Dipam Patel,
Jiwon Yu
and
Gyeong S. Hwang
*
McKetta Department of Chemical Engineering, University of Texas at Austin, Austin, Texas 78712, USA. E-mail: gshwang@che.utexas.edu; Fax: +1-512-471-7060; Tel: +1-512-471-4847
First published on 16th September 2025
Reactive oxygen species (ROS) play a critical role in the oxidative degradation of amine solvents for carbon dioxide (CO2) capture, resulting in solvent loss and the formation of harmful byproducts. While hydroxyl radicals (˙OH) or direct oxidation by metal cations have been proposed as potential initiators of the oxidative degradation, the specific ROS involved remains unclear. In this study, we propose that superoxide (O2−) may be the dominant ROS under alkaline conditions during CO2 capture, based on the thermodynamic analysis of reduction reactions. Using quantum mechanical (QM) calculations, we further demonstrate that ferrous iron (Fe2+) complexes, when coordinated to electron-donating ligands such as carbamates and bicarbonates, exhibit significantly lower reduction potentials, making them effective reductants for O2. The large concentration of these ligands in CO2-loaded amine solution may allow for the production of such reductive Fe complexes and in turn facilitate O2 reduction to O2−. Finally, we propose a reaction mechanism, supported by molecular dynamics simulations, where O2− initiates the decomposition of protonated monoethanolamine. These findings offer new insights into the primary ROS involved in the oxidative degradation of aqueous amines and suggest strategies to mitigate solvent degradation in CO2 capture processes.
Oxidative degradation occurs in the presence of O2 from flue gas and solvated transition metals leached from the absorber walls. Previous studies have assumed two possible mechanisms: direct oxidation of amines by Fe3,4 or via the production of hydroxyl radicals (˙OH).5 However, the former pathway would produce a highly unstable hydrocarbon radical which is thermodynamically unfavorable without significant potential biasing. Our preliminary calculations support this, showing a standard potential of less than −1.3 V vs. SHE for the reduction of ferric iron (Fe3+) coupled with amine oxidation. Furthermore, the low solubility of Fe3+ in aqueous systems suggests that the energetic cost for this reaction could be even higher in real systems. Hence, this pathway is highly unlikely to occur without strong chelating agents to modify the environment of Fe3+. The latter mechanism involving the production of ˙OH is also somewhat dubious, as the generation of ˙OH from O2 requires the transfer of 3 electrons and 2–3 protons, depending on pH. This reaction may be readily observed in electrochemical systems with an applied potential and an abundant electron source; however, it is less feasible in aqueous amine-based CO2 capture systems. In such systems, the electron source is a small quantity (<1 M) of dissolved metal atoms. In addition, oxidative degradation occurs in alkaline conditions due to the presence of amines. This alkaline environment results in a scarcity of protons, which hinders formation of a key ROS intermediate, H2O2, as 2 protons must be transferred to O2. The complex interplay between oxygen species, solvated metal ions, and environmental conditions results in a system where the dominant ROS is not obvious.
In this work, we investigate the dominant ROS in aqueous amine solutions for CO2 capture. We conduct a comprehensive thermodynamic analysis of the reduction reactions of potential ROS to evaluate their feasibility under typical operating conditions. Additionally, we examine the role of solvated metal ion complexes and the effect of modifying their ligand environment on the reduction processes by employing quantum mechanical (QM) simulations. Also, we propose a mechanism in which superoxide (O2−) can initiate the oxidative degradation of monoethanolamine (MEA), a benchmark solvent for CO2 capture.6
DFT-based ab initio molecular dynamics (AIMD) were performed in CPMD.13 The exchange functional of Becke14 and the correlation functional of Lee, Yang, and Parr15 were used as they were employed by previous calculations involving O2−.16,17 Norm-conserving Troullier–Martins pseudopotentials18 in the nonlocal form of Kleinman and Bylander19 were used to describe the valence-core electron interaction. A cutoff for the plane-wave basis was set at 45 Ry. The local spin density (LSD) approximation was used to treat the spin-polarized character of the system. A fictitious electron mass of 600 amu and a timestep of 0.12 fs were used to ensure adiabaticity in CPMD. Hydrogen atoms were treated as deuterium to ensure no electronic–ionic coupling. Electronic and ionic dynamics were controlled by a chain of Nose–Hoover thermostats.20–22 Free-energy sampling was performed using the PLUMED package.23 The well-tempered metadynamics algorithm was used to sample the free-energy surface.24,25 Hills with an initial height of 0.94 kcal mol−1 and width of 0.1 Å were deposited at a rate of 10−2 fs−1. ΔT was set to 4500 K, corresponding to T = 300 K and a damping parameter of 15. Molecular dynamics trajectories were analyzed using the MDAnalysis python library.26–28
A static calculation of the transition state for this reaction was performed using the Vienna ab initio simulation package (VASP)29 and the dimer method of Henkelmann.30 This calculation utilized the generalized gradient approximation (GGA) functional of Perdew, Burke, and Ernzerhof (PBE)31 and the Projector-Augmented Wave (PAW) method of Kresse.32
O2 + e− → O2−, E0 = −0.33 V vs. SHE | (R1) |
O2− + 2H+ + e− → H2O2, E0 = 0.92 V vs. SHE | (R2) |
H2O2 + H+ + e− → H2O + ˙OH, E0 = 0.32 V vs. SHE | (R3) |
![]() | (1) |
![]() | (2) |
Using eqn (2), we first show the effect of pH alone on the reduction potentials of each ROS of interest, neglecting concentration effects from the first term. As shown in Fig. 1(a), the reduction potentials of (R2) and (R3) decline significantly with increasing pH, becoming endergonic after pH = 5 and pH = 7, respectively. The requirement of 2H+ to form H2O2 severely limits its production under alkaline conditions. On the other hand, the reduction potential of (R1) remains unchanged with varying pH as it does not involve protonation. As the pH shifts further toward the alkaline range, the equilibrium potentials of all three reactions become comparable at pH ≈ 10, with (R1) becoming the most favorable reaction at higher pH values.
![]() | ||
Fig. 1 (a) Calculated equilibrium potentials for O2 reduction (R1), H2O2 formation (R2), and ˙OH formation (R3) and (b) concentration of ROS at different levels of alkalinity. |
Fig. 1(a) describes ROS generation as a function of pH at a set potential (E) of 0.771 V vs. SHE, similar to assuming an abundant electron supply from an electrode. However, in typical CO2 stripping processes, the electron source is limited to a small concentration of dissolved metal ions. These metals undergo oxidation in the presence of electron acceptors with their own half-cell reactions at certain standard reduction potentials (EM).
M → M+ + e−, E0 = EM | (R4) |
E0 = ER − EM | (3) |
![]() | (4) |
![]() | (5) |
[M] = [M]0 − [O2−] − 2[H2O2] − 3[˙OH] | (6) |
[M]0 = [M] + [M+] | (7) |
![]() | ||
Fig. 2 Equilibrium concentrations of ROS at pH = 10 as a function of Fe complex reduction potential. |
With the dominance of O2− shown, the only unknown remaining is the identity of the moderately reductive metal complexes. While the Cu1+/2+ couple in aqueous media fits within the moderate potential region, Fe is overwhelmingly more abundant in the tank walls. This leads us to speculate that some Fe species exist with electronic structures more favorable to oxidation than the pure aqueous case.
The 3d electrons of octahedral Fe2+ complexes are split into the non-bonding t2g set and the anti-bonding eg* set by σ-bonding with the surrounding ligands. The oxidation of Fe2+ requires the removal of an electron from the lower energy t2g set to preserve the net spin, as removal from the eg* set would change the net spin, making the transition spin-forbidden. The energy difference between the t2g set and the semi-occupied HOMO of O2 is the equilibrium potential of (R1), denoted by ΔE.
The right portion of Fig. 3 illustrates the effect of substituting a water ligand with a ligand with free p-orbitals, denoted as the π ligand. In this case, the interaction between the t2g orbitals of Fe2+ and the filled non-bonding p-orbitals of the π ligand interact splitting the normally non-bonding t2g set into bonding t2g and anti-bonding t2g*. The electron is now drawn from these anti-bonding t2g* orbitals when oxidation occurs. The energy increase of these anti-bonding t2g* orbitals directly corresponds to a decrease in reduction potential (ΔEπ), making O2 reduction more thermodynamically favorable.
![]() | ||
Fig. 3 Molecular orbital diagram of π-donation increasing electron energy and decreasing reduction potential. |
In a CO2-loaded amine solution, the available species to serve as π ligands are OH−, HCO3− and amine carbamate, with the latter two species being formed exclusively from CO2 absorption. To quantify the effect of these species on electron transfer, we calculated the equilibrium potential of the Fe2+/3+ couple in various ligand environments.
Table 1 lists the calculated equilibrium potentials of the Fe2+/3+ couple when a single water in an octahedral complex is exchanged for one of the ligands (L = OH−, HCO3−, or MEACOO−). The minimum energy electron configuration is a high spin state for all complexes analyzed here. Note that our calculated potential for the fully aqueous complex [Fe(H2O)6] deviates by only 0.006 V from the experimental value, demonstrating a high accuracy of our calculations.
Structure formula Fe2+(H2O)5L, L = | E0 [V] |
---|---|
H2O | 0.765 |
OH− | 0.211 |
HCO3− | 0.203 |
MEACOO− | 0.226 |
As summarized in Table 1, substitution of a single H2O for any π ligand in the Fe2+(H2O)6 complex results in about 0.5 V more favorable reduction potential, indicating that ligand (L) substitution enhances the ability of Fe complex to reduce O2. The donation effect illustrated in Fig. 3 is further corroborated by comparing the densities of states (DOS) of each complex as shown in Fig. S1. The modified complexes show a shift higher in energy of the frontier orbitals (located just below −5 eV vs. vacuum), indicative of higher energy (more reductive) electrons. Furthermore, charge decomposition analysis of the MEACOO− complex (see Fig. S2) reveals that ligand orbitals contribute significantly to the elevation of the Fe frontier molecular orbitals.35,36
Thermodynamically, the energy required to oxidize the Fe2+ ion decreases due to the complex becoming less stable in its reduced form. To reach such an elevated state, a driving force is required for substitution. Considering that the concentration of OH− is small at realistic pH, only 10−4 at pH = 10 for example, OH− substitution is unlikely to occur. On the other hand, the concentration of HCO3−/MEACOO− can exceed several molar. The large concentration of such ligands can provide the necessary driving force to produce an appreciable number of substituted complexes. In essence, large concentrations of HCO3−/MEACOO− shift the mean ligand environment of Fe2+ cations towards a more reductive potential. In this way, these complexes can transfer an appreciable quantity of electrons to O2, producing ROS (O2−) and causing oxidative degradation. Indeed, experiments have shown increased rates of oxidative degradation in CO2-loaded solutions compared to unloaded solutions.5
For completeness, we have also computed reduction potentials for di-substituted complexes, which are presented in Table S2. Further substitution results in a compounding of the ligand effect, shifting the reduction potential to negative values vs. SHE. However, these complexes are rather unlikely to form considering the concentration of Fe complexes and the inherent energy cost associated with forming such a reductive complex.
We first used a static dimer method transition-state search with an implicit solvent model to evaluate the enthalpic barrier for the reaction of O2− with MEAH+. As shown in Fig. 5, in the transition state (TS), the Cα of MEA becomes a trigonal bipyramidal five-coordinate center with NH3 and –OO in the polar positions. The N–Cα–O angle is 162°, close to the idealized linear arrangement; this is a characteristic of an SN2 reaction where the formation of the Cα–O bond and the breaking of the Cα–N bond occur simultaneously. The Cα–O–O angle, 116°, indicates a nucleophilic attack by the frontier orbitals of O2−, which are skewed relative to the O–O axis. The partial charges on all states in this calculation are denoted in Table S4. The trend from initial to final state is indicative of a charge transfer from the bottom O to N, as expected in the formation of NH3 from a protonated amine. The enthalpic barrier, representing the energy difference, without entropy, between this TS complex and the reactants, is predicted to be 21.7 kcal mol−1.
![]() | ||
Fig. 5 The transition state geometry of the O2− attack on the Cα of MEAH+ obtained from the dimer method. Blue, grey, red, and white balls represent N, C, O, and H respectively. |
We also conducted well-tempered metadynamics simulations using an explicit solvent model to evaluate the free energy barrier while accounting entropic contributions at finite temperatures. 30H2O, 1 MEAH+, and 1O2− molecules are placed in a cubic simulation box with an edge length of 9.5 Å and periodic boundary conditions, corresponding to about 10 wt% of MEA solution. The system was equilibrated for 5 ps, confirming the stable presence of MEAH+ and O2−. As shown in Fig. 6, a single collective variable (CV) of dC–OO − dC–N was employed to mimic the SN2 reaction involving simultaneous Cα–O bond formation and Cα–N bond cleavage. For further verification, we also conducted metadynamics simulations using 2 CVs, dC–OO and dC–NH3, as shown in Fig. S5. In our simulations, the formation of the C–OO bond was always accompanied by a simultaneous breaking of the C–N bond. Moreover, the observed TS geometries closely resemble the TS structure from static calculations in Fig. 6.
![]() | ||
Fig. 6 Free-energy profile of the reaction in Fig. 5 at T = 313 K and T = 413 K. IS = initial state, TS = transition state, FS = final state. |
The free-energy barrier is predicted to be 28.9 kcal mol−1 at 313 K, which is greater than the enthalpy barrier of 21.7 kcal mol−1. The difference is primarily attributed to the solvation effect of O2−. The free-energy barrier may decrease with increasing temperature, as the solvent effect diminishes (see Fig. S8); that is, as the temperature increases, water molecules solvating O2− become more mobile and less rigidly organized, thereby facilitating the intermolecular reaction.38 Given the relatively moderate barriers, our work highlights that O2−, the most likely ROS, can play a key role in initiating the oxidative degradation of aqueous MEA solvents, especially in CO2 stripping conditions.
• Our thermodynamic analysis demonstrates that ˙OH and H2O2 may be dominant in environments with plentiful protons and electrons. However, under typical CO2 capture conditions, O2− is likely to be the dominant ROS. considering the small quantity of metal cations and its alkaline environment.
• The substitution of a water ligand in Fe(H2O)6 with π ligands, such as OH−, HCO3−, and MEACOO−, can significantly lower the reduction potential by π-backdonation. Also, CO2-loaded solutions, where concentrations of HCO3− and MEACOO− are relatively high and act as ligands in Fe complex, enables the formation of a considerable amount of O2−.
• Our free-energy barrier calculations suggest that oxidative degradation of MEA can be initiated by the attack of O2− on the electrophilic Cα of protonated MEA (MEAH+), accompanied by the release of NH3 via SN2 mechanism.
We emphasize that while other pathways could be proposed, the mechanism here importantly identifies O2− as the initiator of oxidative degradation, which is more reasonable based on our analysis than ˙OH. Our work suggests that this mechanism not only has implications for amine-based CO2-capture but in other cases of ROS-related organic molecule degradation such as biological systems, where Fe2+ in alkaline conditions produces ROS which cause oxidative damage to proteins.39
Supplementary information including further discussion of the system of equations parameters and addtional metadynamics results. See DOI: https://doi.org/10.1039/d5cp02617b.
This journal is © the Owner Societies 2025 |