrossRef]
  108. Minyuk, P.; Tyukova, E.; Subbotnikova, T.; Kazansky, A.Y.; Fedotov, A. Thermal magnetic susceptibility data on natural iron sulfides of northeastern Russia. Russ. Geol. Geophys. 2013, 54, 464–474. [Google Scholar] [CrossRef]
  109. Zhang, T.; Wu, C.; Li, B.; Wang, C.; Chen, X.; Wei, J.; Yu, Q. Clarifying the decomposition process of pyrite and SO2 release in the cyclone preheater of a dry rotary cement kiln system. J. Clean. Prod. 2019, 241, 118422. [Google Scholar] [CrossRef]
  110. Lv, W.; Yu, D.; Wu, J.; Yu, X.; Du, Y.; Xu, M. A mechanistic study of the effects of CO2 on pyrrhotite oxidation. Proc. Combust. Inst. 2017, 36, 3925–3931. [Google Scholar] [CrossRef]
  111. Zhang, L.; Zhao, Y.; Yao, H. Mineral behavior during oxy-fuel combustion. In Oxy-Fuel Combustion; Elsevier: Amsterdam, The Netherlands, 2018; pp. 129–150. [Google Scholar]
  112. Yu, D.; Yu, X.; Wu, J.; Han, J.; Liu, F.; Pan, H. A comprehensive review of ash issues in oxyfuel combustion of coal and biomass: Mineral matter transformation, ash formation, and deposition. Energy Fuels 2021, 35, 17241–17260. [Google Scholar] [CrossRef]
  113. Zhou, Y.; Xu, P.; Cheng, H.; Liu, Q. Thermal phase transition of pyrite from coal: Implication for the environmental impact of sulfur pollution. J. Therm. Anal. Calorim. 2018, 134, 2391–2396. [Google Scholar] [CrossRef]
  114. Hansen, J.P.; Jensen, L.S.; Wedel, S.; Dam-Johansen, K. Decomposition and oxidation of pyrite in a fixed-bed reactor. Ind. Eng. Chem. Res. 2003, 42, 4290–4295. [Google Scholar] [CrossRef]
  115. Shkodin, V.G.; Abishev, D.N.; Kobzhasov, A.K.; Malyshev, V.P.; Mangutova, R.F. The question of thermal decomposition of pyrite. J. Therm. Anal. 1978, 13, 49–53. [Google Scholar] [CrossRef]
  116. Pysh’Yev, S.; Gayvanovych, V.; Pattek-Janczyk, A.; Stanek, J. Oxidative desulphurisation of sulphur-rich coal. Fuel 2004, 83, 1117–1122. [Google Scholar] [CrossRef]
  117. Hausen, D.M. Reversible reactions between pyrite and pyrrhotite in S02. J. Miner. 1991, 43, 31–34. [Google Scholar] [CrossRef]
  118. Thomas, P.S.; Hirschausen, D.; White, R.E.; Guerbois, J.P.; Ray, A.S. Characterisation of the oxidation products of pyrite by thermogravimetric and evolved gas analysis. J. Therm. Anal. Calorim. 2003, 72, 769–776. [Google Scholar] [CrossRef]
  119. Almeida, C.M.V.B.; Giannetti, B.F. Comparative study of electrochemical and thermal oxidation of pyrite. J. Solid State Electrochem. 2002, 6, 111–118. [Google Scholar] [CrossRef]
  120. Banerjee, A.C.; Rangaswamy, P.; Sood, S. Mechanism of oxidation of iron pyrite in dynamic air. Therm. Anal. 1980, 2, 241–246. [Google Scholar] [CrossRef]
  121. Banerjee, A.C. Mechanism of oxidation of iron pyrites. J. Chem. Soc. D Chem. Commun. 1971, 1006–1007. [Google Scholar] [CrossRef]
  122. Paulik, F.; Paulik, J.; Arnold, M. Kinetics and mechanism of the decomposition of pyrite under conventional and quasi-isothermal–quasi-isobaric thermoanalytical conditions. J. Therm. Anal. 1982, 25, 313–325. [Google Scholar] [CrossRef]
  123. Eneroth, E.; Koch, C.B. Crystallite size of haematite from thermal oxidation of pyrite and marcasite—Effects of grain size and iron disulphide polymorph. Miner. Eng. 2003, 16, 1257–1267. [Google Scholar] [CrossRef]
  124. Cole, D.A.; Simmons, G.W.; Herman, R.G.; Klier, K.; Czakó-Nagy, I. Transformations of iron minerals during coal oxidation. Fuel 1987, 66, 1240–1248. [Google Scholar] [CrossRef]
  125. Descostes, M.; Mercier, F.; Beaucaire, C.; Zuddas, P.; Trocellier, P. Nature and distribution of chemical species on oxidized pyrite surface: Complementarity of XPS and nuclear microprobe analysis. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2001, 181, 603–609. [Google Scholar] [CrossRef]
  126. Usher, C.R.; Paul, K.W.; Narayansamy, J.; Kubicki, J.D.; Sparks, D.L.; Schoonen, M.A.; Strongin, D.R. Mechanistic aspects of pyrite oxidation in an oxidizing gaseous environment: An in situ HATR-IR isotope study. Environ. Sci. Technol. 2005, 39, 7576–7584. [Google Scholar] [CrossRef]
  127. Kennedy, T.; Sturman, B.T. The oxidation of iron (II) sulphide. J. Therm. Anal. 1975, 8, 329–337. [Google Scholar] [CrossRef]
  128. Rimstidt, J.D.; Vaughan, D.J. Pyrite oxidation: A state-of-the-art assessment of the reaction mechanism. Geochim. Cosmochim. Acta 2003, 67, 873–880. [Google Scholar] [CrossRef]
  129. Tian, C.; Rao, Y.; Su, G.; Huang, T.; Niu, Y. Effects of Pyrrhotite on the Combustion Behavior and the Kinetic Mechanism of Pyrite-Pyrrhotite Mixture Powders in the Air. Int. J. Chem. Eng. 2023, 2023, 9567708. [Google Scholar] [CrossRef]
  130. Ferrow, E.A.; Mannerstrand, M.; Sjöberg, B. Reaction kinetics and oxidation mechanisms of the conversion of pyrite to ferrous sulphate: A Mössbauer spectroscopy study. Hyperfine Interact. 2005, 163, 109–119. [Google Scholar] [CrossRef]
  131. Ferrow, E.A.; Sjöberg, B.A. Oxidation of pyrite grains: A mössbauer spectroscopy and mineral magnetism study. Hyperfine Interact. 2006, 163, 95–108. [Google Scholar] [CrossRef]
  132. Li, Y.; Gao, X.; Wu, H. Ash cenosphere from solid fuels combustion. Part 2: Significant role of ash cenosphere fragmentation in ash and particulate matter formation. Energy Fuels 2013, 27, 822–829. [Google Scholar] [CrossRef]
  133. Li, Y.; Wu, H. Ash cenosphere from solid fuels combustion. Part 1: An investigation into Its formation mechanism using Pyrite as a model fuel. Energy Fuels 2012, 26, 130–137. [Google Scholar] [CrossRef]
  134. Senior, C.L.; Srinivasachar, S. Viscosity of ash particles in combustion systems for prediction of particle sticking. Energy Fuels 2002, 9, 277–283. [Google Scholar] [CrossRef]
  135. Wall, T.F. Mineral matter transformations and ash deposition in pulverised coal combustion. Symp. (Int.) Combust. 1992, 24, 1119–1126. [Google Scholar] [CrossRef]
  136. Nowok, J.W.; Benson, S.A.; Jones, M.L.; Kalmanovitch, D.P. Sintering behaviour and strength development in various coal ashes. Fuel 1990, 69, 1020–1028. [Google Scholar] [CrossRef]
  137. Fryda, L.E.; Sobrino, C.; Glazer, M.; Bertrand, C.; Cieplik, M.K. Study of ash deposition during coal combustion under oxyfuel conditions. Fuel 2012, 92, 308–317. [Google Scholar] [CrossRef]
  138. Asaki, Z.; Kondo, Y. Oxidation kinetics of iron sulfide in the form of dense plate, pellet and single particle. J. Therm. Anal. 1989, 35, 1751–1759. [Google Scholar] [CrossRef]
  139. Asaki, Z.; Mori, S.; Ikeda, M.; Kondo, Y. Oxidation of pyrrhotite particles falling through a vertical tube. Metall. Trans. B 1985, 16, 627–638. [Google Scholar] [CrossRef]
  140. Tenbrink, H.; Eenkhoorn, S.; Hamburg, G. A fundamental investigation of the flame kinetics of coal pyrite. Fuel 1996, 75, 945–951. [Google Scholar] [CrossRef]
  141. Jassim, E.; Benson, S.A.; Bowman, F.M.; Seames, W.S. The influence of fragmentation on the behavior of pyrite particles during pulverized coal combustion. Fuel Process. Technol. 2011, 92, 970–976. [Google Scholar] [CrossRef]
  142. Wu, H.W.; Li, Y. Ash cenosphere fragmentation during pulverised pyrite combustion: Importance of cooling. Proc. Combust. Inst. 2019, 37, 2773–2780. [Google Scholar] [CrossRef]
  143. Huang, F.; Zhang, L.-Q.; Yi, B.-J.; Xia, Z.-J.; Zheng, C.-G. Transformation pathway of excluded mineral pyrite decomposition in CO2 atmosphere. Fuel Process. Technol. 2015, 138, 814–824. [Google Scholar] [CrossRef]
  144. Huang, F.; Xin, S.; Mi, T.; Zhang, L. Study of pyrite transformation during coal samples heated in CO2 atmosphere. Fuel 2021, 292, 120269. [Google Scholar] [CrossRef]
  145. Yu, D.X.; Zhao, L.; Zhang, Z.Y.; Wen, C.; Xu, M.H.; Yao, H. Iron transformation and ash fusibility during coal combustion in air and O2/CO Medium. Energy Fuels 2012, 26, 3150–3155. [Google Scholar] [CrossRef]
  146. Gomes, T.; da Rosa, R.; Cargnin, M.; Quadri, M.B.; Peterson, M.; de Oliveira, C.M.; da Rosa Rabelo, N.; Angioletto, E. Pyrite roasting in modified fluidized bed: Experimental and modeling analysis. Chem. Eng. Sci. 2022, 261, 117977. [Google Scholar] [CrossRef]
  147. Yang, Y.J.; Liu, J.; Liu, F.; Wang, Z. Comprehensive evolution mechanism of SO2 formation during pyrite oxidation. Proc. Combust. Inst. 2019, 37, 2809–2819. [Google Scholar] [CrossRef]
  148. Coombs, P.; Munir, Z. The mechanism of oxidation of ferrous sulfide (FeS) powders in the range of 648 to 923 K. Metall. Trans. B 1989, 20, 661–670. [Google Scholar] [CrossRef]
  149. Coombs, P.; Munir, Z. The decomposition of iron (III) sulfate in air. J. Therm. Anal. 1989, 35, 967–976. [Google Scholar] [CrossRef]
  150. Vázquez, M.; Moreno-Ventas, I.; Raposo, I.; Palma, A.; Díaz, M.J. Kinetic of pyrite thermal degradation under oxidative environment. J. Therm. Anal. Calorim. 2019, 141, 1157–1163. [Google Scholar] [CrossRef]
  151. Mao, Y.; Bu, X.; Peng, Y.; Tian, F.; Xie, G. Effects of simultaneous ultrasonic treatment on the separation selectivity and flotation kinetics of high-ash lignite. Fuel 2020, 259, 116270. [Google Scholar] [CrossRef]
  152. Flagan, R.C. Submicron particles from coal combustion. Symp. (Int.) Combust. 1979, 17, 97–104. [Google Scholar] [CrossRef]
  153. Ni, C.; Xie, G.Y.; Jin, M.G.; Peng, Y.L.; Xia, W.C. The difference in flotation kinetics of various size fractions of bituminous coal between rougher and cleaner flotation processes. Powder Technol. 2016, 292, 210–216. [Google Scholar] [CrossRef]
  154. Hawkins, A.B. Engineering implications of the oxidation of pyrite: An overview, with particular reference to Ireland. In Implications of Pyrite Oxidation for Engineering Works; Springer International Publishing AG: Cham, Switzerland, 2014; pp. 1–98. [Google Scholar]
  155. Baxter, L.L. Char fragmentation and fly ash formation during pulverized-coal combustion. Combust. Flame 1992, 90, 174–184. [Google Scholar] [CrossRef]
  156. Baxter, L.L.; Mitchell, R.E. The release of iron during the combustion of Illinois No. 6 coal. Combust. Flame 1992, 88, 1–14. [Google Scholar] [CrossRef]
  157. Yani, S.; Zhang, D.K. An experimental study into pyrite transformation during pyrolysis of Australian lignite samples. Fuel 2010, 89, 1700–1708. [Google Scholar] [CrossRef]
  158. Bailey, C.W.; Bryant, G.W.; Matthews, E.M.; Wall, T.F. Investigation of the high-temperature behavior of excluded siderite grains during pulverized fuel combustion. Energy Fuels 1998, 12, 464–469. [Google Scholar] [CrossRef]
  159. Sheng, C.D.; Lin, J.; Li, Y.; Wang, C. Transformation behaviors of excluded pyrite during O2/CO combustion of pulverized coal. Asia-Pac. J. Chem. Eng. 2010, 5, 304–309. [Google Scholar] [CrossRef]
  160. Vuthaluru, H.B.; Eenkhoorn, S.; Hamburg, G.; Heere, P.G.T.; Kiel, J.H.A. Behaviour of iron-bearing minerals in the early stages of pulverised coal conversion processes. Fuel Process. Technol. 1998, 56, 21–31. [Google Scholar] [CrossRef]
  161. Zhao, H.-L.; Bai, Z.-Q.; Yan, J.-C.; Bai, J.; Li, W. Transformations of pyrite in different associations during pyrolysis of coal. Fuel Process. Technol. 2015, 131, 304–310. [Google Scholar] [CrossRef]
  162. Komraus, J.L.; Popiel, E.S.; Mocek, R. Chemical transformations of ferruginous minerals during the process of oxidation of hard coal. Hyperfine Interact. 1990, 58, 2589–2592. [Google Scholar] [CrossRef]
  163. Liu, L.; Liu, Q.; Li, Y.; Cao, H.; Kang, X. Occurrence of iron in the minerals of carboniferous coal gangue of the Pingshuo open-pit mine, North China. Clays Clay Miner. 2022, 70, 695–711. [Google Scholar] [CrossRef]
  164. Wang, T.; Zhang, H.; Liu, Q.; Yang, H. Experimental studies on phase transformation during pyrite concentrate oxidation under circulating fluidized bed (CFB) roasting conditions. Ind. Eng. Chem. Res. 2011, 50, 14168–14174. [Google Scholar] [CrossRef]
  165. Zeng, T.; Helble, J.J.; Bool, L.E.; Sarofim, A.F. Iron transformations during combustion of Pittsburgh no. 8 coal. Fuel 2009, 88, 566–572. [Google Scholar] [CrossRef]
Figure 1. Phase diagram of FeS2 transformation at 1 atm via FactSage software.
Figure 1. Phase diagram of FeS2 transformation at 1 atm via FactSage software.
Minerals 15 01139 g001
Figure 2. Dark-field optical microscopy of a single FeS2 particle taken at various stages of pyrolysis: (a) ϕ = 0.94; (b) ϕ = 0.87; (c) ϕ = 0.81; (d) ϕ = 0.77. (Reproduced from Hoare et al. (1988) [53], Copyright 1988 Royal Society of Chemistry. Permission conveyed through Copyright Clearance Center, Inc.).
Figure 2. Dark-field optical microscopy of a single FeS2 particle taken at various stages of pyrolysis: (a) ϕ = 0.94; (b) ϕ = 0.87; (c) ϕ = 0.81; (d) ϕ = 0.77. (Reproduced from Hoare et al. (1988) [53], Copyright 1988 Royal Society of Chemistry. Permission conveyed through Copyright Clearance Center, Inc.).
Minerals 15 01139 g002
Figure 3. SEM images of oxidized pyrite particles: (a) a partially oxidized particle exhibiting a characteristic multi-layered structure with an unreacted FeS2 core, a porous Fe1−XS intermediate layer, and a dense oxide shell; (b) a fully oxidized particle. (Reproduced with permission from Hu et al. (2006) [16]. Copyright 2006 Elsevier Science & Technology Journals. Permission conveyed through Copyright Clearance Center, Inc. The figure was originally published by Dunn et al. (1989) [80]).
Figure 3. SEM images of oxidized pyrite particles: (a) a partially oxidized particle exhibiting a characteristic multi-layered structure with an unreacted FeS2 core, a porous Fe1−XS intermediate layer, and a dense oxide shell; (b) a fully oxidized particle. (Reproduced with permission from Hu et al. (2006) [16]. Copyright 2006 Elsevier Science & Technology Journals. Permission conveyed through Copyright Clearance Center, Inc. The figure was originally published by Dunn et al. (1989) [80]).
Minerals 15 01139 g003
Figure 4. Ash formation mechanisms for pyrite. (a) Excluded pyrite; (b) included pyrite (Reproduced from McLennan et al. (2000) [29]. Copyright 2000 American Chemical Society).
Figure 4. Ash formation mechanisms for pyrite. (a) Excluded pyrite; (b) included pyrite (Reproduced from McLennan et al. (2000) [29]. Copyright 2000 American Chemical Society).
Minerals 15 01139 g004
Figure 5. (a) The FeO-FeS eutectic ash particle; (b) Fe-O-S/Fe-glass phase two phases. (Reproduced from McLennan et al. (2000) [29]. Copyright 2000 American Chemical Society).
Figure 5. (a) The FeO-FeS eutectic ash particle; (b) Fe-O-S/Fe-glass phase two phases. (Reproduced from McLennan et al. (2000) [29]. Copyright 2000 American Chemical Society).
Minerals 15 01139 g005
Figure 6. SEM images of DTF sample surface (Tf = 850 °C, [O2] = 10.0%, tR = 5.0 s). (a) Typical particle I. (b) Typical particle II. (Reproduced with permission from Wang et al. (2020) [72]. Copyright 2020 Elsevier. Permission conveyed through Copyright Clearance Center, Inc.).
Figure 6. SEM images of DTF sample surface (Tf = 850 °C, [O2] = 10.0%, tR = 5.0 s). (a) Typical particle I. (b) Typical particle II. (Reproduced with permission from Wang et al. (2020) [72]. Copyright 2020 Elsevier. Permission conveyed through Copyright Clearance Center, Inc.).
Minerals 15 01139 g006
Figure 7. Evolution of apparent Ea during pyrite transformation: mechanism shift from pyrolysis-dominated (high Ea) to oxidation-dominated (low Ea).
Figure 7. Evolution of apparent Ea during pyrite transformation: mechanism shift from pyrolysis-dominated (high Ea) to oxidation-dominated (low Ea).
Minerals 15 01139 g007
Figure 8. A kinetically modified phase diagram of thermal transformation of FeS2 in coal. (Thermodynamic calculations yield sharp phase boundaries under ideal conditions. However, kinetic constraints such as diffusion limitations, local atmosphere, and temperature gradient variations result in blurred boundaries and a continuous transition between mechanisms).
Figure 8. A kinetically modified phase diagram of thermal transformation of FeS2 in coal. (Thermodynamic calculations yield sharp phase boundaries under ideal conditions. However, kinetic constraints such as diffusion limitations, local atmosphere, and temperature gradient variations result in blurred boundaries and a continuous transition between mechanisms).
Minerals 15 01139 g008
Figure 9. Verification of the diagram against proposed reaction pathways (note: the reaction routes cited in the figure are schematic; temperature and oxygen concentration are not in exact correspondence).
Figure 9. Verification of the diagram against proposed reaction pathways (note: the reaction routes cited in the figure are schematic; temperature and oxygen concentration are not in exact correspondence).
Minerals 15 01139 g009
Table 1. Possible chemical reactions and corresponding ΔrGθ at temperatures of 500 and 1000 K. (Data source: Zhang et al. [35].)
Table 1. Possible chemical reactions and corresponding ΔrGθ at temperatures of 500 and 1000 K. (Data source: Zhang et al. [35].)
No.ReactionEquationΔrGθ at 500 K
kJ/mol
ΔrGθ at 1000 K
kJ/mol
1FeS2 = 8/7Fe0.875S + 3/7S2(g)−0.1219 T + 109.081748.13−12.82
2FeS2 = FeS + 1/2S2(g)−0.1375 T + 140.234471.482.73
3S + 2O2 = 2SO20.1465 T − 723.7860572.481003.73
4FeS2 + 11/4O2(g) = 1/2Fe2O3 + 2SO2(g)0.0756 T − 832.7516−794.95−757.15
5FeS2 + 7/4O2(g) = 1/2Fe2O3 + SO2(g)0.1407 T− 611.2840−540.93−470.58
6FeS2 + 8/3O2(g) = 1/3Fe3O4 + 2SO2(g)0.0519 T − 792.2288−766.28−740.33
7FeS2 + O2(g) = FeS + SO2(g)−0.0647 T − 221.4677−253.82−286.17
8FeS2 + 6/7O2(g) = 8/7Fe0.875S + 6/7SO2(g)−0.0592 T − 201.0982−230.70−260.30
9FeS2 + 3O2(g) = FeSO4 + SO2(g)0.2920 T − 1050.5365−904.54−758.54
10FeS2 + 7/2O2(g) = 1/2Fe2(SO4)3 + SO2(g)0.4859 T − 1265.3779−1022.43−779.48
11Fe0.875S + 53/32O2(g) = 7/16Fe2O3 + SO2(g)0.1192 T − 553.6720−494.07−434.47
12Fe0.875S + 19/12O2(g) = 7/24Fe3O4 + SO2(g)0.0985 T − 518.6612−469.41−420.16
13Fe0.875S + 1/8O2(g) = 7/8FeS + 1/8SO2(g)−0.038 T − 18.0233−37.02−56.02
14Fe0.875S + O2(g) = 7/8FeSO4 + 1/8SO2(g)0.3099 T − 744.6433−589.69−434.74
15Fe0.875S + 47/16O2(g) = 7/16Fe2(SO4)3 + 5/16SO2(g)0.4938 T − 1127.0820−880.18−633.28
16FeS + 5/3O2(g) = 1/3Fe3O4 + SO2(g)0.1166 T − 570.7611−512.46−454.16
17FeS + 2O2(g) = FeSO40.3566 T − 829.0689−650.77−472.47
18FeS + 5/2O2(g) + 1/2SO2(g) = 1/2Fe2(SO4)3 0.5505 T − 1043.9102−768.66−493.41
19Fe3O4 + 1/4O2(g) = 3/2Fe2O30.0722 T − 121.5684−85.47−49.37
20Fe3O4 + O2(g) + 3SO2(g) = 3FeSO40.7201 T − 774.9232−414.87−54.82
21Fe3O4 + 5/2O2(g) + 9/2SO2(g) = 3/2Fe2(SO4)31.3018 T − 1419.4472−768.55−117.65
22Fe2O3 + 1/2O2(g) + 2SO2(g) = 2FeSO40.4319 T − 435.5699−219.62−3.67
23Fe2O3 + 3/2O2(g) + 3SO2(g) = Fe2(SO4)30.8197 T − 865.2525−455.40−45.55
24FeSO4 + 1/2O2(g) + 1/2SO2(g) = 1/2Fe2(SO4)30.1939 T − 214.8413−117.89−20.94
Table 2. Pyrite pyrolysis—chemical reaction controlled mechanism (high Ea).
Table 2. Pyrite pyrolysis—chemical reaction controlled mechanism (high Ea).
No.AuthorParticle SizeEquipmentAtmosphereHeating RateModelEa Value, kJ/mol Mechanism
1Hong [25]2 × 0.05 cm
slice
Vertical tubeHe, N2,
75 mL/min
Isothermal
400–590 °C
Linear dynamics297 ± 34Lattice decomposition
CO2,
75 mL/min
275± 10 Reacted with CO2
2Coats [59]0.25 inches cylinderVertical Quartz TubeAr
180 mL/min
Isothermal 600–653 °CMcKewan model270.6–290.7 Chemically controlled
3Pannetier (as cited from [16,25]) Weight LossVacuumIsothermal
451–476 °C
Linear kinetics310
4Hu [60]<1 mmTG-DTGDynamic Ar1000 °C,
2.5, 5, 7.5, 15 K/min
Friedman268Lattice defects pyrolysis
5Concer [61] TGAN2
50 mL/min
2.5, 5.0, 7.5, 10.0 K/min.Friedman279.2Volatilization of sulfur
6Pemsler [58]70–100, 100–140, 140–200, 270–325 meshTGHe
100–250 mL/min
Isothermal
500, 525, 550, 575, 600 °C
None-isothermal: 2, 5, 10, 20 °C/min
Shrinking Core Model/First-Order226–239Chemical Reaction Control
7Charpentier [62]<63 μmTGA/DSCAr
50 mL/min
2–100 °C/minFiredman250–350
(average 283 kJ/mol)
Chemical Reaction Control
Table 3. Pyrite pyrolysis—diffusion controlled mechanism (low Ea).
Table 3. Pyrite pyrolysis—diffusion controlled mechanism (low Ea).
No.AuthorParticle SizeEquipmentAtmosphereHeating RateModelEa Value, kJ/mol Mechanism
8Lv [63]63–75 μmTGN2,
3000 mL/min
Isothermal
675–800 °C
Shrinking/3D diffusion103Step 1: 675–725 °C, Surface desulfurization
Step 2: 725–800 °C, Diffusion desulfurization
93
9Fegley [27]sliceHigh-Temperature TubeCO2Isothermal,
390, 416, 468, 500, 530 °C
2.5–7 cm/min
Linear dynamics,
Shrinking core
142Desorption of sulfur
100 ppm
CO-CO2
156
Ar-CO2120
1.1% CO-CO2153
CO-CO2-SO2141
10Boyabat [64]0.425–1.4 mmTubeN2
1670 cm/min
Non-Isothermal
400–800 °C
Shrinking core113Heat transfer at low-temperature
Mass transfer at high-temperature
96
11Udintsev (as cited from [25]) Vacuum, ArIsothermal
400–750 °C
Linear kinetics110
12Schwab [65]0.01–0.1 mm Air, CO2, H2400–650 °CLinear kinetics125–138
13Zhukovskii (as cited from [16,25]) Weight LossVacuum, N2Isothermal
450–690 °C
Linear kinetics110
14Samal (as cited from [16,25]) Weight LossVacuumIsothermal
486–554 °C
Linear kinetics120
Table 4. Pyrolysis with two-step or transitional kinetics.
Table 4. Pyrolysis with two-step or transitional kinetics.
No.AuthorParticle SizeEquipmentAtmosphereHeating RateModelEa Value, kJ/mol Mechanism
15Jovanović [68]0.072 mmthermobalanceN2Isothermal
600, 660, 700, 750, 800, 850 °C
Farrar-Smith272Initially chemical reactions, later diffusion.
Diffusion control equation126
16Hoare [53]0.58 mg
220 mg particles
Stanton-Redcroft TG761/Cahn RGN2
35 mL/min and 200 mL/min
3 °C/min
non-isothermal
Shrinking core, Ginstling-Brounshtein286Step 1: Chemical reaction
Step 2: Solid-state diffusion
190
17Zhang [51]0.058 mmtubularN2
1 L/min
10 °C/min, 450, 500, 600, 700 °CCoats-Redfern199.76600–700 °C
Microwave oven60 °C/min172.62500–600 °C
18Lambert [37]210 × 250 μm
44 × 53 μm
TGAVacuum Shrinking222S2 molecule form
139Desorption of S2 molecule
19Luganov [69]<0.1 mmThermal AnalyzerAr
8–10 L/min
10 °C/min
20–900 °C
Firedman193–210Chemical control at high temperatures
Diffusion control at low temperature
Table 5. Direct oxidation of pyrite to iron oxides.
Table 5. Direct oxidation of pyrite to iron oxides.
No.AuthorParticle SizeEquipmentAtmosphereHeating RateModelStep1 Ea
kJ/mol
Step2 Ea
kJ/mol
Pathway
1Hong [25]2 × 1 × 0.05 cm
slice
Vertical tubeO2-CO2,
75 mL/min
Isothermal
400–590 °C
Linear
dynamics
392–460 °C
82 ± 52
484–538 °C
293 ± 52
Oxidized layer hinders reaction under 460 °C
CO2, CO-CO2,
75 mL/min
260–275Rate constants similar with inert gases
2Aracena [71,76]12.3, 16.0, 22.7, 33.8 μmVertical tubeO2, 5.07–28.69 kpa,1000 mL/minIsothermal
550–800 °C
Initial slope70.1 kJ/mol/Temperature accelerates the first stage; Oxygen pressure impacts on second stage
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

Liu, M.; Xie, G.; Sha, J. Unified Phase Diagram and Competition-Coupling Mechanism for Pyrite Thermal Transformation. Minerals 2025, 15, 1139. https://doi.org/10.3390/min15111139

AMA Style

Liu M, Xie G, Sha J. Unified Phase Diagram and Competition-Coupling Mechanism for Pyrite Thermal Transformation. Minerals. 2025; 15(11):1139. https://doi.org/10.3390/min15111139

Chicago/Turabian Style

Liu, Mingrui, Guangyuan Xie, and Jie Sha. 2025. "Unified Phase Diagram and Competition-Coupling Mechanism for Pyrite Thermal Transformation" Minerals 15, no. 11: 1139. https://doi.org/10.3390/min15111139

APA Style

Liu, M., Xie, G., & Sha, J. (2025). Unified Phase Diagram and Competition-Coupling Mechanism for Pyrite Thermal Transformation. Minerals, 15(11), 1139. https://doi.org/10.3390/min15111139

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