Next Article in Journal
Initiation Behavior of Microscopic Short Cracks in Weld Toe of Structural Steel Q345B
Previous Article in Journal
The Microstructure Formation in Slag Solidification at Continuous Casting Mold
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Cu on Nitriding of α-Fe

Institute of Materials Science, Technische Universität Bergakademie Freiberg, Gustav-Zeuner-Str. 5, D-09599 Freiberg, Germany
*
Author to whom correspondence should be addressed.
Metals 2022, 12(4), 619; https://doi.org/10.3390/met12040619
Submission received: 12 March 2022 / Revised: 29 March 2022 / Accepted: 31 March 2022 / Published: 3 April 2022
(This article belongs to the Section Metal Casting, Forming and Heat Treatment)

Abstract

:
Nitriding of Fe-1 wt.% Cu and Fe-5 wt.% Cu alloys at 813 K leads to the formation of predominantly the γ′-iron nitride phase (γ′-Fe4N) when using nitriding conditions, which lead to pronounced formation of ε-iron nitride phase (ε-Fe3N1+x) upon nitriding of pure α-Fe. Energy dispersive X-ray analysis reveals that the developing γ′ can attain a Cu content corresponding to that of the base material. In contrast, tiny amounts of ε-nitride that eventually develop contain considerably less Cu. The microstructure implies that the formation of the ε-nitride requires long-range substitutional interdiffusion to achieve the Cu partitioning. These observations were interpreted in terms of a significantly higher solubility of Cu in the γ′ phase than in the ε phase, which is explainable by the phases’ crystal structures. The observations were rationalized in terms of schematic Fe–Cu–N phase diagrams valid for 813 K.

1. Introduction

Nitriding (and nitrocarburizing) of steel and (e.g., cast) iron is a widely applied thermochemical heat treatment procedure, by which N (and C) from an external source, typically being a gas or a plasma, are introduced into the surface-near region of a workpiece by diffusion [1]. If applied to non-austenitic base materials, nitriding is usually performed below the eutectoid temperature of the Fe-base material. The treatment makes use of the moderate solubility of N and high diffusivity of N in α-Fe, leading to diffusion depths of up to several 100 μm. Generally, one distinguishes between a compound layer at the surface and a diffusion layer underneath. The compound layer is typically composed of the ε and/or the γ′ phase, being hcp- and fcc-based interstitial phases, both showing long-range ordering of N on the octahedral sites. See Figure 1 for the crystal structure of the γ′ phase. The term “diffusion layer” refers to the substrate underneath the compound layer being enriched in N. In the case of pure Fe, N is dissolved in solid solution and may precipitate during cooling. It is also the diffusion layer where characteristic alloying element-base nitride precipitates can already develop during nitriding, which can increase hardness in the diffusion zone and improve fatigue properties by inducing compressive stress.
The fundamental basis for the understanding of the phase transformations occurring upon nitriding iron-base alloys is the thermodynamics of the Fe–N system. Phase diagrams Fe–N are shown in Figure 2a,b in two forms. Figure 2a depicts what is typically shown as a Fe–N phase diagram. It reflects the solid-state phase equilibria between iron nitrides with N contents up to the ε phase, in the present version ignoring the ζ phase, Fe2N, and higher-N content iron nitrides [2,3] which were not included in the thermodynamic modelling, being the basis for that diagram [4].
Both ε and γ′ nitride are metastable with respect to N-containing α/γ and N2 gas at 1 atm, implying that Figure 2a shows actually a metastable phase diagram. In practice, these nitrides can be prepared by metastable equilibration of iron at some elevated temperature with NH3 + H2 containing atmospheres via the reaction
NH 3     [ N ] + 3 2 H 2 ,
where [N] represents N incorporated into a solid phase. The chemical potential of N in a solid equilibrated with an NH3 + H2 containing atmospheres can be shown to be related to the atmosphere’s nitriding potential rN [5]:
r N = p NH 3 p H 2 3 / 2 ,
with the partial pressures pi of the gas-phase species i. In more detail, in the case of a gas–solid equilibrium according to Equation (1), the activity aN of N at the surface of the solid increases linearly with rN and the chemical potential increases with lnrN. Note that the equilibrium N2 pressure corresponding to the NH3 + H2 mixtures can easily amount to several 1000 atm.
Figure 1. Unit cell of the crystal structure of γ′-Fe4N [6,7] based on an fcc arrangement of the Fe atoms and ordered occupation of ¼ of the octahedral sites by N. That ordering leads to two crystallographically distinct Fe(0) and Fe(II) atoms, with the Fe(0) being substituted by Cu. Note that this structure is frequently referred to as antiperowskite [8].
Figure 1. Unit cell of the crystal structure of γ′-Fe4N [6,7] based on an fcc arrangement of the Fe atoms and ordered occupation of ¼ of the octahedral sites by N. That ordering leads to two crystallographically distinct Fe(0) and Fe(II) atoms, with the Fe(0) being substituted by Cu. Note that this structure is frequently referred to as antiperowskite [8].
Metals 12 00619 g001
Accordingly, exposing the Fe to an atmosphere exhibiting a certain value of rN at a given temperature, a specific solid phase with specific composition should be in equilibrium with that atmosphere (if decomposition processes leading to N2 can be ignored; if that is not the case, a somewhat lower N content will be attained; “close to equilibrium”). The phase forming is classically indicated by the so-called Lehrer diagram (see Figure 2b), corresponding to a potential phase diagram, noting that the chemical potential of N is the variable conjugate to the amount of N in the system. As also seen in the present work, nitriding of massive iron typically leads to (close to) equilibrium at the surface, while the inwards diffusion occurs at a finite rate. Hence, the N content will decrease with increasing distance to the surface. If the value of rN suffices to produce an ε-iron nitride layer at the surface of the specimen, the decrease of the N content will lead under local equilibrium to an intermediate γ′ layer between the ε phase at the surface and the α substrate (ε/γ′ double layer).
The current work deals with the thermodynamics of nitriding in the presence of minor amounts of an alloying element M (considering M to be a metal or metalloid element, but not carbon). Estimating the results of nitriding of binary Fe-M alloys typically involves comparison of the stability of the nitrides in the M–N system with those in the Fe–N system. Higher stability of the nitrides in the M–N system typically allows formation of such M nitrides at rN values much lower than required for the formation γ′ or ε iron nitrides. This explains the formation of alloying element nitrides like VN, CrN or AlN in the diffusion zone [9], occurring while depleting the matrix by the alloying element (referred to as internal nitriding). N uptake beyond that expected from the precipitation of such stable, e.g., stoichiometric MN nitrides and N saturated Fe (as implied by bulk thermodynamics) was reported for nitrided Fe-M alloys. This is often ascribed to so-called excess nitrogen caused by the interface precipitate matrix and due to the strain field in the matrix caused by the precipitates [10,11].
Considerations to estimate the amount of excess nitrogen often appear to neglect a possible uptake of Fe by the M–N nitrides. Predicting such an uptake is obstructed by limited knowledge of the equilibrium phase diagrams Fe–M–N at the relevant temperatures and by the usually nanoscopic nature of the nitrides developing (obstructing, in any case, the validity of consideration purely based on the equilibrium phase diagrams/bulk thermodynamics). There is, however, clear evidence for such a Fe uptake by nitrides, e.g., by CrN [12,13] and by amorphous Si3N4 (actually (Si3N4)1−x(Fe3N2)x [14]). N uptake by such nitrides is sometimes predicted in the course of thermodynamic calculations by the CALPHAD method, e.g., Fe uptake by the VN phase [15], which, however, does not necessarily imply that the thermodynamic description of the phase is really based on clear experimental evidence for such particular behaviour.
Figure 2. Binary phase diagrams relevant for the present work with arrows marking the treatment temperature of 813 K. (a,b) Metastable phase diagrams Fe–N calculated from a thermodynamic database [4] excluding N2 gas (redrawn): (a) Temperature vs. molar fraction of N xN and (b) complementary potential phase diagram in the form of the Lehrer diagram T vs. nitriding potential rN. The colored lines in (b) mark the equilibrium existence of the corresponding phases at 813 K have also been used in following schematics and phase diagrams, with the points and numbers indicating the nitriding potentials applied experimentally in atm−1/2. (c) Phase diagram Fe–Cu [16].
Figure 2. Binary phase diagrams relevant for the present work with arrows marking the treatment temperature of 813 K. (a,b) Metastable phase diagrams Fe–N calculated from a thermodynamic database [4] excluding N2 gas (redrawn): (a) Temperature vs. molar fraction of N xN and (b) complementary potential phase diagram in the form of the Lehrer diagram T vs. nitriding potential rN. The colored lines in (b) mark the equilibrium existence of the corresponding phases at 813 K have also been used in following schematics and phase diagrams, with the points and numbers indicating the nitriding potentials applied experimentally in atm−1/2. (c) Phase diagram Fe–Cu [16].
Metals 12 00619 g002
Much less is known about the effect of elements M forming less-stable nitrides than Fe. As it concerns the series of 3d metals, stability of transition metal nitrides generally decreases from Ti towards Cu [17]. For Ti…Ni this has also been shown quantitatively on the basis of the Gibbs energies of dissolution of N into terminal transition metal-based solid solution [18]. In the whole row Ti…Cu, M = Cu should develop the least stable nitrides. The best-known binary Cu nitride is Cu3N, which is, however, not accessible from elemental Cu upon interaction with NH3 + H2 containing atmospheres, but requires non-metallic precursors to be treated with NH3 [19,20,21]. Cu3N decomposes irreversibly above (620–720) K under kinetic control to Cu + N2 [19,21]. This nitride is definitely not expected to develop upon nitriding Fe–Cu alloy using NH3 + H2 atmospheres typically used to nitride Fe-base alloys. Instead, only iron nitrides are expected to form, which may or may not be affected by the presence of Cu.
Below the eutectoid temperature, Cu has a very low solubility in iron, as visible from the binary Fe–Cu phase diagram [16] shown in Figure 2c. In steel, minor amounts of Cu may be used to improve corrosion properties and to improve strength [22,23]. However, it is usually regarded as an unwanted tramp element usually incorporated during recycling [24,25], and too high contents may lead to hot cracking [26,27,28]. As it concerns cast iron, Cu is occasionally deliberately added (alongside with other elements) to promote the formation of pearlite [29].
It was a study on nitriding of Cu-containing, white-solidified cast iron [30] which suggested that Cu promotes formation of the γ′ nitride against that of ε, which was attributed to different solubilities of these elements in the two iron-nitride phases. To confirm this idea, in the present work the effect of Cu in iron on the competition between the γ′ with ε iron nitrides is studied using binary Fe–Cu alloys to exclude complications due to Si and C, as presented in Ref. [30].

2. Materials and Methods

2.1. Alloy Preparation

Fe was supplied as cold-rolled plates (“α-Fe”; thickness 1 mm; 99.99%, Goodfellow) which were encapsulated in fused silica tubes under Ar. For recrystallization, the tubes were heated for 2 h at 973 K and slowly cooled.
Fe–Cu Alloys with nominal compositions of 1 wt.% Cu (“Fe-1Cu”) and 5 wt.% Cu (“Fe-5Cu”) were prepared by arc melting in an Arc Melter AM 200 (Edmund Bühler GmbH, Bodelshausen, Germany) from Fe granules (99.98%, Alfa Aesar) and Cu metal (99.9995%, VEB Spurenmetalle Freiberg, Germany) with batch sizes of about 10 g. Material loss upon preparation was very small, such that the nominal composition corresponding to molar fractions of xCu = 0.009 for Fe-1Cu and xCu = 0.044 were adopted to be valid. The alloy batches were hammered to a thickness of about 4 mm and cut into smaller, plate-like specimens. These were encapsulated as above and heat treated for 20 min at 1383 K. These heat treatments were terminated by quenching in water with a temperature of about 300 K, including crushing the fused silica tubes. This should ensure rapid occurrence of the γ → α transition in order to avoid long-range Cu partitioning.
Prior to nitriding, the α-Fe, Fe-1Cu and Fe-5Cu plate specimens were ground and polished (final stage 1 µm diamond paste, performed directly prior to nitriding), washed with water and ethanol and then dried. Nitriding in a NH3 + H2 gas mixture at 813 K was performed in a vertical fused-silica tube furnace (tube diameter of 28 mm) described in some more detail in Ref. [5]. The overall gas-flow rate was adjusted to 500 mL/min pertaining to room temperature. The gas was composed from NH3 and H2 to result in rN = 1, 2.5 or 4 atm−1/2 according to Equation (2). The nitriding treatment was terminated by quenching in water having a temperature of about 300 K, which was previously flushed with N2.

2.2. Microstructure Analysis

X-ray diffraction (XRD) measurements were performed on the surface of the samples in a diffraction angle 2θ range of 30°–125° with a Bruker D8 ADVANCE (Bruker AXS, Karlsruhe Germany) diffractometer being equipped with a Co tube, a quartz Johannsson monochromator in the primary beam selecting the CoKα1 component of the fluorescence radiation (wave length 1.78897 Å). The diffracted intensity was recorded using a Lynxeye position sensitive detector (Bruker AXS, Karlsruhe, Germany). The diffraction patterns were evaluated using the Bruker-AXS TOPAS software (Version 5, Bruker AXS, Karlsruhe, Germany) [31] based on Rietveld and Pawley fits, considering the crystal structures of α-Fe, γ′-Fe4N and ε-Fe3N1+x (e.g., [2,7,32]).
Scanning electron microscopy (SEM) was performed on polished metallographic cross-sections. To prepare these, the nitrided specimens were coated electrochemically by a nickel layer to reduce edge effects upon polishing the cross sections. The coated specimens were cut, embedded, ground and polished (final stage colloidal silica). The SEM investigations were conducted using a JEOL JSM-7800 F (JEOL, Tokyo, Japan) equipped with EDAX Octane Elite EDS system (EDS: Energy dispersive electron spectroscopy) and EDAX Hikari Super EBSD system (EBSD: Electron backscatter diffraction), both Ametek, Weiterstadt, Germany. An atomic number sensitive backscatter electron (BSE) was used for imaging. EBSD was measured with a step-width of 250 nm and 50 nm for overview and detailed measurements, respectively. EBSD patterns were indexed based on the crystal structure data from [2,18,32].

3. Results

The compositions of the Fe-1Cu and Fe-5Cu alloys are located within the γ solid solution region at 1383 K [16], see Figure 2c. SEM-based EDS investigations of metallographic cross-sections of the as-quenched alloys did not indicate a macroscopically inhomogeneous distribution of Cu. While the microstructure of the Fe-1Cu was ferritic, that of Fe-5Cu showed a more complicated microstructure, suggesting a massive or martensitic transformation mechanism, in agreement with Ref. [33].
In reasonable agreement with [33,34], the alloys water-quenched from the austenite region showed (room-temperature) lattice parameters of 2.869 Å (Fe-1Cu) and 2.872 Å (Fe-5Cu), which are significantly increased as compared to the lattice parameter of pure α-Fe (2.866 Å). Heat treatments at 813 K, the temperature also selected for nitriding, revealed a decrease in the lattice parameters with increasing annealing time, asymptotically approaching the value of α-Fe. Comparison with, e.g., Ref. [33] implies the formation of finely distributed nanoscopic Cu-rich precipitates, which is also confirmed by more in-depth analysis of the nitrided specimens, see what follows.
Exemplary cross-sectional microstructures of the compound layers, formed upon nitriding of α-Fe, Fe-1Cu and Fe-5Cu alloys at 813 K and varying nitriding potentials, are depicted in Figure 3. EBSD analysis, thereby, allowed straightforward identification of the phases having developed as the compound layer. X-ray diffraction data, depicted in the Supplementary Material (Figures S1–S3), always show, apart from reflections of the α-substrate material, reflections due to nitride phases identified by SEM-based investigations.
Nitriding of pure α-Fe at rN = 1 atm−1/2 for 4 h leads to the formation of a pure γ′ layer (Figure 3a). Upon increasing the nitriding potential to 2.5 atm−1/2 and keeping the treatment time at 4 h, an ε/γ′ double layer develops on pure α-Fe (Figure 3b). The microstructure is similar for rN = 4 atm−1/2, only leading to a thicker double layer (not shown).
Nitriding of the Fe–Cu alloys, however, appears to significantly suppress the development of the ε phase. As illustrated in Figure 3c,d, nitriding of Fe-1Cu and Fe-5Cu at rN = 2.5 atm−1/2 for 4 h leads to the formation of only little (Figure 3c) or no (Figure 3d) ε-iron nitride. Upon nitriding at rN = 4 atm−1/2 ε forms in both Fe-1Cu and Fe-5Cu, alongside some porosity (see Figure 3e). Nevertheless, the difference to the corresponding treatment of pure α-Fe remains considerable. Note that the small amounts of ε phase preferentially appear to occur close to the surface, where the microstructure is highly irregular, including the presence of pores which must have formed during nitriding. Apart from the ε-suppressing effect, the Cu in the substrate also promotes the formation of γ′-iron nitride in some depths of the substrate (see Figure 3e), which is not typical for pure α-Fe quenched after nitriding.
SEM imaging at a higher resolution reveals the presence of regularly distributed features appearing brighter than the α-Fe matrix in BSE contrast images. These bright features occur just beneath the compound layer (Figure 4a) but also in a larger depth of about 800 μm (Figure 4b), i.e., much below the diffusion depth of the N in the diffusion zone. In line with a large body of literature, these particles should correspond to Cu-rich precipitates developing in the α-Fe due to the thermal load independent of the nitriding [34,35,36]. Strikingly, these particles appear to be absent (apart from possible tiny remainders) in the γ′-iron nitride (compare Figure 4a).
The Cu content in the iron nitrides was analysed using EDS, see Figure 5. The Cu content in the γ′ phase appears to correspond to the uCu = xCu/(xFe + xCu) content in the substrate of uCu ≈ 0.044. In contrast, the Cu content is much lower in the ε particles. We refrain from stating a value due to the tiny character of the particles and mutual detection of Ni fluorescence radiation from the adjacent protective Ni layer.

4. Discussion

The outcome of the interaction of pure α-Fe with a NH3 + H2 atmosphere can be predicted by the potential phase diagram/Lehrer diagram depicted in Figure 2b. As explained in the introduction, the surface is (close to) equilibrated with the atmosphere, and due to the relatively slow inwards diffusion of the N, an N-content depth profile into the α-Fe develops. Accordingly (see Figure 3a,b), formation of γ′ phase for rN = 1 atm−1/2 and ε phase for rN = 2.5 atm−1/2 and 4 atm−1/2 at the surface of the nitrided pure α-Fe is in good agreement with the prediction from Figure 2b. All this is in accordance with a large body of investigations on the nitriding behaviour of α-Fe at similar temperatures [1].
As expected from the low solubility of Cu in α-Fe at the nitriding temperature of 813 K [33,34] (see also Figure 2c), the Fe-1Cu and Fe-5Cu alloys develop towards two-phase assemblage under the thermal load, consisting in first approximation of pure α-Fe and pure Cu. On the basis of this, one may suppose that these two metals interact independently with the NH3 + H2 gas mixture. A “Lehrer diagram for Cu” is unknown, and the low thermodynamic stability of the Cu3N nitride implies that a high, probably unattainable minimum nitriding potential rN would be required to produce Cu3N (e.g., by nitriding in close-to-pure NH3). In fact, experimental evidence suggests rapid decomposition of once formed Cu3N into Cu + N2 at the presently applied temperature [19,20]. The minimum nitriding potential rN to produce Cu3N from pure Cu should be definitely higher than the minimum rN required to produce the ε-iron nitride phase from α-Fe. Accordingly, nitriding of a two-phase mixture α-Fe and Cu should ideally lead to formation of those iron nitrides which are also observed upon nitriding pure iron. These iron nitrides should then coexist with pure Cu. That situation has been illustrated at metal fraction values of uCu = xCu/(xFe + xCu) = 0 and 1 in Figure 6a. This diagram will, however, be “filled” at 0 < uCu < 1 after the following considerations.
The experimental observations made in Section 3 deviate from what is proposed above as ideal nitriding behavior of a two-phase assemblage α-Fe + Cu. Instead, the Cu seems to strongly affect the outcome of the nitriding by promoting the formation of the γ′ phase at the cost of the ε phase. That can be rationalized by assuming solubility of Cu in γ′ but not in the ε phase. This is supported by the following observations:
  • The developing γ′ appears to dissolve the Cu precipitates present in the α-Fe matrix (see Figure 4a), corresponding to a reaction α + Cu → γ′, with the latter γ′ having the uCu fraction values pertaining to the Fe-1Cu and Fe-5Cu substrates (uCu = 0.009 and 0.044, see also Figure 5).
  • The eventually developing ε phase contains significantly less Cu than the γ′ phase, see Figure 5.
Incorporation of Cu into γ′-Fe4N requires Fe/Cu interdiffusion on a length scale of a few tens of nanometers upon dissolving the particles into the γ′ phase. That dissolution is compatible with experimental and theoretical evidence in the literature. γ′-Fe4(1−uCu)Cu4uCuN has been prepared by ball-milling and subsequent annealing [37]. In that work, preferential substitution of the Fe atoms on the Fe(0) site by Cu atoms has been concluded from Mössbauer spectroscopy, as illustrated in Figure 1a. Thereby, the Cu having a lower affinity towards N than Fe (see the low stability of Cu3N) avoids close contact with the N atoms in the crystal structure (see Figure 1a). In that work, Cu uptake up to Fe3.4Cu0.6N corresponding to uCu = 0.15 (or substitution of 60 % of the Fe(0) sites) was found, with higher Cu contents leading to the formation of residual elemental Cu in the powder mixture. Further works on powder and thin film material [38,39] confirm the picture obtained in Ref. [37]. Some works have claimed the preparation of stoichiometric γ′-Fe3CuN (uCu = 0.25) [40], where proof may be missing that indeed a higher Cu content than in [37,38,39] had been attained in the γ′ phase. High-throughput computational screening of magnetic antiperovskite compounds with various elements [41] revealed that the energy of γ′-Fe3CuN is only 36 meV/atom above the convex hull of the Fe–Cu–N system (static structures at 0 K) devised in that work, such that γ′-Fe3CuN might be an accessible material. Note that γ′-Fe4N itself is located (10–20) meV/atom of higher energy than α-Fe + Fe3N, whereby γ′-Fe4N gets vibrationally stabilised at T > 0 K [42,43]. Nevertheless, in the considerations to follow, it will be assumed that the above-mentioned Fe3.4Cu0.6N corresponding to uCu = 0.15 constitute the maximum attainable Cu content on the γ′ phase.
Incorporation of Cu into the ε-iron nitride phase has not yet been reported or theoretically considered. The ideal crystal structure of ε-Fe3N contains only one sort of Fe atom, and this sort has two N atoms in its first coordination sphere [44], similar to the Fe(II) atoms in γ′-Fe4N. As the latter sites are apparently not occupied by Cu, it is also expected that similar sites in ε-iron nitride are inaccessible for Cu atoms.
The phase diagrams Fe–Cu–N at 813 K shown in Figure 6 were constructed based on
(a)
The Fe–N binary phase diagrams shown in Figure 2, in particular considering the N solubility in the ε phase, ranging approximately from 25 at.% to about 33 at.% (see also [2]) with a negligible solubility for Cu.
(b)
The low thermodynamic stability of Cu3N requiring yet unknown high rN values at 813 K, which are likely even higher than values derived for the Ni/Ni3N equilibrium [18], as discussed in Section 1. This implies the existence of two-phase regions of the γ′ and ε nitrides with Cu. For simplicity it is, however, assumed that at high rN values an equilibrium ε + Cu3N exists.
(c)
Continuous substitution of Fe (Fe(0); see above) by Cu up to a composition corresponding to Fe3.4Cu0.6N [37], while minor deviations from 20 at.% N content are ignored.
Considering the ternary extension of the γ′ homogeneity region implied by point (b), the two-phase regions α + γ′ and ε + γ′ will also extend into the ternary compositional region. The arrangements of the α + Cu and α + γ′ two-phase regions require that the three-phase triangle α + γ′ + Cu in Figure 6b implied by point (b) has a corner at the low-N edge of the γ′ homogeneity region. That corner has been chosen to be at the Fe3.4Cu0.6N endpoint of the γ′ region. A further γ′ + ε + Cu triangle has a corner at the ε phase and can connect all three phases only if the triangle touches the γ′ region at the Fe3.4Cu0.6N endpoint. Hence, the γ′ + Cu two-phase region degenerates to a line. Adjacent to the γ′ + ε + Cu triangle, there is a two-phase region ε + Cu, which, in view of point (b), is followed by an experimentally unsupported ε + Cu + Cu3N three-phase region, which is shown in Figure 6b to end at a ε + Cu3N two-phase region. Higher N-contents than corresponding to the low-N end of this region are not considered here.
The diagram constructed in Figure 6b is a basis also for the diagram in Figure 6a. It was assumed that, at a given molar fraction of N, the rN required to produce a given phase increases with increasing uCu, in agreement with the general trend of a lower stability of Cu nitrides as compared to Fe nitrides. Thereby, an extended γ′ field is obtained. Similarly, extended fields are shown for the other phases, exaggerating the homogeneity ranges with some artistic freedom in order to depict the three-phase reactions occurring with increasing rN, which are invariant at the adopted constant pressure of 1 atm. In view of the γ′ + ε microstructures with Cu being depleted in the ε phase encountered at the surface of the Fe-1Cu alloy at rN = 2.5 atm−1/2 and Fe-1Cu and Fe-5Cu alloys at rN = 4 atm−1/2, it is concluded that the value of rN corresponding to the invariant reaction γ′ ⇄ ε + Cu is not yet attained and is higher than 4 atm−1/2. It appears, however, likely that the results are strongly affected by sluggish partitioning of the substitutional elements, i.e., enrichment of the Cu in the γ′. For instance, the non-observation of ε nitride in the Fe-5Cu alloy after treatment at rN = 2.5 atm−1/2 might be the consequence of a state of paraequilibrium. Note that, upon the formation of ε, Cu/Fe interdiffusion has to occur over much larger distances than upon dissolution of the Cu precipitates upon γ′ formation (compare the size of the ε grains in Figure 5 as compared to the distance between the Cu precipitates in Figure 4a). The microstructures developing, in any case, are strongly affected by pore formation, which is usually attributed to N2 bubbles produced inside the solid [45]. Possibly, surface or grain-boundary diffusion contributes to the Fe/Cu partitioning necessary for ε formation, thus also contributing to the irregular microstructure shown in Figure 5.
The experiments of the current study have been conducted to reveal the basic influence of Cu in the Fe substrate on the outcome of nitriding. A more in-depth study is required in order to arrive at quantitative conclusions on the phase equilibria, e.g., to allow for thermodynamic modelling of the Fe–Cu–N system. This can only be achieved in future research.

5. Conclusions

Nitriding of Fe-1Cu and Fe-5Cu at 813 K at various nitriding potentials revealed marked differences as compared to nitriding of pure α-Fe at similar conditions. The presence of copper appears to stabilise the γ′-Fe4N as compared to the ε-Fe3N1+x iron nitride. This experimental observation can be rationalized in terms of much higher solubility of Cu in the γ′ than in the ε phase. The high solubility of Cu in the γ′ can be explained by the phase’s crystal structure providing a particular atomic Fe-site which has no close N atoms in its neighbourhood. Such a site can preferentially be occupied by Cu atoms as already shown in previous works, whereas a corresponding site is missing in the crystal structure of the ε-iron nitride. On the basis of the data, a schematic phase diagram of the Fe–Cu–N system has been sketched for the temperature of 813 K. The considerations in the present work can serve as a model case for investigation of other elements M on the corresponding Fe–M–N systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/met12040619/s1, Figure S1: X-ray diffraction data recorded with Co-Kα1 radiation: nitrided pure α-Fe. The numbers are Laue indices with the index indicating the phase. Figure S2: X-ray diffraction data recorded with Co-Kα1 radiation: nitrided Fe-1 wt.% Cu. The numbers are Laue indices with the index indicating the phase. Figure S3: X-ray diffraction data recorded with Co-Kα1 radiation: nitrided Fe-5 wt.% Cu. The numbers are Laue indices with the index indicating the phase.

Author Contributions

Conceptualization, A.L. and S.K.; methodology, M.F., S.K. and S.M.; formal analysis, A.L., M.F., S.K. and S.M.; investigation, A.L., M.F., S.K. and S.M.; resources, A.L.; data curation, S.K. and S.M.; writing—original draft preparation, A.L.; writing—review and editing, A.L., M.F., S.K. and S.M.; visualization, A.L. and S.M.; supervision, A.L. and S.K.; project administration, S.K.; funding acquisition, A.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the German Research Foundation for their financial support of the project LE 1403/1-2.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Materials or otherwise available upon reasonable request from the corresponding author A.L.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mittemeijer, E.J.; Somers, M.A.J. (Eds.) Thermochemical Surface Engineering of Steels; Woodhead Publishing: Oxford, UK, 2015; ISBN 978-0-85709-592-3. [Google Scholar]
  2. Wriedt, H.A.; Gokcen, N.A.; Nafziger, R.H. The Fe-N (Iron-Nitrogen) system. Bull. Alloy Phase Diagr. 1987, 8, 355–377. [Google Scholar] [CrossRef]
  3. Alkhaldi, H.; Kroll, P. Computing the iron–nitrogen phase diagram at high pressure and high temperature. J. Alloys Compd. 2022, 900, 163533. [Google Scholar] [CrossRef]
  4. Göhring, H.; Fabrichnaya, O.; Leineweber, A.; Mittemeijer, E.J. Thermodynamics of the Fe-N and Fe-N-C Systems: The Fe-N and Fe-N-C Phase Diagrams Revisited. Metall. Mater. Trans. A 2016, 47, 6173–6186. [Google Scholar] [CrossRef] [Green Version]
  5. Leineweber, A.; Gressmann, T.; Mittemeijer, E.J. Simultaneous control of the nitrogen and carbon activities during nitrocarburising of iron. Surf. Coat. Techn. 2012, 206, 2780–2791. [Google Scholar] [CrossRef]
  6. Jack, K.H.; Goodeve, C.F. Binary and ternary interstitial alloys I. The iron-nitrogen system: The structures of Fe4N and Fe2N. Proc. R. Soc. A 1948, 195, 34–40. [Google Scholar] [CrossRef] [Green Version]
  7. Jacobs, H.; Rechenbach, D.; Zachwieja, U. Structure determination of γ′-Fe4N and ϵ-Fe3N. J. Alloys Compd. 1995, 227, 10–17. [Google Scholar] [CrossRef]
  8. Niewa, R. Metal-Rich Ternary Perovskite Nitrides. Eur. J. Inorg. Chem. 2019, 2019, 3647–3660. [Google Scholar] [CrossRef] [Green Version]
  9. Steiner, T.; Mittemeijer, E.J. Alloying Element Nitride Development in Ferritic Fe-Based Materials Upon Nitriding: A Review. J. Mater. Eng. Perf. 2016, 25, 2091–2102. [Google Scholar] [CrossRef] [Green Version]
  10. Hosmani, S.S.; Schacherl, R.E.; Mittemeijer, E.J. Nitrogen uptake by an Fe–V alloy: Quantitative analysis of excess nitrogen. Acta Mater. 2006, 54, 2783–2792. [Google Scholar] [CrossRef]
  11. Somers, M.A.J.; Lankreijer, R.M.; Mittemeijer, E.J. Excess nitrogen in the ferrite matrix of nitrided binary iron-based alloys. Philos. Mag. A 1989, 59, 353–378. [Google Scholar] [CrossRef]
  12. Jessner, P.; Danoix, R.; Hannoyer, B.; Danoix, F. Investigations of the nitrided subsurface layers of an Fe–Cr-model alloy. Ultramicroscopy 2009, 109, 530–534. [Google Scholar] [CrossRef]
  13. Tomio, Y.; Kitsuya, S.; Oh-ishi, K.; Hono, K.; Miyamoto, G.; Furuhara, T. Plasma Nitriding Behavior of Fe-C-M (M = Al, Cr, Mn, Si) Ternary Martensitic Steels. Metall. Mater. Trans. A 2014, 45, 239–249. [Google Scholar] [CrossRef]
  14. Kante, S.; Kürnsteiner, P.; Motylenko, M.; Gault, B.; Leineweber, A. Eutectoid growth of nanoscale amorphous Fe-Si nitride upon nitriding. Acta Mater. 2021, 209, 116774. [Google Scholar] [CrossRef]
  15. Kouba, R.; Keddam, M.; Djeghlal, M.E. A diffusion–precipitation model for gaseous nitriding of Fe-2 wt.% V alloy. J. Alloys Compd. 2012, 536, 124–131. [Google Scholar] [CrossRef]
  16. Chen, Q.; Jin, Z. The Fe-Cu system: A thermodynamic evaluation. Metal. Mater. Trans. A 1995, 26, 417–426. [Google Scholar] [CrossRef]
  17. Juza, R.; Hahn, H. Kupfernitrid Metallamide und Metallnitride. VII. Z. Anorg. Allg. Chem. 1939, 241, 172–178. [Google Scholar] [CrossRef]
  18. Zachwieja, U.; Jacobs, H. Ammonothermalsynthese von Kupfernitrid, Cu3N. J. Less-Comm. Met. 1990, 161, 175–184. [Google Scholar] [CrossRef]
  19. Paniconi, G.; Stoeva, Z.; Doberstein, H.; Smith, R.I.; Gallagher, B.L.; Gregory, D.H. Structural chemistry of Cu3N powders obtained by ammonolysis reactions. Solid State Sci. 2007, 9, 907–913. [Google Scholar] [CrossRef]
  20. Juza, R. Nitrides of Metals of the First Transition Series. In Advances in Inorganic Chemistry and Radiochemistry; Emeléus, H.J., Sharpe, A.G., Eds.; Academic Press Inc.: New York, NY, USA, 1966; pp. 81–131. ISBN 0065-2792. [Google Scholar]
  21. Fonović, M.; Leineweber, A.; Mittemeijer, E.J. Experimental Investigation and Thermodynamic Modeling of the Ni-Rich Part of the Ni-N Phase Diagram. Metall. Mater. Trans. A 2014, 45, 4863–4874. [Google Scholar] [CrossRef]
  22. Görzen, D.; Schwich, H.; Blinn, B.; Bleck, W.; Beck, T. Influence of different precipitation states of Cu on the quasi-static and cyclic deformation behavior of Cu alloyed steels with different carbon contents. Int. J. Fatigue 2020, 136, 105587. [Google Scholar] [CrossRef]
  23. Hong, J.H.; Lee, S.H.; Kim, J.G.; Yoon, J.B. Corrosion behaviour of copper containing low alloy steels in sulphuric acid. Corros. Sci. 2012, 54, 174–182. [Google Scholar] [CrossRef]
  24. Pauliuk, S.; Milford, R.L.; Müller, D.B.; Allwood, J.M. The Steel Scrap Age. Environ. Sci. Technol. 2013, 47, 3448–3454. [Google Scholar] [CrossRef] [Green Version]
  25. Daehn, K.E.; Cabrera Serrenho, A.; Allwood, J.M. How Will Copper Contamination Constrain Future Global Steel Recycling? Environ. Sci. Technol. 2017, 51, 6599–6606. [Google Scholar] [CrossRef]
  26. Rod, O.; Becker, C.; Nylén, M. Opportunities and dangers of using residual elements in steels: A literature survey. Jernkontoret Rep. 2006, D819, 1–58. [Google Scholar]
  27. Daehn, K.E.; Serrenho, A.C.; Allwood, J. Finding the Most Efficient Way to Remove Residual Copper from Steel Scrap. Metall. Mater. Trans B 2019, 50, 1225–1240. [Google Scholar] [CrossRef]
  28. Savov, L.; Volkova, E.; Janke, D. Copper and Tin in Steel Scrap Recycling. RMZ Mater. Geoenviron. 2003, 50, 627–640. [Google Scholar]
  29. Lacaze, J. Pearlite growth in cast irons: A review of literature data. Int. J. Cast Metals Res. 1999, 11, 431–436. [Google Scholar] [CrossRef]
  30. Kante, S.; Leineweber, A. Interaction of N with White-solidified Cast Iron Model Alloys: The Effect of Mn and Cu on the Formation of Fe and Si Nitrides. J. Cast. Mater. Eng. 2021, 5, 66–70. [Google Scholar] [CrossRef]
  31. Dinnebier, R.E.; Leineweber, A.; Evans, J.S.O. Rietveld Refinement: Practical Powder Diffraction Pattern Analysis Using TOPAS; De Gruyter: Berlin, Germany; Boston, MA, USA, 2019; ISBN 9783110456219. [Google Scholar]
  32. Liapina, T.; Leineweber, A.; Mittemeijer, E.J.; Kockelmann, W. The lattice parameters of ε-iron nitrides: Lattice strains due to a varying degree of nitrogen ordering. Acta Mater. 2004, 52, 173–180. [Google Scholar] [CrossRef]
  33. Kimura, Y.; Takaki, S. Phase Transformation Mechanism of Fe-Cu Alloys. ISIJ Int. 1997, 37, 290–295. [Google Scholar] [CrossRef]
  34. Hornbogen, E. The role of strain energy during precipitation of copper and gold from alpha iron. Acta Metall. 1962, 10, 525–533. [Google Scholar] [CrossRef]
  35. Othen, P.J.; Jenkins, M.L.; Smith, G.D.W. High-resolution electron microscopy studies of the structure of Cu precipitates in α-Fe. Philos. Mag. A 1994, 70, 1–24. [Google Scholar] [CrossRef]
  36. Monzen, R.; Jenkins, M.L.; Sutton, A.P. The bcc-to-9R martensitic transformation of Cu precipitates and the relaxation process of elastic strains in an Fe-Cu alloy. Philos. Mag. A 2000, 80, 711–723. [Google Scholar] [CrossRef]
  37. De Figueiredo, R.S.; Foct, J.; dos Santos, A.V.; Kuhnen, C.A. Crystallographic and electronic structure of CuXFe4−XN. J. Alloys Compd. 2001, 315, 42–50. [Google Scholar] [CrossRef]
  38. El Khiraoui, S.; Sajieddine, M.; Vergnat, M.; Bauer, P.; Mabrouki, M. Mössbauer study of (Fe1−xCux)4N (0.05 ≤ x ≤ 0.15) films. Phys. B Condens. Matter 2007, 388, 180–182. [Google Scholar] [CrossRef]
  39. Wang, W.; Kan, X.; Liu, X.; Feng, S.; Liu, C.; Mehmood Ur Rehman, K.; Shezad, M.; Wu, Q.; Wang, Y. Effect of Cu on microstructure, magnetic properties of antiperovskite nitrides CuxNFe4−x. J. Mater. Sci. Mater. Electr. 2019, 30, 10383–10390. [Google Scholar] [CrossRef]
  40. Shi, X.; Li, X.; Lai, Z.; Liu, X.; Mi, W. Structure, magnetic and electronic transport properties in antiperovskite cubic γ′-CuFe3N polycrystalline films. Intermetallics 2020, 121, 106779. [Google Scholar] [CrossRef]
  41. Singh, H.K.; Zhang, Z.; Opahle, I.; Ohmer, D.; Yao, Y.; Zhang, H. High-Throughput Screening of Magnetic Antiperovskites. Chem. Mater. 2018, 30, 6983–6991. [Google Scholar] [CrossRef]
  42. Leineweber, A.; Hickel, T.; Azimi-Manavi, B.; Maisel, S.B. Crystal structures of Fe4C vs. Fe4N analysed by DFT calculations: Fcc-based interstitial superstructures explored. Acta Mater. 2017, 140, 433–442. [Google Scholar] [CrossRef]
  43. Wetzel, M.H.; Rabending, T.T.; Friák, M.; Všianská, M.; Šob, M.; Leineweber, A. Phase Stability of Iron Nitride Fe4N at High Pressure—Pressure-Dependent Evolution of Phase Equilibria in the Fe–N System. Materials 2021, 14, 3963. [Google Scholar] [CrossRef]
  44. Leineweber, A.; Jacobs, H.; Hüning, F.; Lueken, H.; Schilder, H.; Kockelmann, W. ε-Fe3N: Magnetic structure, magnetization and temperature dependent disorder of nitrogen. J. Alloys Compd. 1999, 288, 79–87. [Google Scholar] [CrossRef]
  45. Schwarz, B.; Göhring, H.; Meka, S.R.; Schacherl, R.E.; Mittemeijer, E.J. Pore Formation Upon Nitriding Iron and Iron-Based Alloys: The Role of Alloying Elements and Grain Boundaries. Metall. Mater. Trans. A 2014, 45, 6173–6186. [Google Scholar] [CrossRef] [Green Version]
Figure 3. (ae) Selected cross-sectional microstructures from the compound layers having formed on the indicated alloys upon nitriding in an atmosphere having the indicated nitriding potential rN for the indicated time at 813 K: Scanning electron microscopy images taken using backscattered electrons as well as maps using electron backscatter diffraction data with, respectively, phase and inverse pole-figure coloring according to the legends indicated at the bottom.
Figure 3. (ae) Selected cross-sectional microstructures from the compound layers having formed on the indicated alloys upon nitriding in an atmosphere having the indicated nitriding potential rN for the indicated time at 813 K: Scanning electron microscopy images taken using backscattered electrons as well as maps using electron backscatter diffraction data with, respectively, phase and inverse pole-figure coloring according to the legends indicated at the bottom.
Metals 12 00619 g003
Figure 4. Scanning electron microscopy with back-scattered electrons of the α-Fe region of the Fe-5Cu alloy after nitriding at rN = 2.5 atm−1/2 for 4 h at 813 K, revealing the presence of Cu-rich precipitates (cross-section; see also Figure 3d): (a) Just adjacent to the γ′ compound layer visible at the top; (b) in a depth of about 800 μm from the surface, definitely not affected by N in solid solution.
Figure 4. Scanning electron microscopy with back-scattered electrons of the α-Fe region of the Fe-5Cu alloy after nitriding at rN = 2.5 atm−1/2 for 4 h at 813 K, revealing the presence of Cu-rich precipitates (cross-section; see also Figure 3d): (a) Just adjacent to the γ′ compound layer visible at the top; (b) in a depth of about 800 μm from the surface, definitely not affected by N in solid solution.
Metals 12 00619 g004
Figure 5. EDS analysis of γ′ and ε phase on at the surface-near region of the Fe-5Cu alloy nitrided at rN = 4 atm−1/2 for 24 h at 813 K (cross-section; see also Figure 3e; covered with protective Ni prior to embedding): (a) Local EBSD phase assignment of the microstructure depicted using BSE contrast showing the location of the EDS probing. Deposited Ni appearing very bright; (b) comparison of the EDS spectra from (a): While γ′ attains a Cu content compatible with the metallic content of the substrate (uCu = 0.05), the ε phase contains less Cu.
Figure 5. EDS analysis of γ′ and ε phase on at the surface-near region of the Fe-5Cu alloy nitrided at rN = 4 atm−1/2 for 24 h at 813 K (cross-section; see also Figure 3e; covered with protective Ni prior to embedding): (a) Local EBSD phase assignment of the microstructure depicted using BSE contrast showing the location of the EDS probing. Deposited Ni appearing very bright; (b) comparison of the EDS spectra from (a): While γ′ attains a Cu content compatible with the metallic content of the substrate (uCu = 0.05), the ε phase contains less Cu.
Metals 12 00619 g005
Figure 6. Schematic metastable (ignoring the N2 gas) Fe–Cu–N phase diagram at 813 K and a pressure of 1 atm with single phase regions in blue, two-phase regions in green and three-phase reactions or tie-line triangles in red: (a) Potential phase diagram nitriding potential rN (monotonously increasing with the nitriding potential) vs. uCu = xCu/(xFe + xCu) with emphasis on the general shape of the γ′ region. The homogeneity ranges of the phases indicated in pale blue have been exaggerated in compositional extent assuming destabilization of the corresponding nitride phases with increasing uCu and thus artificially generating reactions/three-phase equilibria (imagining temperature T instead of rN as axis), of which only the reaction γ′ → ε + Cu is regarded as supported to be of that type by experiment. The colored lines on either side highlight the Lehrer diagram at 813 K as known for pure Fe (using the same colors as in Figure 2b), and as hypothesized for Cu. (b) Corresponding phase diagram at 813 K in the Gibbs compositional triangle with the extended compositional range of γ′-Fe4(1−uCu)Cu4uCuN to uCu = 0.15 corresponding to a formula Fe3.4Cu0.6N. In (b) it was refrained from exaggerating the homogeneity ranges apart from the binary extension of the ε phase and the ternary extension of the γ′ at 20 at.% N. The uCu values of the investigated Fe-1Cu and Fe-5Cu alloys are indicated in dark yellow (schematically in (a)).
Figure 6. Schematic metastable (ignoring the N2 gas) Fe–Cu–N phase diagram at 813 K and a pressure of 1 atm with single phase regions in blue, two-phase regions in green and three-phase reactions or tie-line triangles in red: (a) Potential phase diagram nitriding potential rN (monotonously increasing with the nitriding potential) vs. uCu = xCu/(xFe + xCu) with emphasis on the general shape of the γ′ region. The homogeneity ranges of the phases indicated in pale blue have been exaggerated in compositional extent assuming destabilization of the corresponding nitride phases with increasing uCu and thus artificially generating reactions/three-phase equilibria (imagining temperature T instead of rN as axis), of which only the reaction γ′ → ε + Cu is regarded as supported to be of that type by experiment. The colored lines on either side highlight the Lehrer diagram at 813 K as known for pure Fe (using the same colors as in Figure 2b), and as hypothesized for Cu. (b) Corresponding phase diagram at 813 K in the Gibbs compositional triangle with the extended compositional range of γ′-Fe4(1−uCu)Cu4uCuN to uCu = 0.15 corresponding to a formula Fe3.4Cu0.6N. In (b) it was refrained from exaggerating the homogeneity ranges apart from the binary extension of the ε phase and the ternary extension of the γ′ at 20 at.% N. The uCu values of the investigated Fe-1Cu and Fe-5Cu alloys are indicated in dark yellow (schematically in (a)).
Metals 12 00619 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Leineweber, A.; Fischer, M.; Kante, S.; Martin, S. Effect of Cu on Nitriding of α-Fe. Metals 2022, 12, 619. https://doi.org/10.3390/met12040619

AMA Style

Leineweber A, Fischer M, Kante S, Martin S. Effect of Cu on Nitriding of α-Fe. Metals. 2022; 12(4):619. https://doi.org/10.3390/met12040619

Chicago/Turabian Style

Leineweber, Andreas, Max Fischer, Stefan Kante, and Stefan Martin. 2022. "Effect of Cu on Nitriding of α-Fe" Metals 12, no. 4: 619. https://doi.org/10.3390/met12040619

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