DOI:
10.1039/D5NJ02543E
(Paper)
New J. Chem., 2025, Advance Article
Theoretical investigations of the catalytic antioxidation mechanism of diarylamine and the coordination effects of Fe(III) and Fe atoms
Received
19th June 2025
, Accepted 25th July 2025
First published on 26th July 2025
Abstract
Diarylampine radical-trapping antioxidants (RTA) are widely used in petrochemical products. Herein, density functional theory calculations are used to investigate the catalytic oxidation inhibition mechanism of the highly effective antioxidant diarylamine (Ar2NH). The coordination effects of Fe(III) and Fe atoms on antioxidation are also investigated. These results indicate that Ar2NH is activated by reacting with ROO˙ and that oxygen may serve as a catalyst to participate in the reaction. Ar2NOR can be decomposed by transferring the β-site hydrogen to the ROO˙ radical, resulting in the formation of Ar2N˙. However, the thermal decomposition of Ar2NOR is difficult. The resulting Ar2N˙ can react with an R˙ radical to regenerate Ar2NH, which reenters the reaction system to complete the catalytic cycle. The detailed catalytic mechanism is feasible, which could explain the high efficiency of Ar2NH preferably. Additionally, the coordination of Fe(III) and Fe atoms could improve the antioxidant performance of Ar2NH by enhancing its radical scavenging ability in practical situations, especially Fe(III).
Introduction
The auto-oxidation degradation process of lubricating oil is extremely likely to occur under service conditions because the oil can be in contact with oxygen easily, which is the major cause of the irreversible deterioration of lubricants. The auto-oxidation process is generally considered to involve the following free-radical chain reactions:1–4 |
Propagation R˙ + O2 → ROO˙
| (2) |
|
ROO˙ + RH → ROOH + R˙
| (3) |
Once a free radical R˙ is generated from a lubricant molecule RH under the effects of light or heat, it can react with oxygen rapidly and produce the corresponding peroxy radical ROO˙. The process is shown in eqn (2). ROO˙ further attacks lubricant molecules (eqn (3)), accompanied by the generation of hydroperoxide (ROOH) as the initial oxidation product; thus, a chain reaction is formed via the two reactions. The initial oxidation products may undergo further oxidation and condensation, leading to unexpected acids, esters, and lactones, which are believed to be responsible for the sludge deposits and bulk insoluble matter in used lubricating oil because of their role in increasing the viscosity of oil.5–7 The deterioration of oil reduces lubricant functionality and decreases its service lifetime. This could contribute to the mechanical failure of moving parts and result in significant economic losses in the long run.8,9 The auto-oxidation process can be inhibited or at least retarded by antioxidants that could break the chain reactions by two approaches: one is trapping free radicals without transforming themselves into reactive intermediates, and the other is decomposing hydroperoxide with the generation of unreactive nonradical products. These antioxidants include zinc dialkyldithiophosphate (ZDDP), cuprous dialkyldithiophosphate (CuDDP), organic molybdenum compounds (OMCs), organic sulfides, hindered phenols, and substituted aromatic amines.10–19
Among the numerous types of antioxidants, diarylamines (Ar2NH) are widely employed as radical-trapping antioxidants (RTA) because of their ashless performance and high efficiency at high temperatures. Especially, there have been mandated reductions in ZDDP in engine oil due to the poisoning effect of phosphorus on exhaust after-treatment catalysts, which limits both phosphorus and sulfur in the formulated passenger car engine oil.12,20,21 These situations have caused diarylamines to receive increasingly extensive attention, and their applications are growing significantly.
It is found that Ar2NH could donate its N–H hydrogen atom to the chain-carrying peroxyl radicals (ROO˙) to retard the oxidative degradation of lubricating oil.22 A long time ago, researchers found that each Ar2NH molecule could trap two ROO˙ radicals, i.e., the stoichiometric factor for inhibition (f) is 2.0 at temperatures below ca. 100 °C; however, the f value increases at temperatures above ca. 100 °C.23 Several mechanisms have been proposed to explain the inhibition process, while most of them have far-reaching significance. For example, Bolsman and co-workers24,25 studied the effects of secondary amines on the inhibition of the oxidation process at a high temperature of 130 °C. They found that the f value could reach 40 for diphenylamine and suggested that the nitroxide radical (Ar2NO˙) is an important intermediate catalyst. Haidasz and co-workers26 also believed that the high stoichiometric factor was owing to the regeneration of the diarylamine during hexadecane autoxidation inhibited by the corresponding diarylnitroxide. The most critical progress for this intriguing chemistry phenomenon was achieved by Jensen et al.,27 who investigated the inhibitory effects of 4,4′-dioctyldiphenylamine on the autoxidation of hexadecane. Their work suggested that the high stoichiometric factor observed for diarylamine antioxidants at elevated temperatures is significant owing to the regeneration of these antioxidants from alkoxylamines. The proposed mechanism for the catalytic inhibition by Ar2NH and Ar2NO˙ is designated as the Korcek Cycle, as depicted in Scheme 1. In this mechanism, Ar2NH first reacts with ROO˙ to form Ar2N˙, which then reacts with another ROO˙ to yield a nitroxide radical (Ar2NO˙). Then, Ar2NO˙ can be captured by carbon-centered radicals (R˙) from the substrate to generate Ar2NOR, which produces the parent amine and ketones through thermal decomposition, resulting in the completion of the inhibition cycle.
 |
| Scheme 1 Mechanism for the catalytic inhibition by Ar2NH and Ar2NO˙ proposed by Jensen et al. | |
Many radical-trapping antioxidants have been proposed in accordance with the Korcek cycle mechanism. Pratt's group investigated radical-trapping antioxidants for nearly two decades, achieving remarkable results in this period. One of their primary objectives is to enhance the efficacy of high-temperature RTA. They aim to accomplish this by incorporating heteroatoms, such as nitrogen,28–31 oxygen,32–34 and sulfur33–35 or substituted alkyl chains,36 into diphenylamine. These strategies establish an optimal balance between antioxidant activity and oxidative stability. Similarly, Yu and co-workers successfully developed sulfur-containing Schiff-base-bridged phenolic diphenylamines (SSPDs) that exhibit enhanced antioxidant properties at elevated temperatures using an innovative intramolecular synergy approach.37 Following that, they elucidated the mechanisms behind radical trapping and the decomposition of peroxyl radicals in Schiff base diphenylamines (SDs), underscoring the multifunctional capabilities of these materials in antioxidant applications.38
It is noteworthy that Pratt's group considered that the existing Korcek Cycle should be a simplified representation of the antioxidant cycle mechanism during their development of innovative methods for assessing RTA reaction activity.39,40 Meanwhile, Coote et al.41 studied the antioxidant mechanism of dialkylamine, which could serve as a kind of light stabilizer, through high-level ab initio molecular orbital theory calculations and found that several mechanisms proposed by predecessors were unreasonable. Nevertheless, to date, the detailed mechanistic picture of the antioxidant process of diarylamines in lubricating oil remains unclear. It is also important to elucidate the oxidation inhibition mechanism for the design and application of diarylamine. Therefore, our objective is to investigate the detailed antioxidant mechanism of diarylamine after being inspired by other researchers.
In fact, metals exist extensively in oil in most practical lubricating situations, which come from equipment wear and tear or inorganic additives in operation, such as antioxidants, anticorrosives, and dispersants. The metals, such as Ag, Al, Ti, Fe, and Cu, exist in lubricating oil as soluble organometallic species or metallic particles; some of these species may remain in the whole system all the time.42–44 Meanwhile, researchers have found that the rate of the auto-oxidation process is different from that under the action of metal for lubricating oil.45 For example, zinc, nickel, and aluminum could act as auto-oxidation inhibiting agents when the peroxide concentration is low,46,47 while iron could accelerate the rate of oil auto-oxidation owing to its catalytic action.45,48–50 However, the effects of soluble copper depend on its concentration and other factors.50,51
Singh et al.48 investigated the aging characteristics of base oil under the situation of containing or not containing diarylamine antioxidants. Their results agree with those of earlier researchers because the acceleration of oil aging in tribology experiments is attributed to the catalytic action of iron. Simultaneously, they found that Ar2NH acted as an effective antioxidant under both tribological oxidation conditions and auto-oxidation processes. It is confusing to us that if the antioxidation performance of Ar2NH is unchanged in the two different environments, the retarding results should be evidently different for the participation of iron in steel-on-steel lubricated sliding. Moreover, we found that the addition of Ar2NH to oil provides an antifriction protective graft and reduces the wear of iron. However, the study was not conclusive on which factor—the radical scavenging action, the masking action of Ar2NH to inhibit the catalytic action of iron, or a combination of the two—is the most important driver of antioxidant action in the friction oxidation process. Spurred by these results, we were curious to see if we could resolve this problem from the perspective of the reaction mechanism. Therefore, we investigated the coordination effects of iron on the antioxidant performance of Ar2NH by utilizing density functional theory (DFT) calculations.
In this study, we first used the DFT method to search for a more believable and elaborate mechanism for the antioxidation cycle of diarylamine (Ar2NH). Then, we attempted to answer another question: How would the reaction mechanism be affected if the iron atom or ferric ion coordinates with Ar2NH? This work is of great significance for understanding the actual antioxidant mechanism of Ar2NH under most practical lubricating conditions, which is essential for designing or improving new types of diarylamine antioxidants.
Computational details
4,4′-Dimethyldiphenylamine (DMDPA) and the CH3CH2CH2OO˙ radical were chosen as the model compounds for the diarylamine antioxidant and peroxyl radical, respectively. Gaussian 09 program52 was used to fully optimize all the structures reported in this study at the B3LYP53 level of theory (298.15 K, 1 atm). The 6-31G(d) basis set was used for geometry optimization. Vibrational frequencies were also subsequently calculated at the same level of theory to ensure each optimized structure as a minimum (Nimag = 0) or transition state (Nimag = 1) and to obtain thermodynamic data. IRC calculations54,55 were used to confirm the connectivity between transition states and their minima. To obtain more reliable relative energies, single-point energy calculations for all optimized structures were performed by expanding the basis set to 6-311+G(2d,2p). The reported energies are based on these single-point energies, including thermal corrections. The solvation energies were calculated for gas-phase optimized geometries using the SMD solvation model56 (hexadecane as a solvent) at the M05-2X/6-31G(d) level of theory,57 which was used for parameterizing the SMD model.56 However, because the consideration of the solvation did not significantly change the relative energies for the Ar2NH activation and Ar2NOR decomposition compared with that from the gas-phase calculation (see Tables S1 and S2 for the energy data in solution), only the gas-phase results at 298.15 K are reported here. The radicals were calculated with an unrestricted wave function. The spin state of the calculated species was optimized to the lowest spin state in energy.
Results and discussion
Activation of Ar2NH
Although many mechanisms have been proposed to explain the antioxidant process of diarylamines, the viewpoint that the first reaction step is the activation of Ar2NH by ROO˙ is not controversial. Therefore, the activation mechanism of Ar2NH was first calculated. In the process, Ar2NH is converted into Ar2N˙ radical via homolysis of the N–H bond upon an attack of the ROO˙ radical (Scheme 2, path 1a). Therefore, the bond dissociation energy (BDE) in Ar2N–H (or ArO–H) has been widely considered an important factor in determining the efficacy of an antioxidant, since the weaker the N–H (O–H) bond, the faster the reaction with free radicals.26 In addition, large amounts of oxygen molecules are present in the system, which may participate in the activation reactions. This situation is also considered (path 1b in Scheme 2). The corresponding reaction energies for paths 1a and 1b are shown in Table S1. Certainly, various radicals exist in the system besides ROO˙, such as alkyls (R˙) and alkoxyls (RO˙). The activation of Ar2NH by these radicals is inevitable in lubricant degradation. However, activation by ROO˙ is still expected to dominate because the steady-state concentration of ROO˙ is much higher than that of R˙ and RO˙ under most conditions of oxidative degradation.4,41,58 Therefore, we discuss here the ROO˙-mediated activation of Ar2NH.
 |
| Scheme 2 Activation of Ar2NH. | |
The computed energy profile for the two pathways (paths 1a and 1b) about the activation of Ar2NH is shown in Fig. 1. In path 1a, pre-reaction complex 11a is formed at first with a relative energy of 6.6 kcal mol−1. Then, ROO˙ radical abstracts the hydrogen atom of the N–H bond in Ar2NH directly via the transition state TS121a with an activation energy barrier of 14.2 kcal mol−1 to form the post-reaction intermediate 21a, which subsequently converts into P1a (a combination of Ar2N˙ and ROOH). Although the reaction energy in path 1a is equal to 0.0 kcal mol−1, the reaction could be considered easy on the dynamics for the low energy barrier. In path 1b, the reaction begins with the formation of 11b, which is a complex of Ar2NH with O2. Then, 11b goes through TS121b featuring a concerted event: the cleavage of the N–H bond and formation of O1–H as well as O2–C1 bonds (see Chemdraw-type structure in Fig. 1 for the atom numbering). This step has a Gibbs free energy barrier of 9.6 kcal mol−1 and generates 21b, in which the shape and aromaticity of the activated phenyl ring change. The newly formed 21b subsequently undergoes an isomerization transition state, TS231bIso, with a Gibbs free energy barrier of 4.2 kcal mol−1 to break its weak N⋯H interaction and give 31bIso. Although the relative energy of 31bIso is slightly higher than that of 21b, such an isomerization is beneficial for the H atom of the dangling O1–H bond to interact with the ROO˙ radical. The reaction of ROO˙ with 31bIso occurs via TS451bH with an activation energy barrier of 18.6 kcal mol−1 to generate 51bH. The hydrogen transfer process is the rate-determining step of the entire pathway. 51bH converts into 61bH by releasing ROOH. Then, 61bH undergoes cleavage of the C1–O2 bond to regenerate oxygen (P1b in Fig. 1). This step occurs via TS781b with a free energy barrier of 1.9 kcal mol−1. This result suggests that oxygen participates in the reaction as a catalyst. We chose singlet oxygen as the reactant owing to its higher activity and triplet oxygen (in P1b) as the product owing to its better stability. By comparing the whole pathways of 1a and 1b, path 1a is more kinetically favorable (14.2 vs. 18.6 kcal mol−1), and path 1b has a significant exergonic character. Consequently, we suggest that oxygen may participate as a catalyst in the activation reaction of Ar2NH by ROO˙, which is a feasible process.
 |
| Fig. 1 Computed energy profile (Gibbs free energies and enthalpies in kcal mol−1) for the activation of Ar2NH by ROO˙ radical without or with the participation of O2 (path 1a in black line and path 1b in blue line). The energies are relative to the energy sum of Ar2NH, ROO˙, and singlet O2. | |
Generation of Ar2NOR
Alkoxyamine is an important intermediate in the process of inhibiting oxidation by Ar2NH. The possible pathways for the formation of Ar2NOR are simulated, as shown in Scheme 3. We only investigate these reactions from the viewpoint of thermodynamics since they are rapid radical reactions. Ar2N˙ could react with O21 to yield Ar2NO˙ accompanied by significant heat release (reaction 2a). In addition, Ar2N˙ may capture ROO˙ to produce Ar2NOOR, which is considered feasible to break the Ar2NO–OR bond to form Ar2NO˙ as well (reactions 2b and 2c). The reaction of Ar2N˙ with ROO˙ is endergonic slightly (2.5 kcal mol−1) from the view of change in Gibbs free energies, while it is exothermic from the point of enthalpies. Comparing the two pathways for the formation of Ar2NO˙, reaction 2a holds an absolute advantage. The generated Ar2NO˙ could convert into Ar2NOR by scavenging R˙ easily (reaction 2d). These results indicate that the generated Ar2NO˙ radicals are likely to be taken away, which is consistent with the experimental results, that is, the sustained low concentration of Ar2NO˙.27 Simultaneously, Ar2N˙ could directly react with RO˙ to yield Ar2NOR (reaction 2e). The reaction is also exergonic. Therefore, the pathways shown in Scheme 3 are likely to take place to produce Ar2NOR with respect to the features of radical reactions and thermodynamics. The pathway is more inclined to follow reaction 2a. The results do not agree with Jensen et al.27 who suggest that the activation product Ar2N˙ could mainly react with another ROO˙ to yield Ar2NO˙, which could further react with R˙ to produce Ar2NOR eventually. The results also illustrate that the formation of Ar2NOR is thermodynamically favored.
 |
| Scheme 3 Generation of Ar2NOR.a aCalculated Gibbs free energies and enthalpies (in parentheses) of the reactions are shown in red numbers. | |
Decomposition of Ar2NOR
Thermal decomposition of Ar2NOR. Jensen et al.27 suggested that Ar2NOR could decompose thermally, leading to the regeneration of Ar2NH, accompanied by the formation of carbonyl compounds. This is the main decomposition pathway. In addition, Ar2NOR could decompose into alkene and hydroxylamine (Ar2NOH), which is transformed to Ar2NO˙ by reacting with ROO˙. The two pathways are simulated, and the results are shown in Scheme 4. Paths 3a and 3b are the one-step formation reactions of Ar2NH and Ar2NOH, respectively. In path 3a, the C1–H hydrogen atom is transferred from the C1 atom to the N atom in Ar2NOR, accompanied by the simultaneous dissociation of the N–O bond (TS3a). In path 3b, the C2–H hydrogen atom is transferred from the C2 atom to the O atom in Ar2NOR, coupling with the dissociation of C–O (TS3b). Compared with reaction 3b, reaction 3a is relatively easy to take place owing to its lower activation barrier and a lot of heat release. This may be because Ar2N–OR cleavage is thermodynamically favored over Ar2NO–R cleavage by 15.2 kcal mol−1 (11.2 kcal mol−1 vs. 26.4 kcal mol−1). However, the calculated results demonstrate that both reactions are difficult to occur for the seemingly insurmountable activation barriers (44.0 and 62.8 kcal mol−1). Meanwhile, the two reactions under laboratory conditions (393.15 K) are calculated, and the activation barriers show no significant change for the two pathways. Meanwhile, the reaction Gibbs free energies are changed to be −53.0 kcal mol−1 for path 3a and −6.9 kcal mol−1 for path 3b. The results indicate that high temperature conditions are beneficial for the two reactions with respect to thermodynamics. However, the two reactions are still difficult to occur in view of the high energy barriers. Thus, a new mechanism for the decomposition of Ar2NOR needs to be further explored.
 |
| Scheme 4 Thermal decomposition of Ar2NOR for generating Ar2NH or Ar2NOH via intramolecular H-atom transfer.a aCalculated Gibbs free energy barriers and enthalpy barriers (in parentheses) and reaction energies are shown in red and blue numbers, respectively. | |
The structures of reactant Ar2NOR and the two transition states, TS3a and TS3b, are analyzed to investigate the origin of difficulty in the two reactions. The results in Fig. 2 show that the structures of TS3a and TS3b changed significantly compared with Ar2NOR, especially the dihedral angle ΦN–O–C1–C2, which changed from 179.4° in Ar2NOR to 116.4° in TS3a and −100.0° in TS3b. The large structural changes should be the reason for the high activation barriers in reactions 3a and 3b. Therefore, we speculate that the decomposition of Ar2NOR is a step-by-step reaction to avoid large structural changes.
 |
| Fig. 2 Geometries of the structures of Ar2NOR, TS3a, and TS3b. The key distances (in Å) and dihedral angles (in °) are shown in red boxes. | |
Decomposition of Ar2NOR by reacting with ROO˙. Coote et al.41 studied the antioxidation cycling mechanism of the hindered alkylamine light stabilizer 2,2,6,6-tetramethylpiperidine (TEMPH) through theoretical calculations and proposed a new mechanism for the degradation of alkoxyamines under the action of ROO˙. Inspired by their work, we proposed mechanisms for the decomposition of Ar2NOR, namely β-H and γ-H mechanisms (Scheme 5). The hydrogens in C1 and C2 are named β-H and γ-H, respectively. The decomposition mechanism, including the participation of ROO˙, involves two steps: (i) H-atom transfer, (ii) cleavage of N–O bond (β-H mechanism) or C–O bond (γ-H mechanism). The computed energy profile for the two mechanisms is illustrated in Fig. 3.
 |
| Scheme 5 β-H and γ-H mechanisms for the decomposition of Ar2NOR. | |
 |
| Fig. 3 Computed energy profile (Gibbs free energies and enthalpies in kcal mol−1) for the decomposition of Ar2NOR by reacting with the ROO˙ radical (paths 4a and 4b, i.e., β-H and γ-H mechanism, are in black line and blue line, respectively). The energies are relative to the energy sum of Ar2NOR and ROO˙. | |
The results in Fig. 3 show that the β-H and γ-H mechanisms start with the coordination of ROO˙ with Ar2NOR, yielding 14a and 14b, whose energies are almost at the same level. Then, intermediate 14a or 14b goes through the transfer of the β-H or γ-H in Ar2NOR from C1 or C2 to ROO˙ to generate intermediate 24a or 24b. The hydrogen transfer process via transition state TS124aH or TS124bH has free energy barriers of 26.5 and 30.1 kcal mol−1, respectively. This is the rate-determining step for paths 4a and 4b. 24a or 24b is transformed into 34a or 34b by separating a ROOH. Then, 34a via a transition state TS344aN–O leads to the cleavage of the N–O bond and the formation of Ar2N˙ and propionaldehyde (P4a) with a very low relative Gibbs free energy barrier, 0.4 kcal mol−1. However, 34b occurs through a C–O bond cleavage transition state, TS344bC–O, to turn into Ar2NO˙ with an elimination of a molecule of propene (P4b) with a Gibbs free energy barrier of 5.2 kcal mol−1. Comparing the two pathways, path 4a is more beneficial with respect to thermodynamics compared to path 4b (the black line is always below the blue line). In the meantime, path 4a is kinetically favored over path 4b owing to its lower activation energy at every step of the reaction. The above results indicate that the suggested β-H and γ-H mechanisms for the decomposition of Ar2NOR are facile and that path 4a is kinetically and thermodynamically preferred. This means that the generation of Ar2N˙ is feasible and that Ar2N˙ could be the main product. Of course, we cannot deny the formation of a small quantity of Ar2NO˙ radicals, which are likely to convert into Ar2NOR by scavenging the alkyl radical (path 2c) for the possibility of path 4b. We also extended our investigation to 4,4′-dioctyldiphenylamine (DODPA), which is a bulkier and industrially relevant diarylamine antioxidant, indicating that the mechanism proposed for DMDPA is also applicable to DODPA (see Table S3).
Regeneration of Ar2NH
The crux of the catalytic antioxidation mechanism is to clarify how Ar2NH is regenerated and then re-enter the catalytic cycle. The experimental results27 indicate that the regeneration of the parent amine comes from alkoxy amine intermediates. Thus, we suppose that the regeneration of Ar2NH occurs via the reactions of Ar2N˙, which is the primary product of Ar2NOR decomposition. Consequently, we calculated four pathways in which Ar2N˙ reacts with four substrates (ROO˙, ROOH, R˙, and R), which exist in the reaction system to yield Ar2NH, as shown in Scheme 6. The reaction mechanism for all four pathways is the hydrogen transfer process. The computed energy profile is shown in Fig. 4.
 |
| Scheme 6 Regeneration of Ar2NH. | |
 |
| Fig. 4 Computed energy profile (Gibbs free energies and enthalpies in kcal mol−1) for the regeneration of Ar2NH (path 5a–5d in green, blue, black and red lines, respectively). The energies are relative to the energy sum of the corresponding reactants. In paths 5a and 5c, all structures are singlets except for 15c and 15a, which are triplets. | |
As shown in Fig. 4, path 5a describes the reaction of Ar2NH with ROO˙ to produce Ar2NH, propylene, and oxygen. The reaction (green line) is impossible because of its excessive activation energy barrier, 38.6 kcal mol−1, although the reaction is exothermic. Path 5d demonstrates that the reaction of Ar2NH with R is also impossible owing to its higher activation energy barrier (39.3 kcal mol−1), which is shown in the red line. Along with path 5b (blue line), the regeneration of Ar2NH starts with the coordination of Ar2N˙ with ROOH to form 15b, which comes with the absorption of heat (4.4 kcal mol−1). 15b goes through a hydrogen-transfer transition state with a Gibbs free energy barrier of 15.8 kcal mol−1, TS125b, leading to 15b, which separates into Ar2NH and ROO˙ (P5b). The reaction is kinetically feasible. However, in view of the thermodynamic feature, path 5b (reaction of ROOH) is reversible. Therefore, path 5b is unfavorable owing to the regeneration of Ar2NH. In path 5c (black line), the coordination of Ar2N˙ with R˙ yields 15c and is endergonic by 7.6 kcal mol−1. Then, intermediate 15c also undergoes a hydrogen-transfer transition state, TS125c, to produce 25c with a free energy barrier of 12.8 kcal mol−1. Intermediate 25c tends to convert into Ar2NH with the release of propene (P5c). The results suggest that path 5c is kinetically favorable compared with path 5b. Moreover, path 5c is significantly exergonic (energy release of −45.1 kcal mol−1). Therefore, the regeneration of Ar2NH is derived from the reaction of Ar2N˙ with R˙ (path 5c), which re-enters the reaction system to trap peroxyl radicals.
Catalytic inhibition mechanism
In summary, we proposed a detailed mechanism for catalytic inhibition by Ar2NH and its derivatives. As shown in Scheme 7, diarylamine, Ar2NH, first captures a peroxyl radical (ROO˙) by donating its N–H hydrogen atom to an ROO˙ to produce hydroperoxide (ROOH) and converts itself into diarylaminyl radical, Ar2N˙. It is noteworthy that oxygen may participate in the reaction, which acts as a catalyst. Afterwards, Ar2N˙ can react with O2 or another ROO˙ to yield Ar2NO˙, which can turn into alkoxydiarylamine (Ar2NOR) readily, of which the reaction with O2 is dominating. Meanwhile, Ar2N˙ can directly react with RO˙ to form Ar2NOR, but the concentration of RO˙ is low. Therefore, the formation of Ar2NOR is mainly through the reaction of Ar2NO˙ with R˙. Next, Ar2NOR can decompose via β-site hydrogen atom to another ROO˙, followed by N–O bond cleavage, leading to the formation of Ar2N˙ and propionaldehyde (β-H mechanism). Ar2N˙ then can react with the R˙ radical to produce propylene and Ar2NH, which will re-center in the cycle and complete the catalytic inhibition process. In addition, the decomposition of part Ar2NOR will likely follow the γ-H mechanism, i.e., transfer their γ-site hydrogen atom to the ROO˙ radical to form Ar2NO˙ and propylene. Ar2NO˙ could also re-enter the catalytic cycle to transform itself into Ar2NOR. This is why the experimentally observed concentration of Ar2NO˙ is maintained at a low level. One complete inhibition cycle at least traps three ROO˙ radicals and two R˙ radicals. In general, one of the features of the inhibition cycle is the decomposition mechanism of Ar2NOR, which we believe occurs through reactions with another ROO˙ rather than thermal decomposition. What's more, we suggested that Ar2N˙ radicals are also quite important intermediates. The catalytic inhibition cycle involves the regeneration of three species: Ar2NO˙, Ar2N˙ and Ar2NH.
 |
| Scheme 7 Detailed mechanism for the catalytic inhibition of Ar2NH in the auto-oxidation process. | |
By analyzing the calculated results, we found that the β-H transfer reaction in the decomposition of Ar2NOR through the β-H mechanism is the rate-determining step for the whole catalytic inhibition pathway. Thus, the –CH3 substituent groups in the 4 and 4′-sites are changed to assess the influence of such substituents on these reactions. The substituted groups and calculated results are shown in Table 1.
Table 1 Relative energies ΔG(ΔH) of stationary points involved in the β-H transfer reactions of various substituted Ar2NOR (energies in kcal mol−1, relative to the energy sum of Ar2NOR and ROO˙)

|
R |
Complex |
Transition state |
Intermediate |
Product |
–CH3 |
8.5(0.0) |
26.5(14.5) |
20.8(10.6) |
11.7(12.4) |
–tBu |
7.5(0.0) |
26.6(14.7) |
20.9(10.5) |
12.2(12.4) |
–OCH3 |
10.4(0.4) |
26.5(14.2) |
20.4(10.2) |
11.6(12.1) |
–NO2 |
7.0(−1.4) |
28.2(16.6) |
22.3(12.8) |
12.7(13.5) |
Comparing the results in Table 1, the activation energy barriers for the reactions of Ar2NOR, which are substituted by electron-donating groups (–CH3, –tBu, and –OCH3) with ROO˙, are almost the same, about 26.5 kcal mol−1. However, the activation energy barrier of the β-H transfer reaction increases when the 4 and 4′-sites are substituted by –NO2 although the change is not obvious, that is 28.2 kcal mol−1. These results could provide some enlightenment that the substituted electron-withdrawing groups are averse to the reactions when compared with the electron-donating groups. Valgimigli and Pratt observed that incorporating electron-withdrawing groups can enhance the oxidation potential of diarylamine antioxidants.59 This modification, however, results in only a modest reduction in reactivity, which is consistent with our findings. Furthermore, researchers found that the substitution of electron-donating groups decreases the N–H bond dissociation energy of aniline, whereas electron-withdrawing groups have the contrary effect.60,61 However, introducing nitrogen atoms into the aromatic rings of the phenolic compounds can render them more stable to single-electron oxidation, thereby enabling the substitution of strong electron-donating groups.59 Generally speaking, we suggested that the substituted electron-donating groups are advantageous for the antioxidant performance of Ar2NH in the catalytic inhibition cycle. The results may suggest the development of new diarylamine antioxidants.
Coordination of Fe(III) and Fe atom
The aforementioned studies confirmed the catalytic antioxidant mechanism of Ar2NH, including three main reaction steps: (i) activation of Ar2NH by reacting with ROO˙, (ii) decomposition of Ar2NOR following the β-H mechanism, and (iii) regeneration of Ar2NH through reaction of Ar2N˙ with R˙. Here, the coordination effects of Fe(III) or Fe atom for the mechanism of catalytic inhibition by Ar2NH are investigated.
First, the possible complex structures of Ar2NH coordinating with Fe(III) or Fe atom are optimized. The most stable isomers (i.e., the lowest relative energy structures) are chosen as the models (Fig. 5). The other optimized isomers with higher energies are provided in Fig. S1. Meanwhile, the corresponding binding energy (energy difference between the metal complex and the energy sum of metal and Ar2NH) is reported in Fig. 5.
 |
| Fig. 5 The most stable geometries of Ar2NH_Fe(III) and Ar2NH_Fe and their key geometrical parameters (distance in Å and angle in °), as well as the binding energies of the metal. The binding energy was calculated using the following equation: Ebinding = E(Ar2NH_Fe(III)) − E(Fe(III)) − E(Ar2NH), where E(Ar2NH_Fe(III)), E(Fe(III)), and E(Ar2NH) represent the electronic energies of Ar2NH_Fe(III), Fe(III), and Ar2NH, respectively. The same is true for the case of the Fe atom. | |
As shown in Fig. 5, Fe(III) coordinates to one benzene ring of Ar2NH to form a η6-complex, Ar2NH_Fe(III), which is a quartet. Compared to Ar2NH, the N–H bond in this complex is slightly elongated from 1.01 Å to 1.02 Å, while the C1–N–C2 angle increases from 129.7° to 132.3°. Additionally, the C1–N and C2–N bonds become asymmetric owing to the interaction with Fe(III), which are equal in Ar2NH (1.40 Å). In the case of the Fe atom, a multiple–coordinate complex was located where the Fe atom interacts with one benzene ring and a C atom of another benzene moiety in Ar2NH. In Ar2NH_Fe, the vertical distance between the Fe atom and the centroid of the coordinating benzene ring is 1.48 Å. Compared with Ar2NH, the C1–N bond length increases from 1.40 Å to 1.44 Å, while the C1–N–C2 angle decreases from 129.7° to 114.2°. It is noteworthy that the binding between Fe(III) or Fe atom and Ar2NH could be strong, as suggested by the binding energies shown in Fig. 5. This result indicates that diarylamines are facile to chelate with Fe(III) or Fe atom and prefer the relatively electron-deficient Fe(III). This makes us further investigate the reactions of Fe(III) or Fe atom coordinating with Ar2NH.
Coordination effects of Fe(III) and Fe atom for activation of Ar2NH
First, the coordination effects of Fe(III) and Fe atoms on the activation process of Ar2NH by reacting with ROO˙ have been studied. As shown in Fig. 6, the activation mechanism is considered a hydrogen atom transfer process in which diarylamines donate their N–H hydrogen atom to an ROO˙ radical. Compared with bare Ar2NH, the metal complex-mediated reaction becomes more kinetically feasible, as suggested by their lower activation free energy barrier (14.2 vs. 6.5 and 2.5 kcal mol−1). In the case of the Fe(III) complex, the pathway starts with the formation of a quite stable pre-reaction complex 11aFe(III), accompanied by heat release. Then, 11aFe(III) goes through a hydrogen atom transfer transition state, TS121aFe(III), to yield intermediate 21aFe(III) with a low energy barrier (2.5 kcal mol−1). Although 21aFe(III) is slightly less stable than 11aFe(III), even higher in energy than TS121aFe(III), the accumulation of 11aFe(III) is beneficial to the generation of 21aFe(III). In addition, the separation reaction of post-reaction complex 21aFe(III) needs to absorb heat, which converts 21aFe(III) into P1aFe(III) (Ar2N˙_Fe(III) and ROOH), which means that the products are inclined to exist as complex 21aFe(III). However, there are several other radical species in the system, such as the RO˙ radical. Therefore, 21aFe(III) could release a ROOH accompanied by RO˙ attack to form Ar2NOR_Fe(III), which is thermodynamically favorable. A similar process is also obtained for the case of Ar2NH_Fe, which exhibits a higher free energy barrier than that for Ar2NH_Fe(III) (6.5 vs. 2.5 kcal mol−1), finally leading to P1aFe (Ar2N˙_Fe + ROOH) or Ar2NOR_Fe after an attack by the RO˙ radical. Obviously, compared with bare Ar2NH, the coordination of Fe(III) or Fe atom makes the activation of Ar2NH easier.
 |
| Fig. 6 Computed energy profile (Gibbs free energies and enthalpies in kcal mol−1) for the activation of Ar2NH (black line) and its metal complexes (Fe(III) complex shown in red and Fe complex shown in blue). The energies are relative to the energy sum of the corresponding reactants. 11aFe(III), TS121aFe(III) and P1aFe(III) are quintuplet, while Ar2NORFe(III) is a quartet. | |
Coordination effects of Fe(III) and Fe atoms on the decomposition of Ar2NOR
The aforementioned results about the decomposition of Ar2NOR demonstrated that the β-H mechanism is favorable in both kinetics and thermodynamics, by which Ar2NOR converts into Ar2N˙. The possibility of the γ-H mechanism cannot be neglected although it is not dominant compared with the β-H mechanism. Therefore, decomposition pathways of Fe(III) (7a) or Fe atom (7b) coordinating Ar2NOR complexes are calculated according to both β-H and γ-H mechanisms (Fig. 7)
 |
| Fig. 7 Pathways for the decomposition of Ar2NOR with coordinating Fe(III) (7a) and Fe atom (7b) through β-H and γ-H mechanisms. The energies are relative to the energy sum of the corresponding reactants. All structures are quintuplets, except for TS124aFe(III) and TS124bFe(III), which are septets. | |
As depicted in Fig. 7a, attempts to locate both pre-reaction complexes 14aFe(III) and 14bFe(III) along with β-H and γ-H mechanisms were fruitless. Thus, the two pathways directly start with the transfer of β-site or γ-site hydrogen atoms in Ar2NOR_Fe(III) to ROO˙ and deliver the stable intermediates 24aFe(III) (−32.4 kcal mol−1) or 24bFe(III) (−39.6 kcal mol−1). It is noteworthy that the two processes go through barrierless transition states TS124aFe(III) and TS124bFe(III), respectively. Then, 24aFe(III) involved in the β-H mechanism undergoes its N–O bond cleavage to yield 34aFe(III) with coordinating propionaldehyde. 34aFe(III) is so stable that the dissociation of propionaldehyde is considerably endergonic (109.5 − 36.9 = 72.6 kcal mol−1). However, 24bFe(III) involved in the γ-H mechanism needs to absorb an energy of 21.8 kcal mol−1 to afford 34bFe(III) by releasing ROOH. Then, 34bFe(III) goes through TS344bFe(III) with an energy barrier of 13.1 kcal mol−1 to finally give P4bFe(III) via C–O bond cleavage. It is noteworthy that during the conversion of 34bFe(III) to P4bFe(III), the energy barrier was calculated to be 20.6 kcal mol−1 when considering ROOH coordination. With respect to such an energy barrier for the γ-H mechanism, the decomposition of Ar2NOR_Fe(III) kinetically tends to follow the β-H mechanism.
As shown in Fig. 7b, the decomposition of Ar2NOR_Fe starts with the interaction of ROO˙ with β- or γ-H atom in Ar2NOR_Fe to yield intermediates 14aFe and 14bFe, whose relative energies are almost the same (8.2 and 8.4 kcal mol−1, respectively). Then, 14aFe (or 14bFe) goes through a β-site (or γ-site) hydrogen atom transfer transition state, TS124aFe (or TS124bFe), leading to intermediate 24aFe (or 24bFe), with a Gibbs free energy barrier of 10.7 kcal mol−1 (or 14.8 kcal mol−1). Then, 24aFe is transformed into 34aFe by releasing ROOH. This step is endergonic by 8.4 kcal mol−1. Intermediate 34aFe could undergo N–O bond cleavage via TS344aFe with a low free energy barrier of 0.9 kcal mol−1 to afford more stable intermediate 44aFe, which is further feasibly converted to P4aFe (Ar2N˙_Fe + propionaldehyde). Similarly, 24bFe tends to convert into 34bFe by releasing ROOH. Then, 34bFe undergoes C–O bond cleavage via TS344bFe to afford 44bFe, which further generates P4aFe (Ar2NO˙_Fe + propene) by overcoming a free energy barrier of 6.4 kcal mol−1. Compared with the β-H and γ-H mechanisms shown in Fig. 7b, the decomposition of Ar2NOR_Fe could preferably follow the β-H mechanism because of its lower free energy barrier and more significant exergonic features. Therefore, the coordination of Fe(III) or Fe atom does not change the decomposition pathway of Ar2NOR, which always follows the β-H mechanism. The calculated energy profiles for the decomposition of Ar2NOR with coordinating Fe(III) or Fe atom are shown in Fig. S2.
Coordination effects of Fe(III) and Fe atom for the regeneration of Ar2NH
The above-mentioned results indicate that the regeneration of Ar2NH occurs via the reaction of Ar2N˙ with R˙ (Scheme 6 and Fig. 4). The coordination effects of Fe(III) and Fe atoms on this reaction are investigated here. The calculated energy profile is shown in Fig. 8. Simultaneously, we also calculated other possible regeneration pathways of Ar2NH with coordinating Fe(III) or Fe atom to confirm the most favorable pathway. The results indicate that the coordination of Fe(III) or Fe atom does not change the favorable reaction pathway for the regeneration of Ar2NH (Fig. S3).
 |
| Fig. 8 Computed energy profile (Gibbs free energies and enthalpies in kcal mol−1) for the regeneration of Ar2NH (black line) and its metal complexes (Fe(III) case in red line and the case of Fe atom in blue line). The energies are relative to the energy sum of the corresponding reactants. TS125c and 25c are singlet; 15c, 15cFe, TS125cFe and 25cFe are triplet; and TS125cFe(III) is a quartet. | |
As shown in Fig. 8, the regeneration of Ar2NH goes through a C–H hydrogen atom transfer from R˙ to Ar2N˙ accompanied by the release of propylene with an energy barrier of 12.8 kcal mol−1. After coordinating with the Fe atom, the activation energy barrier is reduced to 9.9 kcal mol−1. The reaction of Ar2N˙_Fe with R˙ is also exergonic although the subsequent discoordination of propylene requires an energy of 7.2 kcal mol−1. The regenerated Ar2NH_Fe could capture ROO˙ to take part in the antioxidation cycle.
It is noted that in the case of Fe(III), attempts to locate reaction intermediates were fruitless along with the regeneration pathway. As shown in Fig. 8, the reaction of Ar2N˙_Fe(III) with R˙ directly goes through a barrierless H-transfer transition state TS125cFe(III) to yield Ar2NH_Fe(III), accompanied by the release of propylene (red line in Fig. 8). It is noteworthy that the resulting ˙CH2CH2CH3 could turn into its isomer, ˙CH(CH3)2, to participate in the reaction. Therefore, the coordination of Fe(III) or Fe atom could accelerate the regeneration of Ar2NH, especially for the Fe(III) case. To sum up, the results indicate that the coordination of Fe(III) or Fe atom could promote the antioxidant process of Ar2NH, including the activation of Ar2NH, decomposition of Ar2NOR, and the regeneration of Ar2NH, especially in the case of Fe(III).
Conclusion
Diarylamines (Ar2NH) are highly effective antioxidants under high temperatures, making them highly employed in lubricating oil over a wide range of operating conditions. The origin of their high efficiency has attracted considerable attention. This study proposed a detailed catalytic inhibition mechanism of Ar2NH through DFT calculations. The results show that the reaction of the ROO˙ radical with Ar2NH leads to the activation of Ar2NH. It is noteworthy that oxygen may serve as a catalyst to participate in this reaction. The key intermediate Ar2NOR could be decomposed via the β-H mechanism by reacting with the ROO˙ radical, resulting in the formation of Ar2N˙ radical. This mechanism differs from the previously proposed thermal decomposition process. The resulting Ar2N˙ could subsequently react with an R˙ radical to regenerate Ar2NH, which reenters the reaction system to complete the catalytic cycle. In addition, the substitution of electron-withdrawing groups on the phenyl ring in Ar2NH may reduce the antioxidant ability of Ar2NH.
The presence of metal in oil is ubiquitous in most practical lubricating systems. These metals could improve the antioxidant performance of Ar2NH. It is reasonably considered that Ar2NH antioxidants are inclined to exist as complexes with a coordinating metal ion or atom. It is found that the coordination of Fe(III) ion or Fe atom could make the catalytic inhibition process of Ar2NH more kinetically feasible, especially for the case of Fe(III) ion, by increasing its radical scavenging ability in practical situations. This study provides a better understanding of the high efficiency of diarylamine in catalytic oil oxidation inhibition.
Author contributions
Junming Wang: conceptualization and writing – original draft. Weiguo Xue: investigation and formal analysis. Zhuozheng Wang: calculation, writing – review and editing. Hao Li: software and formal analysis. Huiying Lv: data curation and validation. Chaoliang Wei: investigation, writing – review and editing. Qingwei Kong: resources, writing – review and editing. Xiaowei Xu: methodology, writing – review and editing. Kebin Chi: investigation, writing – review and editing. Dejun Shi: data curation, writing – review and editing. Yufeng Liu: software, writing – review and editing. Tuanle Li: validation, writing – review and editing. Yi Luo: supervision, project administration, writing – review and editing.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data underlying this study are available in the published article and its SI.
The data supporting this article have been included. See DOI: https://doi.org/10.1039/d5nj02543e
References
- K. U. Ingold and D. A. Pratt, Advances in Radical-Trapping Antioxidant Chemistry in the 21st Century: A Kinetics and Mechanisms Perspective, Chem. Rev., 2014, 114(18), 9022–9046 CrossRef CAS PubMed
. - J. F. Poon and D. A. Pratt, Recent Insights on Hydrogen Atom Transfer in the Inhibition of Hydrocarbon Autoxidation, Acc. Chem. Res., 2018, 51(9), 1996–2005 CrossRef CAS PubMed
. - B. Li and D. A. Pratt, Methods for Determining the Efficacy of Radical-Trapping Antioxidants, Free Radical Biol. Med., 2015, 82, 187–202 CrossRef CAS PubMed
. - J. Helberg and D. A. Pratt, Autoxidation vs. Antioxidants – the Fight for Forever, Chem. Soc. Rev., 2021, 50, 7343–7358 RSC
. - H. E. Jones, D. C. Palacio Lozano, C. Huener, M. J. Thomas, D. J. Aaserud, J. C. Demuth, M. P. Robin and M. P. Barrow, Influence of Biodiesel on Base Oil Oxidation as Measured by FTICR Mass Spectrometry, Energy Fuels, 2021, 35(15), 11896–11908 CrossRef CAS
. - C. Murru, R. Badía-Laíño and M. E. Díaz-García, Oxidative Stability of Vegetal Oil-Based Lubricants, ACS Sustainable Chem. Eng., 2021, 9(4), 1459–1476 CrossRef CAS PubMed
. - Z. Liu, S. Yuan, S. Gong and G. Liu, Long-Term Thermal Oxidative Deposition of RP-3 Jet Fuels: Mechanism and Modeling, Fuel, 2021, 303, 121250 CrossRef CAS
. - E. N. Antonio, C. Wicking, S. Filip, M. P. Ryan and S. Heutz, Role of Iron Speciation in Oxidation and Deposition at the Hexadecane-Iron Interface, ACS Appl. Mater. Interfaces, 2020, 12(16), 19140–19152 CrossRef CAS PubMed
. - J. E. Ruffell, T. J. Farmer, D. J. MacQuarrie and M. S. Stark, The Autoxidation of Alkenyl Succinimides - Mimics for Polyisobutenyl Succinimide Dispersants, Ind. Eng. Chem. Res., 2019, 58(42), 19649–19660 CrossRef CAS
. - K. R. Olson, Y. Gao, A. Briggs, M. Devireddy, N. A. Iovino, M. Licursi, N. C. Skora, J. Whelan, B. P. Villa and K. D. Straub, ‘Antioxidant’ Berries, Anthocyanins, Resveratrol and Rosmarinic Acid Oxidize Hydrogen Sulfide to Polysulfides and Thiosulfate: A Novel Mechanism Underlying Their Biological Actions, Free Radical Biol. Med., 2021, 165(2020), 67–78 CrossRef CAS PubMed
. - Y. Luo, S. Maeda and K. Ohno, DFT Study on Isomerization and Decomposition of Cuprous Dialkyldithiophosphate and Its Reaction with Alkylperoxy Radical, J. Phys. Chem. A, 2008, 112(25), 5720–5726 CrossRef CAS PubMed
. - J. Zhang, J. P. Ewen, M. Ueda, J. S. S. Wong and H. A. Spikes, Mechanochemistry of Zinc Dialkyldithiophosphate on Steel Surfaces under Elastohydrodynamic Lubrication Conditions, ACS Appl. Mater. Interfaces, 2020, 12(5), 6662–6676 CrossRef CAS PubMed
. - J. F. Poon, J. Yan, V. P. Singh, P. J. Gates and L. Engman, Alkyltelluro Substitution Improves the Radical-Trapping Capacity of Aromatic Amines, Chem. - Eur. J., 2016, 22(36), 12891–12903 CrossRef CAS PubMed
. - E. A. Haidasz, D. Meng, R. Amorati, A. Baschieri, K. U. Ingold, L. Valgimigli and D. A. Pratt, Acid Is Key to the Radical-Trapping Antioxidant Activity of Nitroxides, J. Am. Chem. Soc., 2016, 138(16), 5290–5298 CrossRef CAS PubMed
. - M. Griesser, R. Shah, A. T. Van Kessel, O. Zilka, E. A. Haidasz and D. A. Pratt, The Catalytic Reaction of Nitroxides with Peroxyl Radicals and Its Relevance to Their Cytoprotective Properties, J. Am. Chem. Soc., 2018, 140(10), 3798–3808 CrossRef CAS PubMed
. - K. A. Harrison, E. A. Haidasz, M. Griesser and D. A. Pratt, Inhibition of Hydrocarbon Autoxidation by Nitroxide-Catalyzed Cross-Dismutation of Hydroperoxyl and Alkylperoxyl Radicals, Chem. Sci., 2018, 9(28), 6068–6079 RSC
. - C. S. Hanson, M. Donohoe and D. A. Pratt, Enhancement of Diarylamine Antioxidant Activity by Molybdenum Dithiocarbamates, J. Org. Chem., 2023, 88(24), 17420–17429 CrossRef CAS PubMed
. - Z. Tang, Q. Kong, Y. Luo, W. Xue, J. Qu, H. Chen and X. Fu, Theoretical Studies on the Structure and Property of Alkylated Dipenylamine Antioxidants, J. Theor. Comput. Chem., 2014, 13(5), 1–17 CrossRef
. - J. F. Poon, O. Zilka and D. A. Pratt, Potent Ferroptosis Inhibitors Can Catalyze the Cross-Dismutation of Phospholipid-Derived Peroxyl Radicals and Hydroperoxyl Radicals, J. Am. Chem. Soc., 2020, 142(33), 14331–14342 CrossRef CAS PubMed
. - X. Xie, Q. Jia and J. Zhang, Enhanced Anti-Wear Property of Low Viscosity Engine Oil for Sequence IVB Engine Test Meeting the GF-6 Specification, China Pet. Process. Petrochem. Technol., 2019, 21(2), 95–102 CAS
. - Y. Sun, X. Qiu, Y. Liu, S. Sun, C. Zhang, X. Wang, C. Zhao, B. Yu, Q. Yu, M. Cai, F. Zhou, M. Kamal and A. Ali, A Comparative Study of the Tribological Performance of Two Oil-Soluble Ionic Liquids as Replacements for ZDDP (T204) Additives in Lubricants, Tribol. Int., 2024, 198, 109843(1–12) Search PubMed
. - T. Jia, Y. Yu, Q. Liu, Y. Yang, J. J. Zou, X. Zhang and L. Pan, Theoretical and Experimental Study on the Inhibition of Jet Fuel Oxidation by Diarylamine, Chin. J. Chem. Eng., 2023, 56, 225–232 CrossRef CAS
. - J. R. Thomas and C. A. Tolman, Oxidation Inhibition by Diphenylamine, J. Am. Chem. Soc., 1962, 84(15), 2930–2935 CrossRef CAS
. - T. A. B. M. Bolsman, A. P. Blok and J. H. G. Frijns, Catalytic Inhibition of Hydrocarbon Autoxidation by Secondary Amines and Nitroxides, Recl. Trav. Chim. Pays-Bas, 1978, 97(12), 310–312 CrossRef CAS
. - T. A. B. M. Bolsman, A. P. Blok and J. H. G. Frijns, Mechanism of the Catalytic Inhibition of Hydrocarbon Autoxidation by Secondary Amines and Nitroxides, Recl. Trav. Chim. Pays-Bas, 1978, 97(12), 313–319 CrossRef CAS
. - E. A. Haidasz, R. Shah and D. A. Pratt, The Catalytic Mechanism of Diarylamine Radical-Trapping Antioxidants, J. Am. Chem. Soc., 2014, 136(47), 16643–16650 CrossRef CAS PubMed
. - R. K. Jensen, S. Korcek, M. Zinbo and J. L. Gerlock, Regeneration of Amine in Catalytic Inhibition of Oxidation, J. Org. Chem., 1995, 60(17), 5396–5400 CrossRef CAS
. - J. J. Hanthorn, R. Amorati, L. Valgimigli and D. A. Pratt, The Reactivity of Air-Stable Pyridine- and Pyrimidine-Containing Diarylamine Antioxidants, J. Org. Chem., 2012, 77(16), 6895–6907 CrossRef CAS PubMed
. - J. J. Hanthorn, L. Valgimigli and D. A. Pratt, Preparation of Highly Reactive Pyridine- and Pyrimidine-Containing Diarylamine Antioxidants, J. Org. Chem., 2012, 77(16), 6908–6916 CrossRef CAS PubMed
. - R. Shah, E. A. Haidasz, L. Valgimigli and D. A. Pratt, Unprecedented Inhibition of Hydrocarbon Autoxidation by Diarylamine Radical-Trapping Antioxidants, J. Am. Chem. Soc., 2015, 137(7), 2440–2443 CrossRef CAS PubMed
. - J. J. Hanthorn, L. Valgimigli and D. A. Pratt, Incorporation of Ring Nitrogens into Diphenylamine Antioxidants: Striking a Balance between Reactivity and Stability, J. Am. Chem. Soc., 2012, 134(20), 8306–8309 CrossRef CAS PubMed
. - L. A. Farmer, E. A. Haidasz, M. Griesser and D. A. Pratt, Phenoxazine: A Privileged Scaffold for Radical-Trapping Antioxidants, J. Org. Chem., 2017, 82(19), 10523–10536 CrossRef CAS PubMed
. - R. Shah, K. Margison and D. A. Pratt, The Potency of Diarylamine Radical-Trapping Antioxidants as Inhibitors of Ferroptosis Underscores the Role of Autoxidation in the Mechanism of Cell Death, ACS Chem. Biol., 2017, 12(10), 2538–2545 CrossRef CAS PubMed
. - J. F. Poon, L. A. Farmer, E. A. Haidasz and D. A. Pratt, Temperature-Dependence of Radical-Trapping Activity of Phenoxazine, Phenothiazine and Their Aza-Analogues Clarifies the Way Forward for New Antioxidant Design, Chem. Sci., 2021, 12(33), 11065–11079 RSC
. - L. A. Farmer and D. A. Pratt, Non-Tertiary Alkyl Substituents Enhance High-Temperature Radical Trapping by Phenothiazine and Phenoxazine Antioxidants, J. Org. Chem., 2024, 89(9), 6126–6137 CrossRef CAS PubMed
. - R. Shah, J. F. Poon, E. A. Haidasz and D. A. Pratt, Temperature-Dependent Effects of Alkyl Substitution on Diarylamine Antioxidant Reactivity, J. Org. Chem., 2021, 86(9), 6538–6550 CrossRef CAS PubMed
. - S. Yu and S. Liu, Multifunctional Antioxidants with High Activity at Elevated Temperatures Based on Intramolecular Synergism, Eur. J. Org. Chem., 2018, 381–385 CrossRef
. - S. Yu, Y. Wang, S. Wang, J. Zhu and S. Liu, The Antioxidant Activity and Catalytic Mechanism of Schiff Base Diphenylamines at Elevated Temperatures, Ind. Eng. Chem. Res., 2020, 59(3), 1031–1037 CrossRef CAS
. - E. A. Haidasz, A. T. M. Van Kessel and D. A. Pratt, A Continuous Visible Light Spectrophotometric Approach to Accurately Determine the Reactivity of Radical-Trapping Antioxidants, J. Org. Chem., 2016, 81(3), 737–744 CrossRef CAS PubMed
. - R. Shah and D. A. Pratt, Determination of Key Hydrocarbon Autoxidation Products by Fluorescence, J. Org. Chem., 2016, 81(15), 6649–6656 CrossRef CAS PubMed
. - G. Gryn’ova, K. U. Ingold and M. L. Coote, New Insights into the Mechanism of Amine/Nitroxide Cycling during the Hindered Amine Light Stabilizer Inhibited Oxidative Degradation of Polymers, J. Am. Chem. Soc., 2012, 134(31), 12979–12988 CrossRef PubMed
. - Y. Wang, G. Zhang, D. Yang, Y. Li, W. Yin and Y. Weng, Electronic-Control Friction Behavior of Porous Metal-Organic Framework@Ag Nanocrystals as Self-Repairing Lubricant Additive, ACS Appl. Mater. Interfaces, 2023, 15(29), 35732–35740 CrossRef CAS PubMed
. - M. Desanker, B. Johnson, A. M. Seyam, Y. W. Chung, H. S. Bazzi, M. Delferro, T. J. Marks and Q. J. Wang, Oil-Soluble Silver-Organic Molecule for in Situ Deposition of Lubricious Metallic Silver at High Temperatures, ACS Appl. Mater. Interfaces, 2016, 8(21), 13637–13645 CrossRef CAS PubMed
. - M. K. A. Ali, H. Xianjun, L. Mai, C. Qingping, R. F. Turkson and C. Bicheng, Improving the Tribological Characteristics of Piston Ring Assembly in Automotive Engines Using Al2O3 and TiO2 Nanomaterials as Nano-Lubricant Additives, Tribol. Int., 2016, 103, 540–554 CrossRef CAS
. - A. Singh, R. T. Gandra, E. W. Schneider and S. K. Biswas, Lubricant Degradation and Related Wear of a Steel Pin in Lubricated Sliding against a Steel Disc, ACS Appl. Mater. Interfaces, 2011, 3(7), 2512–2521 CrossRef CAS PubMed
. - A. U. C. Maduako, G. C. Ofunne and C. M. Ojinnaka, The Role of Metals in the Oxidative Degradation of Automotive Crankcase Oils, Tribol. Int., 1996, 29(2), 153–160 CrossRef CAS
. - H. Yu, Z. Zheng, H. Chen, D. Qiao, D. Feng, Z. Gong and G. Dong, An Investigation of Tribochemical Reaction Kinetics from the Perspective of Tribo-Oxidation, Tribol. Int., 2022, 165(2021), 107289 CrossRef CAS
. - A. Singh, R. T. Gandra, E. W. Schneider and S. K. Biswas, Studies on the Aging Characteristics of Base Oil with Amine Based Antioxidant in Steel-on-Steel Lubricated Sliding, J. Phys. Chem. C, 2013, 117(4), 1735–1747 CrossRef CAS
. - H. Song, A. Casey, J. Tory, D. Coultas, D. Lester, P. Scott and N. J. Rogers, Biodiesel Promotes Iron-Catalyzed Oxidation of Engine Lubricating Oil, Ind. Eng. Chem. Res., 2023, 62(23), 9054–9061 CrossRef CAS
. - X. Zhang, X. Huang, J. Li, Z. Tang and J. Wang, Thermal Oxidation of Aviation Lubricating Oil: Mechanism, Influencing Factors, Evaluation Methods, and Antioxidants, Asia-Pac. J. Chem. Eng., 2024, 19(5), e3114(1–12) Search PubMed
. - J. M. Lukic, S. B. Milosavljevic and A. M. Orlovic, Degradation of the Insulating System of Power Transformers by Copper Sulfide Deposition: Influence of Oil Oxidation and Presence
of Metal Passivator, Ind. Eng. Chem. Res., 2010, 49(20), 9600–9608 CrossRef CAS
. - M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 09, Revision A.02, Gaussian, Inc., Wallingford CT, 2016. https://www.gaussian.com/g_tech/g_ur/m_citation.htm Search PubMed
. - A. D. Becke, Density-Functional Thermochemistry. III. The Role of Exact Exchange, J. Chem. Phys., 1993, 98(7), 5648–5652 CrossRef CAS
. - K. Fukui, Formulation of the Reaction Coordinate, J. Phys. Chem., 1970, 74(23), 4161–4163 CrossRef CAS
. - K. Fukui, The Path of Chemical Reactions - the IRC Approach, Acc. Chem. Res., 1981, 14(12), 363–368 CrossRef CAS
. - A. V. Marenich, C. J. Cramer and D. G. Truhlar, Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions, J. Phys. Chem. B, 2009, 113(18), 6378–6396 CrossRef CAS PubMed
. - A. Karton, R. J. O’Reilly, D. I. Pattison, M. J. Davies and L. Radom, Computational Design of Effective, Bioinspired HOCl Antioxidants: The Role of Intramolecular Cl+ and H+ Shifts, J. Am. Chem. Soc., 2012, 134(46), 19240–19245 CrossRef CAS PubMed
. - J. Pfaendtner and L. J. Broadbelt, Mechanistic Modeling of Lubricant Degradation. 1. Structure - Reactivity Relationships for Free-Radical Oxidation, Ind. Eng. Chem. Res., 2008, 47(9), 2886–2896 CrossRef CAS
. - L. Valgimigli and D. A. Pratt, Maximizing the Reactivity of Phenolic and Aminic Radical-Trapping Antioxidants: Just Add Nitrogen!, Acc. Chem. Res., 2015, 48(4), 966–975 CrossRef CAS PubMed
. - A. Vagánek, J. Rimarcík, M. Ilcin, P. Škorna, V. Lukeš and E. Klein, Homolytic N–H Bond Cleavage in Anilines: Energetics and Substituent Effect, Comput. Theor. Chem., 2013, 1014, 60–67 CrossRef
. - D. A. Pratt, G. A. DiLabio, L. Valgimigli, G. F. Pedulli and K. U. Ingold, Substituent Effects on the Bond Dissociation Enthalpies of Aromatic Amines, J. Am. Chem. Soc., 2002, 124(37), 11085–11092 CrossRef CAS PubMed
.
Footnote |
† Qingwei Kong's current address: The No. 2 Middle School of Sanhe, Sanhe 065200, China. |
|
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.