Next Article in Journal
Investigation of the Microstructure and Properties of Cage-Shaped Hollow Cathode Bias Voltage Modulated Si-Doped DLC Thick Film
Previous Article in Journal
A Study on the Process of Electroplating a Ni Coating on a Carbon Fiber Surface Under the Condition of an Electroplating Solution Without Additives
Previous Article in Special Issue
Paint Removal Performance and Sub-Surface Microstructural Evolution of Ti6Al4V Alloy Using Different Process Parameters of Continuous Laser Cleaning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Effect of Hetero Interstitial Atoms (C/N/O) on the Thermodynamic Stability in BCC Fe: A DFT Study

1
School of Resource Environment and Safety Engineering, College of Mechanical Engineering, Key Laboratory of Hunan Province of Equipment Safety Service Technology Under Extreme Environment, University of South China, Hengyang 421001, China
2
School of Physics, Dalian Maritime University, Dalian 116026, China
3
Light Alloys Research Institute, Central South University, Changsha 410083, China
*
Authors to whom correspondence should be addressed.
Coatings 2025, 15(8), 929; https://doi.org/10.3390/coatings15080929
Submission received: 13 July 2025 / Revised: 4 August 2025 / Accepted: 5 August 2025 / Published: 8 August 2025

Abstract

Laser cladding rapid solidification technique is an effective strategy for manufacturing ultra-high-strength martensitic stainless steels (UHS-MSS). Due to super-saturation solution strengthening of interstitial atoms (IAs), martensitic stainless steels containing IAs exhibit excellent ultra-high strength and toughness and have high tolerance for oxygen impurities. Hence, studying the specific speciation and structural characteristics of IAs is of great significance for guiding laser cladding of ultra-high-strength steels. Herein, we use density functional theory (DFT) computations to analyze the stable occupancies of IAs and their interactions in body-centered cubic iron (BCC Fe). The findings show that single IAs prefer to occupy octahedral sites over tetrahedral sites. Therefore, octahedral sites are selected as the optimal sites for the following double IAs study. For homo IAs, C-C and N-N configurations exhibit greater stability at long-range distances, whereas O-O demonstrate optimal stability at intermediate distances. Crucially, hetero IAs configurations are more stable compared to single IAs and homo IAs, exhibiting a synergistic effect. Especially, the C-O combination shows the highest stability and strongest bonding character. Meanwhile, the dissociation behavior of O indicates that C-O and N-O have higher dissociation temperatures than single O, further verifying the synergistic effect of hetero IAs. This provides a theoretical basis for understanding the interstitial solution strengthening of laser cladding UHS-MSS.

Graphical Abstract

1. Introduction

Ultra-high-strength martensitic stainless steels (UHS-MSS) have been increasingly used as lightweight structural component materials in aerospace, national defenses, and military industries recently, due to their outstanding comprehensive performances (such as ultra-high strength, excellent toughness, and fatigue properties) [1,2,3]. However, the traditional manufacturing process of UHS-MSS is very complex and costly, which limits its development and application [4,5]. Due to its unique advantages, laser cladding technology has become one of the main technologies for manufacturing various alloys [6,7,8,9,10] in recent years, especially UHS-MSS [11,12]. During laser cladding, interstitial atoms (IAs), carbon (C), and nitrogen (N) are usually considered as beneficial elements because of solution strengthening [13,14,15,16,17], while interstitial oxygen (O) as an impurity [18], is inevitably doped in alloys during the process. Nevertheless, the content of O requires control at an extremely low level in UHS-MSS, e.g., 0.0073 wt.% [19] or 0.001 wt.% [12] in AerMet100, greatly increasing manufacturing difficulty and cost. However, our previous experimental study found that MSS containing C, N, and O IAs, manufactured by laser cladding, have a high tolerance for O impurities and exhibit excellent strength and toughness properties [20,21,22]. Meanwhile, some studies revealed that C/N/O interstitials can form a supersaturated solid solution by laser cladding rapid solidification technique, which provides a highly efficient and low-cost strategy for manufacturing high-strength steel [23,24,25]. Furthermore, IAs form ordered interstitial complexes and play a key role in strengthening various alloy materials, which can be confirmed in the results of Zhao-ping Lu’s research team [26,27,28,29]. They reported that in high-entropy alloys and medium-entropy alloys, oxygen can form ordered oxygen complexes leading to interstitial strengthening, significantly increasing the strength and enhancing the ductility, and ordered nitrogen complexes can also be formed in high-entropy alloys to obtain a high strength and good ductility. Nevertheless, the specific speciation and structural characteristics of IAs remain unclear in UHS-MSS by laser cladding. Therefore, it is urgent to investigate the interstitial strengthening mechanism and thermodynamic behavior of C/N/O interstitials in UHS-MSS at the atomic scale.
First-principles calculation based on the density functional theory (DFT) provides an effective means to obtain the atomic effects of IAs and their interactions at the ab initio level [30,31]. Numerous studies have been dedicated to investigating the effects of IAs by using DFT calculations [13,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48]. Previous studies mainly focused on the stability, solution, diffusion of single IAs, and their interaction with point defects in α-Fe [13,32,35,37,43,44,45]. Ye et al. [13], Shang et al. [35], and Luo et al. [45] studied the migration and occupation of O atoms; some of them believed that O is always stable in the octahedral interstice, but Luo et al. obtained that the tetrahedral site O atom can improve the stability of Fe atoms. Kadowak et al. [43], Ohtsuka et al. [44], and B. Kou et al. [32] investigated the occupation, stability, and diffusion of interstitial C atoms and found that the carbon atoms prefer to exist in the O-site and that interstitial C improves the stability of Fe-Mn-C system. Souissi et al. [37] studied the solution and migration energies of C, N, and O in BCC Fe, and the findings revealed that the calculated values of C and N octahedral interstitial sites are consistent with the experimental data, while the results of oxygen are inconsistent. From the above research results, it can be seen that there is an argument about the site preference of oxygen. Moreover, the stability of double IAs and their interactions also significantly impact solid solution strengthening [49].
Some research has already been conducted concerning the interactions between double IAs and their effects on BCC Fe [34,38,39,41,42]. Barouh et al. [34] reported that the X-X (X = C, N, O) interaction is strongly repulsive when the two solutes are too close, but as the distance increases, the interactions of N–N and C–C become less repulsive, while the interaction of O–O presents an obvious attractive effect at the 4th nn configurations (O–O distance 2.81 Å). He et al. [38] found that the interactions of C–B, B–B, and C–C are all repulsive, C–C is the most repulsive, and the repulsive effect decreases with increasing distance. Lv et al. [39] investigated that O–O and N–N increased the stability of the alloys; the formation energies of these alloys (with N or O) are all negative, with O having a stronger effect than N on the stability of the BCC Fe system. You et al. [41] found that the interstitial C atoms tend to be far away from each other; C–C, O–O, and C–O show repulsive interaction in various configurations of BCC Fe (the largest distance between two IAs is 2 Å); the repulsion is largest for C–O, lowest for O–O, and middle for C–C, respectively. Ahlawat et al. [42] obtained that all IAs are repulsive to each other at shorter distances (two IAs distances within 3 Å), and N-N has the strongest repulsive effect than others, whereas at larger distances, two IAs show attractive interaction. According to their research results, it can be found that two homo IAs show repulsive interaction when they are too close, but there is a slight controversy over the O–O interactions and the repulsive strength between IAs at shorter distances, and the impact of double IAs on the stability of Fe supercells at larger distances is unclear. Moreover, research [38] has found that the equilibrium distance between IAs is around 5 Å. Therefore, it is necessary to comprehensively study the interactions between IAs within 5 Å. In summary, the site preference of single IAs and the influence and interaction of double IAs on the stability of BCC Fe supercells are not fully understood and still need more research, because of their great significance in explaining solid-solution strengthening.
In our work, first-principles calculations were conducted to comprehensively study the thermodynamic stability of IAs (C, N, O) in BCC Fe. Firstly, for single IAs, site preference via lattice parameters, magnetic moments, formation energy, and binding energy was analyzed. Then, diffusion pathways via the climbing-image nudged elastic band (CI-NEB) method were calculated to evaluate the migration characteristics of interstitial atoms. Subsequently, the thermodynamic stability of double homo interstitial pairs (C–C, N–N, O–O) with different configurations was investigated by energy (formation/binding energy) calculations. Next, the effects of hetero interstitial pairs (C–N, C–O, N–O) on lattice stability were discussed, compared with single interstitials and double homo IAs, through energy analysis (formation/binding) and electronic structure analysis (PDOS/charge density difference/Bader charge). Ultimately, the dissociation behavior of O from hetero interstitial pairs (C–O, N–O) was investigated by computing dissociation energies and dissociation temperatures.

2. Calculation Details and Structural Models

2.1. Methods

Density-functional-theory (DFT) calculations were performed using the plane-wave pseudopotential method [50,51] and the Vienna ab initio Simulation Package (VASP) [52,53,54]. The ion–electron interaction was characterized using projector-augmented wave (PAW) potentials [55,56], while the exchange-correlation functionals were provided by the generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional [57]. The Fe54 supercell underwent an energy cutoff convergence test (200–800 eV) to ensure accuracy, as shown in Figure A1 in Appendix A. The results indicated that the system converges when the cutoff energy exceeds 500 eV. The cutoff energy was set at 600 eV to account for computing time and resource conservation. The VASPKIT 1.5.1 [58] software was used to generate K-point paths, and the Gamma center Monkhorst–Pack scheme was employed to sample the K-points in the Brillouin zone [59]. A 6 × 6 × 6 K-point grid was adopted for the BCC Fe lattice and its supercells after K-point convergence tests, as shown in Figure A2. The geometric optimization of supercells was conducted by fully relaxing the cell volume and atomic positions. Then, a single point energy calculation was performed to eliminate Pull stresses further and obtain more accurate energy. The convergence criteria of force and electronic self-consistence were set at −0.02 eV/Å per one atom and 10−6 eV per one atom, respectively.
The climbing image nudged elastic-band (CI-NEB) approach [60,61] was used to calculate the lowest energy paths and diffusion barriers of IAs (C/N/O) in the supercells [25]. Fe, being a transition metal element, is predicted to have a significant spin-polarization effect [42]. As a result, spin-polarized calculations for BCC Fe were performed in the current work since it is ferromagnetic.
The formation enthalpy (ΔH) of IAs X (X = C, N, O) in metals is defined as [62]:
Δ H f = Δ E f = E   F e 54 X n 54 × E ( F e ) n   E   X
where E(Fe54Xn) represents the energies of the supercell with n X atoms, and X could be interstitial atom C, N, or O. E(Fe) represents energy per one atom of the perfect BCC Fe, E(C) represents the energy per one C atom in the graphite solid, with a value of −9.213 eV/atom, and E(O)/E(N) represents half energy of a gas-phase O2/N2 molecules in vacuum, with a value of −4.931 eV/−8.305 eV. At T = 0 K and p = 0 Pa, the formation enthalpy equals the calculated formation energy (ΔEf), i.e., ΔHf = ΔEf, ignoring the zero-vibration contribution [63,64]. A positive formation energy denotes an endothermic process, whereas a negative value indicates an exothermic process, i.e., the formation of interstitial atomic configurations is unfavorable or favorable.
The binding energy is utilized to evaluate the stability of the IAs model. For a supercell containing C/N/O interstitial (Xi), the binding energy can be calculated by the equation:
E b X 1 , X 2 , , X n = i = 1 n E ( X i ) E ( X 1 + X 2 + + X n ) + ( n 1 ) E 0
where E(Xi) represents the energy of the Fe supercell containing Xi alone, E (X1 + X2 + … + Xn) represents the energy of the Fe supercell having n IAs, and E0 represents the energy of the Fe54 supercell. Herein, a positive binding energy indicates attraction between these entities, whereas a negative binding energy denotes repulsion.
This work defines the charge density difference (Δρ(Fe54Xn)) as the difference between the charge density of the n X-containing BCC Fe system, the BCC Fe system, and the X atom [45,65], in Equation (3):
Δ ρ = ρ ( F e 54 X n ) ρ ( F e 54 ) i = 1 n ρ ( X i )
where ρ(Fe54Xn) denotes the total charge density of the system, ρ(Fe54) denotes the charge density of the Fe54 supercell, and ρ(Xi) denotes the charge density of IAs Xi (Xi = C, N, O) at the corresponding position in the system.

2.2. Models

Martensitic stainless steels are conventionally modeled as body-centered cubic Fe (BCC Fe) for theoretical calculations [39,46,66]. Hence, the research model is conducted based on the BCC Fe. In our study, a body-centered cubic iron (BCC Fe) structure (Im3m space group) was established, with lattice constant a = b = c = 2.8664Å, and α = β = γ = 90°. A 54-atom supercell (3 × 3 × 3 unit cell) was adopted to calculate BCC Fe. It was used to investigate the stability of IAs (C, N, O) and the interactions in order to determine their influence on thermodynamic stability in BCC Fe. Because of its relatively small radii, C, N, and O tend to occupy the interstitial positions in the iron lattice [25]. One C/N/O atom was located on the interstice of Fe54 to construct the supercell of Fe54C/Fe54N/Fe54O. Figure 1 shows the structures of the Fe54 supercell with C, N, and O atoms occupying octahedral sites (O-sites) and tetrahedral sites (T-sites), respectively.

3. Results and Discussion

3.1. Stability of Single Interstitial C/N/O Atoms in BCC Fe Supercell

3.1.1. The Analysis of the Formation Energy of Single Interstitial C/N/O Atoms in BCC Fe Supercell

Table 1 shows the calculation of the formation energy (ΔEf) for supercells with a single IA to assess their thermo-stability. The results indicate that single IAs (C/N/O) in BCC Fe exhibit lower formation energy and higher binding energy at O-sites compared to T-sites, which results in greater thermodynamic stability of single IAs at octahedral positions. Moreover, the positive formation energy of Fe54C indicates that interstitial C is more difficult to form into a solid solution compared to N/O. Notably, the Fe54O-Oct exhibits the lowest formation energy. The stability order for octahedral interstitial systems is O > N > C.

3.1.2. Effect of IAs on Lattice Constants and Magnetic Properties

To examine the effect of IAs on lattice constants, geometrically optimized BCC Fe with C/N/O IAs occupying O-sites and T-sites was used. The optimized lattice parameters are listed in Table 1. The lattice constants (a, b, c) and unit cell volume (V) relative to the pure Fe54 supercell reference system (abbreviated as ΔV) for both interstitial configurations after optimization were compared. Positive ΔV values for both tetrahedral and octahedral interstitial configurations demonstrate that structures incorporating C/N/O IAs exhibit volumetric expansion compared to the pure BCC Fe system. In Fe54 systems with interstitial C/N/O atoms, octahedral occupation induces significantly larger volumetric changes than tetrahedral occupation. Especially, the octahedral configuration displays large lattice constant deviations along the c-axis direction but approaches near-isotropic expansion behavior along the <001> crystallographic plane, leading to the transformation of the lattice into a body-centered tetragonal (BCT) lattice structure. It suggests that O-site occupation of C/N/O atoms represents an energetically favorable configuration compared to the T-site in BCC Fe, which is in accordance with the previous energy study in Section 3.1.1. Table 1 shows that IAs (C/N/O) increase the average magnetic moments of the supercells relative to the pure iron, reaching 2.2–2.3 μB, which agrees well with previous results [67]. Taking Fe54C-OCT as an example, we calculate the formation energy convergence uncertainty (±10−4 eV), as shown in Table A1. Meanwhile, it can be seen that the influence of IAs occupying O-sites on the magnetic moment is greater than occupying T-sites. Furthermore, based on the analysis of formation energy, it can be concluded that IAs prefer to occupy O-sites. Therefore, we will focus solely on the effects of IAs occupying O-sites in the following studies.
From the top table of Figure 2, it can be seen that the magnetic moment of iron atoms adjacent to IAs changes drastically compared to the pure iron structure. As shown in Figure 2, for C and N interstitials, both 1NN and 2NN Fe atoms show marked magnetic moment reductions relative to the average value. In contrast, O interstitial displays a significant moment decrease in 1NN Fe atoms, while the 2NN Fe moments remain comparable to the average. This result can be further illustrated by electronic structure analysis.

3.1.3. Verifying Stability of O-Site Interstitial Atom at the Electronic Level

In this section, the electronic structures of the supercell Fe54 with a single IA occupying the O-site are illustrated and compared, and the interaction between IAs and their neighboring iron atoms in BCC Fe is examined to clarify the stability of the supercells. The density of states (DOS), charge density difference, and Bader charge population are employed to describe the chemical bonding and charge transfer of the supercells.
To understand detailed bonding characteristics, the projected density of states (PDOS) of O-site IAs and their neighboring iron atom (1NN Fe and 2NN Fe) in Fe54C, Fe54N, and Fe54O lattices is computed, as indicated in Figure 3. There are no energy gaps near the Fermi level, indicating the metallic properties of the supercells. PDOS shows that the hybridization between Fe (1NN Fe and 2NN Fe) 3d states and C/N/O 2p states is strong in both the Fe-lattices, which is in accordance with other calculations [34,42]. By contrast, it can be seen that the hybridization between Fe 3d and C 2p is stronger than that between N and O, and the Fe-C bond presents significant covalent bond properties. O 2p and Fe 3d exhibit obvious isolated peak hybridization, so the Fe-O bond has distinct ionic properties. The hybridization between C 2p and Fe 3d occurs at −6.4 eV in Fe54C (Figure A3), the hybridization between N 2p and Fe 3d occurs at −7.5 eV in Fe54N, and the hybridization between O 2p and Fe 3d occurs at −8.5 eV in Fe54O. A significant asymmetry in the spin-up and spin-down electron densities reveals the ferromagnetic nature of BCC Fe supercells.
To gain further insight into the charge transfer between C/N/O and their neighbor iron atoms, charge transfer is examined with the aid of charge density difference and Bader charge population. As shown in Figure 4, there is an obvious charge accumulation around the C/N/O atoms, while the charge around their neighboring iron atoms is obviously depleted. Bader charge calculation results also show that the IAs gain electrons, while the neighboring iron atoms lose electrons. It indicates that the charge is transferred from iron to IAs. Because the Bader method is based on atomic volume division for charge calculation, it results in a certain deviation between the Bader charge calculation and the actual charge. The volume of O is the largest, and the calculated charge value is the lowest (−1.020). Moreover, the Bader method has poor correlation with electronegativity, and N charge calculation is overestimated (Table A2). Therefore, the Bader charge results in this paper are used for the comparison of charge transfer in the same supercell between IAs and their neighboring iron atoms, not for the comparison between IAs. The charge density difference results show that the Fe–C bond indicates significant covalency between C and its neighboring Fe atoms, the Fe-N bond implies a strong polar covalent bond due to the large electronegativity difference, and the Fe–O bond shows a typical ionic bond characteristic of a spherical charge distribution. This implies stronger interaction in Fe54O between IAs and Fe atom, but least in Fe54C, which is consistent with the previous energy results in Section 3.1.1.

3.2. The Diffusion of a Single IA in Fe54 Supercell

The diffusion behavior of IAs has a significant impact on the clustering, segregation, and precipitation of solutes within steel [36]. Therefore, it is essential to conduct an in-depth study of the diffusion behavior of IAs. To analyze the diffusion of single IAs in the BCC Fe lattice, the CI-NEB method was employed to optimize the best migration path for these atoms moving from one O-site to a neighboring O-site in the BCC Fe lattice, as illustrated in Figure 5a.
As shown in Figure 5b, in the migration path, the T-site performs as a saddle point. Moreover, according to the calculation of total energy, the IAs (C/N/O) atom is stable in the O-site, but metastable in the T-site. The result is similar to the conclusion drawn from the formation energy calculations (Table 1). The migration energies (Em) for C, N, and O are 0.926 eV, 0.784 eV, and 0.557 eV, respectively. It indicates that O is easiest to migrate and that C is the least. It can be inferred that O tends to diffuse into interstitial sites both within the crystal structure and at its boundaries, thereby affecting the overall structural stability. Based on the above analysis, it can be obtained that the O-site occupation by C/N/O atoms in Fe54 systems represents an energetically favorable configuration compared to T-site occupation.

3.3. The Occupying Sites and Stability of Double IAs in BCC Fe Supercell

3.3.1. Two Homo IAs in Fe54 Lattice

Based on the discussion above (Section 3.1 and Section 3.2), a single IA tends to occupy the O-site. Taking Fe54C2 (with one C in an O-site and another C in a T-site) as an example, the stability of the T-site was tested. The results reveal that C is unstable in a T-site and inevitably reaches the adjacent O-site, as shown in Figure A4. Therefore, we only study the O-sites of double IAs. Moreover, O-sites are classified into two types: the octahedral central interstitial (OCI) site and the edge interstitial (OEI) site. When both C/N/O atoms occupy the OCI sites, there may be five possible configurations as shown in Figure 6a. For double C/N/O IAs, where one occupies an OEI site, there may be seven possible configurations within a single cell, as demonstrated in Figure 6b. The interaction becomes minimal at larger distances; hence the distance between two IAs is less than 5 Å and is considered here [34,42]. The specific energy calculation values and the distances between two homo IAs are shown in Table A3.
As shown in Figure 6, for Fe54C2 and Fe54N2, configuration 11, with two homo interstitial C/N atoms located on the atomic close-packed plane <011> occupying the OEI-site, exhibits the lowest formation energy (1.233 eV/−0.640 eV) and the highest binding energy (0.173 eV/0.092 eV) and is the most stable among the 12 configurations. After structural optimization, the C–C distance is 4 Å, and the N–N distance is 4.5 Å. Configuration 2, with two homo IAs located on the atomic close-packed plane <011> occupying the diagonal OCI-site, is metastable. The positive binding energy of configuration 11 indicates an attractive C–C/N–N interaction, enhancing the stability of BCC Fe. For Fe54O2, configuration 1 possesses the lowest formation energy (−1.895 eV) and highest binding energy (0.329 eV), making it the most stable. After structural optimization, the O–O The distance is 2.8 Å. Configuration 8, where two homo interstitial O atoms occupy the opposite side OEI-site, is metastable. Compared to N–N and C–C, O–O interactions yield a lower formation energy and higher binding energy, signifying stronger attraction and greater system stability. To further probe the influence of IAs on the thermodynamic stability of the Fe54 supercell, the stability of hetero interstitial atom pairs will be investigated in the following section.

3.3.2. Two Hetero IAs in Fe54 Lattice

Figure 6 indicates that configurations 1, 2, 8, and 11 represent relatively stable structures for two homo IAs (C, N, or O). Consequently, these four positions were selected to investigate the occupational preference of two hetero IAs, as illustrated in Figure 7a. In configurations 1 and 2, both hetero IAs occupy OCI sites, while in configurations 8 and 11, both hetero IAs occupy OCI sites.
To assess the thermo-stability of hetero IAs, binding energies between the atoms were analyzed to determine the strength of their mutual interactions, apart from formation energies. Figure 7b shows that the C-O combination exhibits the highest binding energy, whereas C-N has the lowest Eb. The N-O binding energy lies between them. The binding energies of all hetero IAs are positive, indicating the attraction effect of two hetero IAs, enhancing the stability of supercells. Figure 7c shows that configuration 11, with two hetero IAs on the atomic close-packed plane <011>, occupying the OEI-site, demonstrates the lowest formation enthalpy and the highest binding energy, resulting in the highest stability. To further investigate the interactions between IAs and neighboring Fe atoms, electron property and local magnetic moment analyses were performed for configuration 11.
Previous studies indicate that C–O, C–N, and N–O combinations exhibit strong interactions with their neighboring Fe atoms [42]. Electronic structure analysis is conducted to investigate interactions between hetero interstitial pairs and neighboring Fe atoms further. Here, we analyze a typical set of seven neighbor Fe atomic sites that interact with the IAs (Figure 8), as well as the influences of C–O, C–N, and N–O combinations on the electronic distribution and magnetic properties.
The two-dimensional charge density difference plots for Fe54CO, Fe54CN, and Fe54NO (Figure 9) and Bader charge calculation (Table 2) reveal significant charge transfer between the C/N/O atoms and their nearest-neighbor Fe atoms. Charge depletion occurs around the neighboring Fe atoms, while charge accumulation occurs near the C, N, and O atoms. This is evidence of charge transfer from iron to IAs. Notably, the Fe4 atom, which is illustrated in Figure 8, situated between the IAs, exhibits pronounced charge transfer. Given the higher electronegativity of O, a distinct electric dipole forms between Fe and O (Figure 9). The PDOS results for Fe54C in Figure 3a reveal strong hybridization between the C-2p and the surrounding Fe-3d. This hybridization consequently modifies the electron density distribution around neighboring Fe atoms, as demonstrated in Figure 9a. This behavior is distinct from the interactions of C-N and N-O, as presented in Figure 9c,e. A comparison of the bond lengths between C/N/O and their neighboring Fe atoms (Figure 9) and the Bader charge population (Table 2) reveals a negative correlation: the magnitude of charge transfer between C/N/O and their neighboring Fe atoms decreases with shorter bond lengths. Surprisingly, the smallest charge transfer occurs with the first nearest neighbors.
As shown in Table 2, compared to the charge transfer of a single IA in the Fe54C, Fe54N, and Fe54O supercells (Figure 4), the charge transfer of O and C in the Fe54CO, Fe54CN, and Fe54NO systems has increased to a certain extent. This demonstrates a synergistic effect relative to the single IA cases. Notably, the charge transfer magnitudes for both C and O increase simultaneously in the C–O combination, attributed to the exceptionally high electronegativity of O. For the C–N and N–O combinations, the charge transfer increases for C and O, while it decreases slightly for N. This suggests the existence of a charge transfer pathway between C–N and N–O, thereby enhancing bonding between IAs and improving the stability of BCC Fe alloys. Charge transfer and interactions between the IAs (C/N/O) and surrounding Fe atoms have a considerable impact on magnetic moments (Table 2). As a result of hybridization between the IAs 2p orbitals and the Fe atoms 3d orbitals, the magnetic moments of the first nearest-neighbor Fe atoms are significantly reduced compared to the average value of 2.2 μB. Conversely, the magnetic moments of other neighboring Fe atoms increase relative to the average, consistent with spin conservation compensation.

3.4. Synergistic Effects of Hetero IAs by Comparing with the Single and Homo IAs and Studying Oxygen Dissociation Behavior

From the previous analysis, it can be concluded that IAs tend to occupy O-sites, and two homo IAs tend to stabilize at a relatively long distance. The results of the stability study of two hetero IAs site preference indicate that configuration 11 (two hetero located on the close-packed plane <011> occupying the OEI-site) is the most stable (Figure 7). Based on the stable structures, we systematically compare thermodynamic stability and electronic structure across single IAs, homo IAs, and hetero IAs, thereby revealing how synergistic interactions between hetero interstitials enhance the stability of BCC Fe alloys.
The binding energies of two IAs supercells, occupying configuration 11, are all positive, indicating that there are interactions between the two IAs, which is confirmed by Section 3.3. This suggests that the combination of IAs helps to enhance the stability of the BCC Fe supercell. As shown in Table 3, compared to Fe54C and Fe54C2, the formation energy of Fe54CO and Fe54CN is also negative, but it is significantly reduced, especially for Fe54CO, which has a lower formation energy and a positive binding energy, making the structure more stable. This is owing to the strong interaction between C and O. Compared to Fe54O, Fe54NO has a lower formation energy and a higher binding energy, making the structure more stable. Combined with the previous analysis of the migration energy barriers of single IAs, where the migration energy barrier of O is the lowest and that of N is the highest, the strong C–O and N–O bonding not only make the structure more stable but also fix the easily migratory O within the crystal, preventing O from segregating at the grain boundaries.
To further elucidate the bonding characteristics of IAs and their influence on the electronic structure, PDOS analysis is conducted on both single IAs and double IAs supercells, as shown in Figure 3 and Figure 10. Contrasting with single IAs supercells, among the two IAs supercells at configuration 11 (Figure 7), Fe54C2 exhibits a higher DOS near and at the Fermi level, making it less stable. However, multiple hybridization peaks are observed in the −7 to −5 eV range between C 2p and Fe 3d (Figure A3), implying strong covalent bonding characteristics. For Fe54N2 and Fe54O2, N 2p and Fe 3d orbitals form hybridization peaks near −4.5 eV. While O 2p and Fe 3d orbitals form isolated hybridization peaks near −8 eV in Figure 10. Additionally, the DOS at the Fermi level is lower than that of the single-interstitial alloys Fe54N and Fe54O, resulting in a more stable system. Comparing Figure 3 and Figure 10, a significant hybridization phenomenon is observed between C–O, C–N, and N–O, and hybridization is also evident between Fe–C, Fe–N, and Fe–O. It suggests that bonding occurs between hetero IAs, and charge transfer exists between the IAs and their neighboring Fe atoms. This finding is supported by the Bader charge analysis in Table 2. Compared to the effect of single and homo IAs, hetero IAs exhibit lower DOS near and at the Fermi level, implying higher electrochemical stability, which indicates that hetero IAs have a synergistic effect on the stability of the BCC Fe supercell.
According to the investigations on IAs diffusion in Section 3.2, O is more easily diffused compared to C and N. Experimental results indicate that O as an impurity is more easily dissociated from the complexes under high temperature, leading to grain boundary segregation, thereby affecting the mechanical properties of alloy steels [68].
Therefore, the dissociation behavior of O is studied to verify the synergistic effect of hetero IAs (C–O and N–O) by calculating the dissociation energy [25] and dissociation temperature, which is estimated by the Polanyi–Wigner equations [69]. As illustrated in Table 4, the dissolution energy of both C–O and N–O combinations is higher than the migration energy of a single O, indicating that O is more difficult to dissolute from them. To common sense, a higher dissociation energy corresponds to a higher dissociation temperature [48]. The dissociation temperature of C–O (266 K) is the highest among them. This implies that the combination of C–O enhances the thermal stability of the system, verifying the synergistic effect of hetero IAs on the stability of BCC Fe again. In addition, the migration energy of O calculated by DFT is quite different from the experimental value [35]. The reason is that the larger attraction of Fe vacancies to O makes the migration energy of O larger. This paper focuses on the interaction between interstitial atoms and their interaction with neighboring irons and does not consider the effect of vacancy on oxygen migration. The calculated dissociation temperature is used to analyze the synergistic effect trend of hetero IAs. However, the effect of vacancy should be considered in the actual dissociation of oxygen. The migration energy of oxygen with vacancy in reference [37] is calculated, and the corresponding dissociation energy and dissociation temperature are shown in Table A5.

4. Conclusions

The study systematically investigated the thermodynamic stability and atomic interaction mechanisms of interstitial C, N, and O atoms in BCC Fe using DFT calculations. Key findings are summarized as follows:
(1) For single IAs, they prefer occupying O-sites over T-sites. The octahedral configuration displays large lattice constant deviations along the c-axis, resulting in a body-centered tetragonal (BCT) lattice structure. The migration energy barrier of O is the lowest, and that of N is the highest, and the transition state is the tetrahedral T-sites during migration.
(2) For homo IAs, C–C and N–N configurations exhibit greater stability at long-range distance (4.0 Å and 4.5 Å) and obtain the most stable structure on the close-packed plane <011>, whereas O–O demonstrate optimal stability at intermediate distances (2.8 Å). For hetero IAs, C–O, C–N, and N–O all exhibit greater stability at long-range distances (about 4.0 Å). Configuration 11 (two hetero atoms located on the close-packed plane <011> occupying the OEI-site) is the most stable.
(3) The positive binding energy of double IAs systems shows that two IAs have a mutual attraction, particularly C–C, C–O, and N–O. Compared with single and homo IAs, hetero IAs systems exhibit lower DOS near and at the Fermi level, indicating higher electrochemical stability, especially, the C–O combination displays the highest stability and strongest bonding character. In addition, the dissociation temperature of C–O (266 K) and N-O (255 K) is higher than single O. All of the above reveal that the synergistic effect of hetero IAs helps to enhance the stability of the BCC Fe supercell.

Author Contributions

Conceptualization, F.W., T.M. and C.Q.; supervision, F.W., P.C., H.Z., Y.C., P.Z. and R.L.; writing—review and editing, F.W., P.C., H.Z., Y.C., P.Z., R.L. and C.Q.; writing—original draft, F.W.; investigation, T.M.; methodology, T.M.; resources, T.M.; formal analysis, P.C. and C.Q.; funding acquisition, H.Z. and C.Q.; software, P.Z. 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 (Grant No. 51474130); National Natural Science Foundation of China (Grant No. 52375341).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
UHS-MSSUltra-high strength martensitic stainless steels
IAsInterstitial atoms
DFTDensity functional theory
BCC FeBody-centered cubic iron
CI-NEBClimbing-image nudged elastic band
VASPVienna ab initio Simulation Package
PAWProjector-augmented wave
1NNThe first-nearest-neighbor
2NNThe second-nearest-neighbor
DOSDensity of states
PDOSProjected density of states
O-sitesOctahedral sites
T-sitesTetrahedral sites
BCTBody-centered tetragonal
OCI-siteOctahedral central interstitial site
OEI-siteEdge interstitial site

Appendix A

Figure A1. Energy cutoff (ENCUT) convergence test of Fe54, and the red square represents the chosen ENCUT of this paper.
Figure A1. Energy cutoff (ENCUT) convergence test of Fe54, and the red square represents the chosen ENCUT of this paper.
Coatings 15 00929 g0a1
Figure A2. K-point convergence tests of Fe54, and the red square represent the chosen K grid of this paper.
Figure A2. K-point convergence tests of Fe54, and the red square represent the chosen K grid of this paper.
Coatings 15 00929 g0a2
Figure A3. (a) A zoom of PDOS of the C and its 1NN Fe, 2NN Fe in Fe54C; (b) a zoom of PDOS of Fe54C2.
Figure A3. (a) A zoom of PDOS of the C and its 1NN Fe, 2NN Fe in Fe54C; (b) a zoom of PDOS of Fe54C2.
Coatings 15 00929 g0a3
Figure A4. Structure Optimization of Fe54C2 with one C in an O-site and another C in a T-site. (a) is the initial structure of Fe54C2, (b) is the final structure of Fe54C2.
Figure A4. Structure Optimization of Fe54C2 with one C in an O-site and another C in a T-site. (a) is the initial structure of Fe54C2, (b) is the final structure of Fe54C2.
Coatings 15 00929 g0a4
Figure A5. 2D slices of the Electron Localization Function (ELF) in Fe54CO, Fe54CN, and Fe54NO. (a,c,e) are on the <011> plane direction; (b,d,f) are on the <100> plane direction.
Figure A5. 2D slices of the Electron Localization Function (ELF) in Fe54CO, Fe54CN, and Fe54NO. (a,c,e) are on the <011> plane direction; (b,d,f) are on the <100> plane direction.
Coatings 15 00929 g0a5
  • Equation A1
E dis k B T d i s 2 = ν j β e E d i s k B T d i s
where β is the ramp in temperature (the heating rate) and νj is the attempt frequency, and the frequency of oxygen is taken as 1.5 × 1013 Hz by the reference, kB is the Boltzmann constant, Tdis is the dissociation temperature, and Edis is the dissociation energy.
Table A1. Formation energy convergence uncertainty of Fe54C-OCT.
Table A1. Formation energy convergence uncertainty of Fe54C-OCT.
Total Energy (eV)Formation Energy (eV)Convergence Uncertainty (eV)
1−453.362760.75840
2−453.362770.7583−1 × 10−4
3−453.362740.75860.0002
Table A2. The Bader charge and atomic volume of IAs with their neighboring iron atoms in Fe54C, Fe54N and Fe54O.
Table A2. The Bader charge and atomic volume of IAs with their neighboring iron atoms in Fe54C, Fe54N and Fe54O.
AlloyAtomBader ChargeAtom Volume (Å3)
Fe54CC−1.0487.388
1NN Fe0.05610.677
2NN Fe0.14711.079
Fe54NN−1.1627.678
1NN Fe0.09010.716
2NN Fe0.15911.132
Fe54OO−1.0207.965
1NN Fe0.08811.000
2NN Fe0.16911.291
Table A3. Formation energy (∆Ef), binding energy (Eb), and the distances (d) between interstitial atoms of different configurations in Fe54C2/Fe54N2/Fe54O2.
Table A3. Formation energy (∆Ef), binding energy (Eb), and the distances (d) between interstitial atoms of different configurations in Fe54C2/Fe54N2/Fe54O2.
Configurations∆Ef (eV)Eb (eV)d (Å)
Fe54C2Fe54N2Fe54O2Fe54C2Fe54N2Fe54O2C-CN-NO-O
11.276−0.577−1.8950.1460.0290.3292.9152.9132.767
21.250−0.638−1.7420.1720.0900.1764.0023.8903.883
32.1400.414−1.231−0.718−0.962−0.3352.3822.4842.358
41.393−0.481−1.6160.029−0.0670.0503.4833.4993.522
53.1061.120−0.166−1.684−1.668−1.4003.4693.4233.434
62.1405.530−1.218−0.718−6.078−0.3482.3822.4072.354
73.276−0.457−1.052−1.854−0.091−0.5141.5221.9522.370
81.258−0.425−1.7860.164−0.1230.2202.8642.4662.779
91.391−0.538−1.5880.031−0.0100.0223.4843.4173.535
101.355−0.523−1.4960.067−0.025−0.0703.3463.4543.387
111.233−0.640−1.6020.1730.0920.0363.9984.4703.896
121.489−0.366−1.661−0.067−0.1820.0952.6982.6202.551
Table A4. Formation energy (ΔEf) and Binding energy (Eb) of two hetero C/N/O interstitial atoms occupying different configurations of O-sites, as shown in Figure 7a. The units of all energy are eV.
Table A4. Formation energy (ΔEf) and Binding energy (Eb) of two hetero C/N/O interstitial atoms occupying different configurations of O-sites, as shown in Figure 7a. The units of all energy are eV.
ConfigurationsC–OC–NN–O
ΔEfEbΔEfEbΔEfEb
10.0830.1480.6580.082−1.1820.125
20.0530.1780.6150.125−1.210.153
30.0690.1620.6410.099−1.1870.13
40.0420.1890.5990.141−1.2150.158
Table A5. Calculation of dissociation energy (Edis) and dissolution temperature (Tdis) of single O (occupying O-site), C-O, and N-O considering vacancy effect. Note that Edis = Em + Eb is adopted for IA, and the heating rate during dissociation calculation is 1 K/s.
Table A5. Calculation of dissociation energy (Edis) and dissolution temperature (Tdis) of single O (occupying O-site), C-O, and N-O considering vacancy effect. Note that Edis = Em + Eb is adopted for IA, and the heating rate during dissociation calculation is 1 K/s.
Eb (eV)Em (eV)Edis (eV)Tdis (K)
single O-0.897 *-318
CO → C + O0.1890.897 *1.086384
NO → N + O0.1580.897 *1.055373
* Ref. [37].

References

  1. Han, T.F.; Ding, Z.M.; Feng, W.X.; Yao, X.Y.; Chen, F.F.; Gao, Y.S. Effects of Process Parameters on Microstructure and Wear Resistance of Laser Cladding A-100 Ultra-High-Strength Steel Coatings. Coatings 2024, 14, 669. [Google Scholar] [CrossRef]
  2. Yang, Y.X.; Wang, Z.H.; Wang, Q.; Tang, C.Y.; Wan, P.; Cao, D.H.; Dong, C. Microstructure structure and mechanical properties of coherent precipitation strengthened ultrahigh strength maraging stainless steel. Acta Phys. Sin. 2025, 74, 058101. [Google Scholar] [CrossRef]
  3. Guan, S.Z.; Wang, L.L.; Gao, Q.Y.; Fang, Y.C.; Zhan, X.H. Microstructure Evolution and Mechanical Property Enhancement of 30CrMnSiNi2A Steel Repaired by Laser Cladding with AerMet100 Steel. Met. Mater. Int. 2025, 9, 1–14. [Google Scholar] [CrossRef]
  4. Li, J.H.; Zhan, D.P.; Jiang, Z.H.; Zhang, H.S.; Yang, Y.K.; Zhang, Y.P. Progress on improving strength-toughness of ultra-high strength martensitic steels for aerospace applications: A review. J. Mater. Res. Technol. 2023, 23, 172–190. [Google Scholar] [CrossRef]
  5. Wang, Q.P.; Liang, Y.Z.; Chen, X.S.; Yang, Z.W.; Dong, K.W.; Peng, Y.; Zhou, Q.; Wang, K.H.; Kong, J. A novel 2.1 GPa martensitic stainless steel manufactured by laser powder bed fusion and post treatment. J. Mater. Res. Technol. 2025, 36, 1930–1937. [Google Scholar] [CrossRef]
  6. Gao, S.R.; Liu, H.S.; Fang, X.F.; Lu, W.J.; Li, S.X.; Chen, Y.B.; Lu, S.Y. Underlying Mechanism for “Loss of Passivation” Effect of a High-Carbon Martensitic Stainless Steel Coating via Laser Cladding. J. Electrochem. Soc. 2023, 170, 021513. [Google Scholar] [CrossRef]
  7. Song, T.; Chen, Z.; Cui, X.; Lu, S.; Chen, H.; Wang, H.; Dong, T.; Qin, B.; Chan, K.C.; Brandt, M.J.N. Strong and ductile titanium–oxygen–iron alloys by additive manufacturing. Nature 2023, 618, 63–68. [Google Scholar] [CrossRef]
  8. Cui, C.Y.; Xu, H.; Ye, F.Y.; Yang, J.; Cui, X. Microstructure evolution and characteristics of laser cladding with pre-placed powder lightweight refractory NbxMo0.5Ti1.5Ta0.2Cr high-entropy alloy. Intermetallics 2024, 167, 108229. [Google Scholar] [CrossRef]
  9. Mo, K.K.; Cao, L.C.; Zhao, D.K.; Long, Y.; Zhong, J.L.; Zhong, J.K.; Bi, G.J. Study on the microstructure and wear resistance of laser clad martensitic stainless steel 420 with different content of TiC. Mater. Today Commun. 2025, 46, 112615. [Google Scholar] [CrossRef]
  10. Patil, M.; Acherjee, B.; Manna, I. Enhancing wear resistance of railway wheel by laser surface cladding of martensitic stainless steel on high carbon rail steel. Mater. Today Commun. 2025, 42, 111353. [Google Scholar] [CrossRef]
  11. Wang, P.; Zhang, Y.H.; Pang, M.; Gao, L.H. Research of the effects of laser power variation on the properties of laser cladding Ni45/Cr3C2/MoS2 self-lubricating wear-resistant coatings on the surface of 300M ultra-high-strength steel. J. Mater. Sci. 2024, 59, 13558–13574. [Google Scholar] [CrossRef]
  12. Zhan, X.H.; Fang, Y.C.; Gao, Q.Y.; Guan, X.H.; Ci, S.W.; Cai, X.K.; Guan, S.Z. Study on the microstructure evolution performance improvement of Invar alloy repaired by laser cladding Aermet100 ultra-high strength steel. J. Mater. Process. Technol. 2024, 331, 118517. [Google Scholar] [CrossRef]
  13. Ye, Z.; Lei, Y.; Zhang, J.; Zhang, Y.; Li, X.; Xu, Y.; Wu, X.; Liu, C.S.; Hao, T.; Wang, Z. Effects of oxygen concentration and irradiation defects on the oxidation corrosion of body-centered-cubic iron surfaces: A first-principles study. Chin. Phys. B 2022, 31, 086802. [Google Scholar] [CrossRef]
  14. Wang, C.Y.; Luo, K.Y.; Wang, J.; Lu, J.Z. Carbide-facilitated nanocrystallization of martensitic laths and carbide deformation in AISI 420 stainless steel during laser shock peening. Int. J. Plast. 2022, 150, 103191. [Google Scholar] [CrossRef]
  15. Wang, R.; Li, F.; Yu, Z.; Kang, Y.; Li, M.; Hu, Y.; An, H.; Fan, J.; Miao, F.; Zhao, Y.; et al. Influences of partial substitution of C by N on the microstructure and mechanical properties of 9Cr18Mo martensitic stainless steel. Mater. Des. 2023, 236, 112497. [Google Scholar] [CrossRef]
  16. Coha-Vesga, P.M.; Yuuki-Koga, G.; Mendoza-Oliveros, M.E.; Gil-Coury, F.; Mujica-Roncery, L. Corrosion resistance of nitrogenated high- carbon martensitic stainless steel designed and at low. Rev. UIS Ing. 2023, 22, 181–190. [Google Scholar] [CrossRef]
  17. Song, Y.; Lu, Y.; Li, Y.; Sun, H.; Li, H.; Xu, H.; Wang, Y. Effect of nitrogen on the thermal deformation behavior of martensitic stainless bearing steel. Mater. Tehnol. 2024, 58, 809–818. [Google Scholar] [CrossRef]
  18. Lv, N.; Fu, A.; Chen, C.; Bai, H.; Zhao, Z.; Li, X.; Xu, Z.; Meng, G. Influence of oxygen on metastable pitting behavior of super 13Cr stainless steel in CO2-saturated environment. Vacuum 2025, 232, 113845. [Google Scholar] [CrossRef]
  19. Cheng, H.; Liu, D.; Tang, H.-B.; Zhang, S.-Q.; Ran, X.-Z.; Wang, H.-M. Effect of hot isostatic pressing on fatigue properties of laser melting deposited AerMet100 steel. J. Iron Steel Res. Int. 2013, 20, 79–84. [Google Scholar] [CrossRef]
  20. Hu, J.P.; Zhu, H.M.; Zhang, J.W.; Ouyang, M.N.; Qiu, C.J.; Duan, J. Effects of TiC addition on microstructure, microhardness and wear resistance of 18Ni300 maraging steel by direct laser deposition. J. Mater. Process. Technol. 2021, 296, 117213. [Google Scholar] [CrossRef]
  21. Zhu, H.; Ouyang, M.; Hu, J.; Zhang, J.; Qiu, C.J.C.I. Design and development of TiC-reinforced 410 martensitic stainless steel coatings fabricated by laser cladding. Ceram. Int. 2021, 47, 12505–12513. [Google Scholar] [CrossRef]
  22. Zhu, H.M.; Ben, X.; Li, B.C.; Shen, L.Z.; Pan, C.L.; Qiu, C.J. Influence of Al microalloying on microstructure and mechanical properties of laser directed energy deposited AISI 420 steel. Mater. Today Commun. 2025, 42, 111355. [Google Scholar] [CrossRef]
  23. Raabe, D.; Tasan, C.C.; Olivetti, E.A. Strategies for improving the sustainability of structural metals. Nature 2019, 575, 64–74. [Google Scholar] [CrossRef]
  24. He, M.Y.; Shen, Y.F.; Jia, N.; Liaw, P.K. C and N doping in high-entropy alloys: A pathway to achieve desired strength-ductility synergy. Appl. Mater. Today 2021, 25, 101162. [Google Scholar] [CrossRef]
  25. Chen, Y.; Zhu, H.; Zhang, P.; Wang, Z.; Wang, M.; Sha, G.; Lin, H.; Ma, J.; Zhang, Z.; Song, Y.; et al. An exceptionally strong, ductile and impurity-tolerant austenitic stainless steel prepared by laser additive manufacturing. Acta Mater. 2023, 250, 118868. [Google Scholar] [CrossRef]
  26. Lei, Z.F.; Liu, X.J.; Wu, Y.; Wang, H.; Jiang, S.H.; Wang, S.D.; Hui, X.D.; Wu, Y.D.; Gault, B.; Kontis, P.; et al. Enhanced strength and ductility in a high-entropy alloy via ordered oxygen complexes. Nature 2018, 563, 546–550. [Google Scholar] [CrossRef]
  27. Jiao, M.Y.; Lei, Z.F.; Wu, Y.; Du, J.L.; Zhou, X.Y.; Li, W.Y.; Yuan, X.Y.; Liu, X.C.; Zhu, X.Y.; Wang, S.D.; et al. Manipulating the ordered oxygen complexes to achieve high strength and ductility in medium-entropy alloys. Nat. Commun. 2023, 14, 806. [Google Scholar] [CrossRef]
  28. Zhou, X.Y.; Wu, H.H.; Wu, Y.; Liu, X.J.; Peng, X.Y.; Hou, S.; Lu, Z.P. Formation and strengthening mechanism of ordered interstitial complexes in multi-principle element alloys. Acta Mater. 2024, 281, 120364. [Google Scholar] [CrossRef]
  29. Zhao, D.; Yang, Q.; Wang, D.; Yan, M.; Lu, Z.J.V.; Prototyping, P. Ordered nitrogen complexes overcoming strength-ductility trade-off in an additively manufactured high-entropy alloy. Virtual Phys. Prototyp. 2020, 15, 532–542. [Google Scholar] [CrossRef]
  30. Dong, X.; Ma, S.; Wu, X.-X.; Na, R.-S. First principles on bcc and fcc phases structure of Fe52T2 (T=Cr,Mn,Co,Ni) alloys. Cailiao Gongcheng 2019, 47, 147–153. [Google Scholar] [CrossRef]
  31. Leveau, T.; Ventelon, L.; Clouet, E. Interaction of C, N and O interstitial solute atoms with screw dislocations in HfNbTaTiZr. Acta Mater. 2024, 275, 120062. [Google Scholar] [CrossRef]
  32. Kou, B.; Kou, Z.; Liu, Y.; Han, P. First principles study on interstitial C and substitutional Mn in bcc iron: Structural stability and elastic properties. Mater. Res. Innov. 2014, 18, 775–779. [Google Scholar] [CrossRef]
  33. You, Y.; Yan, M.F.; Chen, H.T. Interactions of carbon-nitrogen and carbon-nitrogen-vacancy in alpha-Fe from first-principles calculations. Comput. Mater. Sci. 2013, 67, 222–228. [Google Scholar] [CrossRef]
  34. Barouh, C.; Schuler, T.; Fu, C.-C.; Nastar, M. Interaction between vacancies and interstitial solutes (C, N, and O) in α-Fe: From electronic structure to thermodynamics. Phys. Rev. B 2014, 90, 054112. [Google Scholar] [CrossRef]
  35. Shang, S.L.; Fang, H.Z.; Wang, J.; Guo, C.P.; Wang, Y.; Jablonski, P.D.; Du, Y.; Liu, Z.K. Vacancy mechanism of oxygen diffusivity in bcc Fe: A first-principles study. Corros. Sci. 2014, 83, 94–102. [Google Scholar] [CrossRef]
  36. Zhang, C.; Fu, J.; Li, R.; Zhang, P.; Zhao, J.; Dong, C. Solute/impurity diffusivities in bcc Fe: A first-principles study. J. Nucl. Mater. 2014, 455, 354–359. [Google Scholar] [CrossRef]
  37. Souissi, M.; Chen, Y.; Sluiter, M.H.F.; Numakura, H. Ab initio characterization of B, C, N, and O in bcc iron: Solution and migration energies and elastic strain fields. Comput. Mater. Sci. 2016, 124, 249–258. [Google Scholar] [CrossRef]
  38. He, T.W.; Jiang, Y.H.; Zhou, R.; Feng, J. Point defect interactions in iron lattice: A first-principles study. RSC Adv. 2016, 6, 45250–45258. [Google Scholar] [CrossRef]
  39. Lv, Z.; Fan, J.; Guan, K.; Wu, Z.; Zhao, D.; Fu, W. Effects of interstitial atoms (N/O) in bcc Fe from first-principle calculations. Fusion. Eng. Des. 2018, 137, 22–29. [Google Scholar] [CrossRef]
  40. Razumovskiy, V.I.; Yeddu, H.K.; Spitaler, J. Effect of carbon on elastic properties and microstructure of maraging steel: First-principles and phase-field study. Comput. Mater. Sci. 2019, 162, 1–11. [Google Scholar] [CrossRef]
  41. You, Y.; Yan, M.F.; Yan, J.H.; Sun, G.; Wang, C. First principles study of interactions of oxygen-carbon-vacancy in bcc Fe. Chin. Phys. B 2019, 28, 106102. [Google Scholar] [CrossRef]
  42. Ahlawat, S.; Srinivasu, K.; Biswas, A.; Choudhury, N. First-principle investigation of electronic structures and interactions of foreign interstitial atoms (C, N, B, O) and intrinsic point defects in body- and face-centered cubic iron lattice: A comparative analysis. Comput. Mater. Sci. 2019, 170, 109167. [Google Scholar] [CrossRef]
  43. Kadowaki, M.; Saengdeejing, A.; Muto, I.; Chen, Y.; Masuda, H.; Katayama, H.; Doi, T.; Kawano, K.; Miura, H.; Sugawara, Y.; et al. First-principles analysis of the inhibitive effect of interstitial carbon on an active dissolution of martensitic steel. Corros. Sci. 2020, 163, 108251. [Google Scholar] [CrossRef]
  44. Ohtsuka, H.; Tsuzaki, K. Recent Studies on the Nature and State of Carbon Atoms in Iron. ISIJ Int. 2021, 61, 2677–2686. [Google Scholar] [CrossRef]
  45. Luo, W.D.; Hou, T.P.; Liang, X.; Cheng, S.; Zheng, Y.H.; Li, Y.; Ye, X.F.; Wu, K.M. Structure and stability investigation of oxygen interaction with Fe in bcc-Fe. Vacuum 2023, 212, 112005. [Google Scholar] [CrossRef]
  46. Sagar, S.; Sluiter, M.H.F.; Dey, P. First—Principles study of hydrogen—Carbide interaction in bcc Fe. Int. J. Hydrogen Energy 2024, 50, 211–223. [Google Scholar] [CrossRef]
  47. Yang, Y.; Ou, X.Q.; Zhang, H.; Song, M. Positioning of interstitial carbon atoms in the deformed Fe-C system. Mater. Today Commun. 2023, 34, 105377. [Google Scholar] [CrossRef]
  48. Ji, H.; Zhang, P.; Li, G.; Wei, M.; Wang, Y.; Zou, T. The synergistic interaction of interstitial carbon/oxygen, alloying Cr/Ni and vacancy in dilute austenitic iron alloys. Mater. Today Commun. 2024, 41, 110853. [Google Scholar] [CrossRef]
  49. Liu, C.; Lu, W.; Xia, W.; Du, C.; Rao, Z.; Best, J.P.; Brinckmann, S.; Lu, J.; Gault, B.; Dehm, G.; et al. Massive interstitial solid solution alloys achieve near-theoretical strength. Nat. Commun. 2022, 13, 1102. [Google Scholar] [CrossRef]
  50. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. B 1964, 136, B864. [Google Scholar] [CrossRef]
  51. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, 1133. [Google Scholar] [CrossRef]
  52. Kresse, G. Ab-intio molecular-dynamics for liquid-metals. J. Non-Cryst. Solids 1995, 193, 222–229. [Google Scholar] [CrossRef]
  53. Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  54. Kresse, G.; Furthmuller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  55. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  56. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  57. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  58. Wang, V.; Xu, N.; Liu, J.C.; Tang, G.; Geng, W.T. VASPKIT: A user-friendly interface facilitating high-throughput computing and analysis using VASP code. Comput. Phys. Commun. 2021, 267, 108033. [Google Scholar] [CrossRef]
  59. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  60. Henkelman, G.; Jonsson, H. Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points. J. Chem. Phys. 2000, 113, 9978–9985. [Google Scholar] [CrossRef]
  61. Henkelman, G.; Uberuaga, B.P.; Jonsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113, 9901–9904. [Google Scholar] [CrossRef]
  62. Lv, Z.; Zhang, Q.; Li, R.; Hong, Y.; Zhang, X. Combined Effects of Carbon and Nitrogen in Fe20Cr6Mn6CN Alloys from First-Principles Calculation. J. Mater. Eng. Perform. 2021, 30, 8322–8335. [Google Scholar] [CrossRef]
  63. Lv, Z.; Xiao, S.; Xiao, Z.; Qian, L.; Wang, W.; Zhou, Y.; Fu, W. Structural properties and bonding characteristic of interfaces between VN and VC from density functional calculations. J. Alloys Compd. 2017, 718, 139–149. [Google Scholar] [CrossRef]
  64. Fang, C.M.; van Huis, M.A.; Zandbergen, H.W. Stability and structures of the ε-phases of iron nitrides and iron carbides from first principles. Scr. Mater. 2011, 64, 296–299. [Google Scholar] [CrossRef]
  65. Raebiger, H.; Lany, S.; Zunger, A. Charge self-regulation upon changing the oxidation state of transition metals in insulators. Nature 2008, 453, 763–766. [Google Scholar] [CrossRef]
  66. Wan, L.; Wang, Q.; Ye, X.; Cao, X.; Jin, S.; Gao, T. Insight into structural stability and helium diffusion behavior of Fe-Cr alloys from first-principles. RSC Adv. 2020, 10, 3277–3292. [Google Scholar] [CrossRef]
  67. Eyméoud, P.; Maugis, P. Magnetic behavior of transition metal solutes in α-iron: A classification. J. Magn. Magn. Mater. 2020, 513, 167223. [Google Scholar] [CrossRef]
  68. Wang, F.K.; Wang, W.L.; Du, J.; Zhang, L.; Sohn, I.; Zeng, J. Oxidation behavior of Cr-Mn-Si alloyed steel by mold flux melt in high temperatures. J. Mater. Res. Technol. 2024, 29, 3898–3909. [Google Scholar] [CrossRef]
  69. Fernandez, N.; Ferro, Y.; Kato, D.J.A.M. Hydrogen diffusion and vacancies formation in tungsten: Density Functional Theory calculations and statistical models. Acta Mater. 2015, 94, 307–318. [Google Scholar] [CrossRef]
Figure 1. (a) Fe54 supercell model; (b) IAs occupying O-site; (c) IAs occupying T-site.
Figure 1. (a) Fe54 supercell model; (b) IAs occupying O-site; (c) IAs occupying T-site.
Coatings 15 00929 g001
Figure 2. Schematic illustration of O-site IAs (C, N, or O) and their adjacent iron atoms in the BCC Fe supercell. The top table lists the magnetic moments of Fe atoms (μB). The Fe1/2 atom in (ac) represents the first-nearest-neighbor (1NN) atom to IAs, and Fe3/4/5/6 represents the second-nearest-neighbor (2NN).
Figure 2. Schematic illustration of O-site IAs (C, N, or O) and their adjacent iron atoms in the BCC Fe supercell. The top table lists the magnetic moments of Fe atoms (μB). The Fe1/2 atom in (ac) represents the first-nearest-neighbor (1NN) atom to IAs, and Fe3/4/5/6 represents the second-nearest-neighbor (2NN).
Coatings 15 00929 g002
Figure 3. (a) PDOS of the C and its 1NN Fe, 2 NN Fe in Fe54C; (b) PDOS of the N and its 1NN Fe, 2 NN Fe in Fe54N; (c) PDOS of the O and its 1NN Fe, 2 NN Fe in Fe54O.
Figure 3. (a) PDOS of the C and its 1NN Fe, 2 NN Fe in Fe54C; (b) PDOS of the N and its 1NN Fe, 2 NN Fe in Fe54N; (c) PDOS of the O and its 1NN Fe, 2 NN Fe in Fe54O.
Coatings 15 00929 g003
Figure 4. The charge density difference (a,c,e) of the <001> plane and the charge density difference (b,d,f) of the <100> plane in Fe54C, Fe54N, and Fe54O, respectively. It plotted from −0.015 (blue) to 0.015 (red) e/Å3. Bader charge population denotes C/N/O IAs and their neighboring Fe-atom.
Figure 4. The charge density difference (a,c,e) of the <001> plane and the charge density difference (b,d,f) of the <100> plane in Fe54C, Fe54N, and Fe54O, respectively. It plotted from −0.015 (blue) to 0.015 (red) e/Å3. Bader charge population denotes C/N/O IAs and their neighboring Fe-atom.
Coatings 15 00929 g004
Figure 5. Schematic migration path of an interstitial atom C/N/O (a), where 0 represents the initial O-site, and 4 represents the final O-site. Diffusion barrier curve of C/N/O atom in Fe54 lattice; (b) Note that 0 represents the initial state, 4 represents the final state, 1 and 3 are the transition states. Please note that the curves on the graph are only for illustrative purposes.
Figure 5. Schematic migration path of an interstitial atom C/N/O (a), where 0 represents the initial O-site, and 4 represents the final O-site. Diffusion barrier curve of C/N/O atom in Fe54 lattice; (b) Note that 0 represents the initial state, 4 represents the final state, 1 and 3 are the transition states. Please note that the curves on the graph are only for illustrative purposes.
Coatings 15 00929 g005
Figure 6. Formation energy (ΔEf) and binding energy (Eb) of two homo IAs occupying different configurations of O-sites. (a) and (b) the different configurations of two homo IAs; (c) Formation energy (ΔEf) and binding energy (Eb) of Fe54C2; (d) Formation energy (ΔEf) and binding energy (Eb)of Fe54N2; (e) Formation energy (ΔEf) and binding energy (Eb)of Fe54O2.
Figure 6. Formation energy (ΔEf) and binding energy (Eb) of two homo IAs occupying different configurations of O-sites. (a) and (b) the different configurations of two homo IAs; (c) Formation energy (ΔEf) and binding energy (Eb) of Fe54C2; (d) Formation energy (ΔEf) and binding energy (Eb)of Fe54N2; (e) Formation energy (ΔEf) and binding energy (Eb)of Fe54O2.
Coatings 15 00929 g006
Figure 7. Configurations and energy calculation of two hetero C/N/O IAs. (a) Four relatively stable configurations; (b) ΔEf and Eb of Fe54CO, Fe54CN, and Fe54NO; (c) the most stable configuration based on energy calculations.
Figure 7. Configurations and energy calculation of two hetero C/N/O IAs. (a) Four relatively stable configurations; (b) ΔEf and Eb of Fe54CO, Fe54CN, and Fe54NO; (c) the most stable configuration based on energy calculations.
Coatings 15 00929 g007
Figure 8. (a) Schematic diagram of C/O and their neighbor iron atoms in Fe54CO; (b) schematic diagram of C/N and their neighbor iron atoms in Fe54CN; (c) schematic diagram of N/O and their neighbor iron atoms in Fe54NO.
Figure 8. (a) Schematic diagram of C/O and their neighbor iron atoms in Fe54CO; (b) schematic diagram of C/N and their neighbor iron atoms in Fe54CN; (c) schematic diagram of N/O and their neighbor iron atoms in Fe54NO.
Coatings 15 00929 g008
Figure 9. (a,c,e) show the charge density difference of Fe54CO, Fe54CN, and Fe54NO on the <011> plane, respectively. (b,d,f) shows the charge density difference Fe54CO, Fe54CN, and Fe54NO on the <100> plane, respectively, ranging from −0.015 (blue) to 0.015 (red) e/Å3.
Figure 9. (a,c,e) show the charge density difference of Fe54CO, Fe54CN, and Fe54NO on the <011> plane, respectively. (b,d,f) shows the charge density difference Fe54CO, Fe54CN, and Fe54NO on the <100> plane, respectively, ranging from −0.015 (blue) to 0.015 (red) e/Å3.
Coatings 15 00929 g009
Figure 10. PDOS of two homo IAs and the Fe atom between them in (a) Fe54C2, (b) Fe54N2, and (c) Fe54O2, and PDOS of two hetero IAs and the Fe atom between them in (d) Fe54CO, (e) Fe54CN, and (f) Fe54NO.
Figure 10. PDOS of two homo IAs and the Fe atom between them in (a) Fe54C2, (b) Fe54N2, and (c) Fe54O2, and PDOS of two hetero IAs and the Fe atom between them in (d) Fe54CO, (e) Fe54CN, and (f) Fe54NO.
Coatings 15 00929 g010
Table 1. Formation energy (ΔEf), lattice constant after optimization (a, b, c), supercell volume variation (∆V), average magnetic moment (M) of the BCC Fe (3 × 3 × 3) supercells.
Table 1. Formation energy (ΔEf), lattice constant after optimization (a, b, c), supercell volume variation (∆V), average magnetic moment (M) of the BCC Fe (3 × 3 × 3) supercells.
AlloysFormation Energy
/eV
Lattice Constant/ÅVolume Variation/Å3Average Magnetic Moment/μB
ΔEfabc∆VM
Fe54-8.4938.4938.49302.20
Fe54C-Oct0.7588.4938.4938.64911.2012.23
Fe54C-Tet1.6388.5918.4408.59110.2582.21
Fe54N-Oct−0.5188.5178.5178.63911.6982.27
Fe54N-Tet0.2258.5748.5068.57410.7862.22
Fe54O-Oct−1.0278.4998.4998.64414.0092.25
Fe54O-Tet−0.5118.5788.4738.57812.6352.21
Table 2. Bader charge and local magnetic moment variations of IAs and their neighbor iron atoms in Fe54CO, Fe54CN, and Fe54NO.
Table 2. Bader charge and local magnetic moment variations of IAs and their neighbor iron atoms in Fe54CO, Fe54CN, and Fe54NO.
Bader ChargeLocal Magnetic Moment/μB
Fe54COFe54CNFe54NOFe54COFe54CNFe54NO
C−1.060−1.058-−0.142−0.142-
N-−1.149−1.157-−0.108−0.103
O−1.083-−1.0970.072-0.073
Fe10.1230.0910.0922.0421.6391.704
Fe20.1700.1630.1742.4602.1812.219
Fe30.1610.1630.1722.4462.2432.256
Fe40.2850.3510.3782.3322.2302.272
Fe50.0570.0720.1181.6981.6612.059
Fe60.1680.1570.1632.2422.2252.439
Fe70.1610.1650.1462.1942.2102.494
Table 3. ΔEf and Eb of single IA (occupying O-site) and double IAs (occupying configuration 11 as shown in Figure 7).
Table 3. ΔEf and Eb of single IA (occupying O-site) and double IAs (occupying configuration 11 as shown in Figure 7).
Alloys∆Ef (eV)Eb (eV)
Fe54C0.758-
Fe54N−0.518-
Fe54O−1.027-
Fe54C21.2330.173
Fe54N2−0.640 0.092
Fe54O2−1.6020.036
Fe54CO0.0420.189
Fe54CN0.5990.141
Fe54NO−1.2150.158
Table 4. The binding energy (Eb), migration energy (Em), dissolution energy (Edis), and dissolution temperature (Tdis) of single O (occupying O-site), C–O, and N–O (occupying configuration 4 as shown in Figure 7). Note that Edis = Em + Eb is adopted for IA, and the heating rate during dissociation calculation is 1 K/s.
Table 4. The binding energy (Eb), migration energy (Em), dissolution energy (Edis), and dissolution temperature (Tdis) of single O (occupying O-site), C–O, and N–O (occupying configuration 4 as shown in Figure 7). Note that Edis = Em + Eb is adopted for IA, and the heating rate during dissociation calculation is 1 K/s.
Eb (eV)Em (eV)Edis (eV)Tdis (K)
single O-0.557-202
CO → C + O0.1890.5570.746266
NO → N + O0.1580.5570.715255
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

Wang, F.; Mi, T.; Chen, P.; Zhu, H.; Chen, Y.; Zhang, P.; Li, R.; Qiu, C. Synergistic Effect of Hetero Interstitial Atoms (C/N/O) on the Thermodynamic Stability in BCC Fe: A DFT Study. Coatings 2025, 15, 929. https://doi.org/10.3390/coatings15080929

AMA Style

Wang F, Mi T, Chen P, Zhu H, Chen Y, Zhang P, Li R, Qiu C. Synergistic Effect of Hetero Interstitial Atoms (C/N/O) on the Thermodynamic Stability in BCC Fe: A DFT Study. Coatings. 2025; 15(8):929. https://doi.org/10.3390/coatings15080929

Chicago/Turabian Style

Wang, Fang, Tengge Mi, Pinghu Chen, Hongmei Zhu, Yong Chen, Pengbo Zhang, Ruiqing Li, and Changjun Qiu. 2025. "Synergistic Effect of Hetero Interstitial Atoms (C/N/O) on the Thermodynamic Stability in BCC Fe: A DFT Study" Coatings 15, no. 8: 929. https://doi.org/10.3390/coatings15080929

APA Style

Wang, F., Mi, T., Chen, P., Zhu, H., Chen, Y., Zhang, P., Li, R., & Qiu, C. (2025). Synergistic Effect of Hetero Interstitial Atoms (C/N/O) on the Thermodynamic Stability in BCC Fe: A DFT Study. Coatings, 15(8), 929. https://doi.org/10.3390/coatings15080929

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