Skip to main content
Erschienen in: Metallurgical and Materials Transactions A 11/2023

Open Access 10.09.2023 | Original Research Article

Quasi-In Situ Localized Corrosion of an Additively Manufactured FeCo Alloy in 5 Wt Pct NaCl Solution

verfasst von: Sudipta Pramanik, Jan Krüger, Mirko Schaper, Kay-Peter Hoyer

Erschienen in: Metallurgical and Materials Transactions A | Ausgabe 11/2023

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

FeCo alloys are important materials used in pumps and motors in the offshore oil and gas drilling industry. These alloys are subjected to marine environments with a high NaCl concentration, therefore, corrosion and catastrophic failure are anticipated. So, the surface dissolution of additively manufactured FeCo samples is investigated in a quasi-in situ manner, in particular, the pitting corrosion in 5.0 wt pct NaCl solution. The local dissolution of the same sample region is monitored after 24, 72, and 168 hours. Here, the formation of rectangular and circular pits of ultra-fine dimensions (less than 0.5 µm) is observed with increasing immersion time. In addition, the formation of a corrosion-inhibiting surface layer is detected on the sample surface. Surface dissolution leads to a change in the surface structure, however, no change in grain shape or grain size is noticed. The surface topography after local dissolution is correlated to the grain orientation. Quasi-in situ analysis shows the preferential dissolution of high-angle grain boundaries (HAGBs) leading to a change in the fraction of HAGBs and low-angle grain boundaries fraction (LAGBs). For the FeCo sample, a potentiodynamic polarisation test reveals a corrosion potential (Ecorr) of − 0.475 V referred to the standard hydrogen electrode (SHE) and a corrosion exchange current density (icorr) of 0.0848 A/m2. Furthermore, quasi-in situ experiments showed that grains oriented along certain crystallographic directions are corroding more compared to other grains leading to a significant decrease in the local surface height. Grains with a plane normal close to the \(\langle {1}00\rangle\) direction reveal lower surface dissolution and higher corrosion resistance, whereas planes normal close to the \(\langle {11}0\rangle\) direction and the \(\langle {111}\rangle\) direction exhibit a higher surface dissolution.
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Equiatomic FeCo alloys are important materials with potential applications in electrical motors. They have low coercivity in combination with high permeability, high Curie temperature and high saturation magnetisation. During the cooling of a FeCo alloy, the A2 (body-centred cubic) structure changes to an ordered B2 structure below 730 °C.[1,2] The ordered B2 phase is formed by the rearrangement of atoms on the lattice sites with a 0.2 pct volume change, but no change in the chemical composition. Due to the formation of the ordered B2 phase, FeCo alloys are brittle. Thus, the conventional processing (casting, rolling) of FeCo alloys is challenging. Alternatively, FeCo alloys can be processed by melt spinning or chemical vapour deposition methods. However, these processing methods cannot produce machine components in bulk form. Additive manufacturing methods such as laser powder bed fusion (LPBF) and laser engineering net shaping (LENS) can be applied to manufacture FeCo alloy parts in their final shape due to their high cooling rate (> 103 K/s).[36] The high cooling rates suppress the formation of the ordered phase or reduce the order parameter of the ordered phase[7] due to less time available for the rearrangement of atoms leading to fine grain sizes. Both of these factors contribute positively to an increase in the ductility of FeCo alloys. In previous studies, FeCo alloy powder and FeCo1.5 V alloy powder have successfully been consolidated to densities > 99 pct using additive manufacturing techniques.[4,6] Moreover, the additively manufactured FeCo samples showed a high yield strength of 600 MPa, ultimate tensile strength of 700 MPa and high ductility of 35 pct with a weak crystallographic texture.[6]
Most of the previous studies[46,8] on additively manufactured FeCo or FeCoV alloys have concentrated on the mechanical and magnetic properties. In previous studies on the corrosion of multicomponent FeCo-based alloys (HITPERM, Fe44Co44Zr7Cu1B4 alloy) in 0.1 M H2SO4 solution, a reduction in the magnetic saturation value is reported.[9,10] Upon corrosion, the formation of nonmagnetic corrosion product layers changes the direction of the magnetic lines inside the material which leads to the loss of saturation magnetisation. Thus, it is important to study the corrosion behaviour of multicomponent FeCo-based alloy as corrosion leads to a reduction in the magnetic performance of electrical machine components.[9,11] The higher the corrosion rate of the alloy, the thicker the corrosion product layer consisting of Fe2O3 and the higher the loss of magnetic saturation. In addition, the material loss due to corrosion will lead to premature failure of electrical machine components.
Induction and synchronous motors have a targeted use, for example, in the offshore oil and gas drilling industry such as in the electrical submersible pumps.[11,12] The ferromagnetic components in electrical submersible pumps are exposed to corrosive downhole well fluids containing dissolved chloride ions, carbon dioxide and hydrogen sulphide.[12] The performance of the different components (housing, shafts, pump stages, heads and bases) in the electrical submersible pumps is degraded upon exposure to the downhole well fluids. These components are also expected to be subjected to marine environments, which means a high NaCl concentration. It is well-known that chloride ions in saline solutions break down protective films and form pits which can lead to catastrophic failure. In the addressed application, a combined attack of corrosion and mechanical loads leading to stress corrosion cracking or corrosion fatigue is expected. As a FeCo alloy consists of two electrochemically different elements, the occurrence of galvanic corrosion is also anticipated. In literature, this is referred to as deironification, where a selective dissolution of Fe from the FeCo alloy occurs as Fe is electrochemically more active compared to Co.[13] Thus, investigating the localised corrosion of FeCo alloys in NaCl solution becomes important. Moreover, in the current literature, there are no studies on the pitting corrosion, stress corrosion cracking or fatigue corrosion for FeCo alloys available to the best of the author’s knowledge.
In the current investigation, the localised corrosion behaviour of additively manufactured FeCo alloys by a quasi-in situ method was undertaken. Here, a FeCo alloy (containing 50 wt pct Co) with a U-shaped notch is immersed in an aqueous NaCl solution. The U-shaped notch is used to track the same region for pre- and post-exposure microstructural investigation. Concurrently, the surface dissolution near the U-shaped notch is monitored by scanning electron microscopy (SEM) and electron backscattering diffraction (EBSD) after pre-defined periods. In addition, open circuit potential (OCP) and potentiodynamic polarization measurements are performed to determine the corrosion potential and current density. The topography after surface dissolution is compared to the grain orientation and the deviation of the plane normal to the \(\langle {1}00\rangle\) direction and the \(\langle {11}0\rangle\) direction. In previous literature, corrosion studies are performed on non-equiatomic compositions of FeCo alloy. However, in the present investigation, the corrosion behaviour is studied for an equiatomic FeCo alloy. Thus, this study is conducted to understand the effect of the alloy composition on the corrosion behaviour. Similarly, in the previous scientific studies, the corrosion behaviour is studied in acidic solutions whereas, in the present study, the corrosion behaviour is studied in neutral solutions. Thus, the effect of the corrosion medium on the corrosion behaviour of FeCo alloy is compared.

2 Experimental Procedure

2.1 Sample Processing

FeCo samples containing 50 wt pct Co is manufactured from pre-alloyed powders by LPBF. The details of the sample preparation are published in Reference 14. In brief, an SLM 250HL (SLM Solutions Group AG, Lübeck, Germany) LPBF machine equipped with a 400 W ytterbium continuous laser is utilised. The adapted processing parameters are laser power 270 W, laser scan speed 700 mm/s, hatch distance 0.11 mm, and layer thickness 0.05 mm.

2.2 Microstructural Analysis

Rectangular specimens with a dimension of 9 mm (width), 6 mm (height) and 2 mm (thickness) are cut from the as-built parts. A U-shaped notch of 2 mm (width) and 1.5 mm (depth) is wire electrical discharge machined into the samples with a 0.1 mm radius of curvature at the notch root. After grinding, polishing of the machined samples is done using a BÜHLER VibroMet polisher for 24 hours with a 50 nm colloidal silica suspension. The microstructural characterisation is performed using an SEM (Zeiss Ultra Plus) operating at 20 kV accelerating voltage and 106 pA probe current. EDS maps by an Octane Pro detector (AMETEK) are acquired to observe the change in the composition of the FeCo sample after corrosion. SEM operating at 20 kV and 13.5 mm working distance, is used to capture grain orientation maps by EBSD (DigiView 5 camera, AMETEK). All EBSD maps are acquired at 0.5 μm step size with a probe current of 77 pA and an objective aperture size of 120 μm.
Analysis of the EBSD data is carried out by the software TSL OIM and MTEX 5.6.1.[15] Initial cleaning of the EBSD map is undertaken with the software TSL OIM and exported to MTEX 5.6.1. The orientation distribution function (ODF) is calculated from the orientation data. In orientation space, the ODF represents the volume fraction of grains oriented towards a fixed orientation. ODF is calculated using the kernel density estimation technique in MTEX 5.6.1.[15] As the ODF calculation is sensitive to the kernel half-width, before performing the ODF estimation, an optimum half-width has to be calculated. Then, the orientation data is plotted on an inverse pole figure (IPF) based on the ODF calculation. Finally, from the EBSD data, the misorientation angle distributions are also calculated and compared for the samples with progressing immersion time. Further on, additional EBSD data analysis which includes the calculation of the surface energy pre-factor and the deviation of the plane normal from the \(\langle {1}00\rangle\) and \(\langle {11}0\rangle\) directions also performed in MTEX 5.6.1 by using in-house program codes.

2.3 Immersion Test

To track the progress of localised corrosion over time, immersion tests are performed in 5 wt pct NaCl solution open to air at room temperature. 5 wt pct NaCl solution is chosen as it is close to the seawater salt concentration of 3.5 wt pct.[16] The concentration is increased compared to seawater to test a more aggressive solution. For the preparation of the electrolyte, 5 g of pure NaCl is dissolved in 95 mL of deionised water. Then the polished U-notched-shaped sample is fully immersed in a cylindrical beaker with a diameter of 90 mm and a height of 20 mm containing the 5 wt pct NaCl solution. The large surface of the sample lays flat on the beaker surface. The curved notch faces the side direction. The ratio of the surface area of the electrolyte to the sample surface area is ≈ 55 mL/mm2. After pre-defined periods (24, 72, and 168 hours), the samples are removed from the solution for a quasi-in situ observation of the same area on the sample surface at each time. Therefore, the samples are rinsed with deionised water and subsequently cleaned with acetone. To determine the weight loss after each test period, the weight of the samples is measured on an XP205, Mettler Toledo weighing scale (accuracy 0.01 mg).

2.4 OCP Measurement and Potentiodynamic Polarization

OCP measurements of the FeCo sample are performed in an aqueous 5 wt pct NaCl solution with the Potentiostat MLab200 (Bank electronics). The FeCo sample with a circular surface area of 28.26 mm2 is assigned as the working electrode which is fixed to a conductive clamp and immersed in 5 wt pct NaCl solution while an Ag//AgCl electrode in saturated KCl solution (Meinsberger, Typ SE11) is assigned as the reference electrode during OCP measurement. At 25 °C, the Ag//AgCl electrode potential in saturated KCl solution is 0.198 V referred to the standard hydrogen electrode (SHE). Before the OCP measurement, the FeCo sample is ground with abrasive paper to remove oxide layers, and a further 5 minutes is waited to achieve equilibrium. During the OCP experiments, the data recording is performed at 10 Hz for 30 minutes.
The potentiodynamic polarization curve of the FeCo sample is measured in 5 wt pct NaCl solution using the Potentiostat MLab200 (Bank Electronics) in the open air at room temperature. A 28.26 mm2 circular surface area of the FeCo sample is exposed to the solution during the potentiodynamic polarization measurement. Here, a three-electrode configuration is used where, the FeCo sample is the working electrode, the Ag//AgCl electrode in saturated KCl solution (Meinsberger, Typ SE11) is the reference electrode and a platinum sheet is the counter electrode. During the scan, the applied potential starts at − 1.79 V and is increased up to 1.18 V referred to as SHE at a scan rate of 0.001 V/s, with a total scan time of 50 minutes. Before the start of the potentiodynamic polarization measurements, the FeCo sample is ground to remove the oxide layers and the experimental apparatus is allowed 5 minutes to attain a state of equilibrium.

3 Results

The OCP under equilibrium conditions is displayed in Figure 1 as it is needed to define the scanning interval for the potentiodynamic polarisation measurement. The potential is observed to drop with increasing measurement time due to the achievement of equilibrium. From Figure 1 the potential at 0 s is − 0.323 V referred to the SHE. However, the potential drifts with increasing time to − 0.37 V vs. SHE.
The potentiodynamic polarization curve of a FeCo alloy in a 5 wt pct NaCl solution is presented in Figure 2. From the potentiodynamic polarization curve, various corrosion parameters are extracted which are reported in Table I. The corrosion potential (Ecorr) referred to SHE is − 0.475 V and the corrosion current density (Icorr) is 0.0848 A/m2. Based on the Tafel extrapolation method (shown by the red dotted lines, Figure 2), the anodic and cathodic Tafel slopes are obtained as 172 mV/decade and 944 mV/decade, respectively. Icorr is calculated using the equation correlating the current (i) to the potential (E) as
$$ {\text{i}} = {\text{i}}_{{{\text{corr}}}} \left[ {{\text{exp}}\left( {\alpha {\text{nF}}\left( {{\text{E}} - {\text{E}}_{{{\text{corr}}}} } \right)/{\text{RT}}} \right) - {\text{exp}}\left( { - \left( {{1} - \alpha } \right){\text{nF}}\left( {{\text{E}} - {\text{E}}_{{{\text{corr}}}} } \right)/{\text{RT}}} \right)} \right], $$
(1)
where icorr represents the exchange current density, Ecorr is the corrosion potential, F is the Faraday constant, T is the temperature, α is the transfer coefficient (usually 0.5), and n is the number of electrons transferred.[17] A detailed review of the Tafel extrapolation method is presented in the literature.[17]
Table I
Corrosion Parameters vs. SHE Calculated via the Tafel Extrapolation Method
Ecorr (V)
Icorr (A/m2)
βanode (mV/decade)
βcathode (mV/decade)
 − 0.475
0.0848
172
944
The area near the U-shaped notch before the onset of the immersion test is presented in the SEM image in Figure 3(a). The polished sample surface shows some pores. After 24 hours immersion, the same area is observed in Figure 3(b) (SEM image). Figure 3(c) is the zoomed-in image of Figure 3(b), Figure 3(d) is the magnified image of the black rectangular area in Figure 3(c). The sample surface topography revealed the grain structure (Figure 3(d)) after 24 hours immersion. This is due to the corrosion of the FeCo alloy at the surface (i.e., selective conversion of Fe to Fe2+ ions in the aqueous solution). A previous investigation proposes the selective dissolution of Fe to Fe2+ ions (referred to as deironification) as Fe is electrochemically more active than Co.[13] This study also observes a reduction in the Fe content (≈ 50 to 47 wt pct) compared to the Co content. The composition of the FeCo alloy in the present investigation is similar to the previous study,[13] but due to different processing methods, the corrosion behaviour is different to the reported literature. The EDS spectrum from Figure 3(e) presents the entire range, however, the Figure 3(e) inset image displays the presence of an oxygen peak kα1, which indicates the adsorption of oxygen on the surface. The observed reduction in the Fe content (≈ 50 to 47 wt pct) is following the literature.[13] The presence of oxygen in the EDS spectrum is related to the formation of an oxide phase on the surface. Oxygen adsorption is due to the reduction of oxygen at the surface i.e., the conversion of dissolved oxygen to hydroxyl ions.[18] No adsorption of chloride ions is detected on the sample surface.
The SEM image of the same area (Figure 3(c)) after the 72 hours immersion test is portrayed in Figure 4(a). The zoomed-in SEM image of Figure 4(a) is shown in Figure 4(b), representing the surface topography of the grain structure. Figure 4(b) is rotated by 180 deg compared to Figure 4(a). The EBSD map from the black rectangular area (Figure 4(a)) with an average confidence index of 0.82 is presented in Figure 4(c). The EBSD map is plotted along the normal direction (ND). Most of the grains are elongated in shape, here grains 1, 2, 3, 4, 5, 6, 7 and 8 are marked in Figure 4(c) which are analysed further on. The pit is highlighted (circle and arrow) both, in Figures 4(b) and 4(c), which is used to correlate the SEM and EBSD images after 72 hours immersion.
The SEM image of the same area (Figure 3(c)) after 168 hours immersion is presented in Figure 5(a). The zoomed-in SEM image of Figure 5(a) is shown in Figure 5(b) which shows the surface topography due to surface dissolution. Figure 5(b) is rotated by 180 deg compared to Figure 5(a). Figure 5(b) represents the same area as Figure 4(b). The EBSD map from the black rectangular area (Figure 5(a)) with an average confidence index of 0.55 is illustrated in Figure 5(c). The low confidence index value is due to local surface dissolution upon immersion. Samples after exposure to the corrosive solution are expected to show a low confidence index. However, due to the quasi-in situ nature of the experiments, this is an acceptable confidence index value for the comparison of crystal orientation in the same region. The EBSD map is plotted along with ND. White areas show unindexed pixels due to the formation of a rough surface as a result of the surface dissolution. As expected, no change in grain morphology is corroborated by no modification in grain shape after 168 hours immersion. The pit is highlighted (circle and arrow) both, in Figures 5(b) and (c), which is used to correlate the SEM and EBSD images after 168 hours immersion.
The misorientation angle distribution of the FeCo sample after 72 and 168 hours immersion time is illustrated in Figure 6(a). The presence of a high fraction of low-angle grain boundaries (LAGBs) to high-angle grain boundaries (HAGBs) with bimodal distribution is visible (Figure 6(a)). With increasing immersion time, the fraction of LAGBs and HAGBs changes, however, the change is not significant. Between 45 and 55 deg, the modification of the HAGBs is not pronounced. This is probably due to the degradation of the image quality with increasing immersion time. In general, the HAGBs are expected to be attacked more due to their relatively higher energy which may lead to a change in the LAGBs and HAGBs fraction.[19] Grain boundaries have a disordered structure relative to the grain interior which results in higher surface energy resulting in a strong pitting corrosion attack.
The orientations of all the grains within the EBSD map after 72 hours immersion test are plotted on an IPF as displayed in Figure 6(b). The formation of \(\langle 0{11}\rangle\)||ND orientation with strong intensity and [\(\overline{1 }\)14]||ND orientations with weaker intensity are detected (Figure 6(b)). The orientations of all grains of the EBSD map after 168 hours immersion time are plotted on an IPF (Figure 6(c)). A slight increase in the intensity of \(\langle 0{11}\rangle\)||ND orientations along with [\(\overline{2 }\) 27]||ND orientations with weak intensity are noticed. A significant weight loss per surface area is observed after 24 hours immersion (Figure 6(d)). However, with increasing immersion time from 24 to 168 hours, only a minor change in weight is detected. This indicates the formation of a corrosion-inhibiting layer containing oxygen (detected by EDS, Figure 3(e)).
The SEM microstructures after 24, 72, and 168 hours immersion times are depicted in Figure 7. Figures 7(a) and (b) (magnified image of Figure 7(a)) represent the formation of pits after 24 hours. Rectangular pits (red arrows) of an average size of 0.15 ± 0.02 μm (length) and 0.09 ± 0.02 μm (width) are formed on the sample surface (Figure 7(b)). Here, 63 pits within an area of dimension of 21 μm × 14 μm are analysed for calculating the pit dimensions. In addition, the formation of pits after 72 hours immersion time is illustrated in Figures 7(c) and (d) (zoomed-in view of Figure 7(c)). Here, also the formation of both rectangular pits (red arrows) and circular pits (blue arrows) is detected. The length of the rectangular pits is increased by 60 pct (0.24 ± 0.07 μm) and the width is increased by 66 pct (0.15 ± 0.04 μm). The diameter of the circular pits is 0.06 ± 0.02 μm. The formation of pits after 168 hours immersion time is demonstrated in Figures 7(e) and (f) (magnified image of Figure 7(e)). Here, the size of the rectangular pits (red arrows) and circular pits (blue arrows) are observed to grow. The length of the rectangular pits increased by 8.3 pct (0.26 ± 0.05 μm) and the width of the pit increased by 20 pct (0.18 ± 0.04 μm). The diameter of the pits increased by 16.6 pct (0.07 ± 0.02 μm). The change in dimensions of the rectangular and square pits with increasing immersion time is represented in Figure 7(g). A sharp increase in length and width of the rectangular pits is seen up to 72 hours immersion time. Afterwards, the increase in length and width becomes slow. The circular pits are observed for immersion times of 72 and 168 hours, and a slow increase in the circular pit sizes is noticed. Both circular and rectangular pits are created due to the breakdown of the passive protective film by the attack of chloride ions.[20] In this study, the presence of rectangular and circular pits is observed inside the grains. The evolution of the area density of pits (number of pits per unit area) with increasing immersion time is displayed (Figure 7(h)). A sharp increase in the pit density is observed after 24 hours immersion time. However, by increasing immersion time to 72 and 168 hours, the pit density remains nearly constant. A three-stage pit density growth is reported[20] (i) nucleation of pits, (ii) rapid propagation of pits and (iii) saturation of pit density. Comparing the present results, the 24 hours immersion test is attributed to stage (ii) showing a rapid propagation of pits, whereas the 72 and 168 hours immersion tests refer to stage (iii) behaviour, where a saturation of the pit density is detected. A sigmoidal behaviour in the growth of the pit density is fitted (red curve, Figure 7(h)) which matches a previous investigation on pitting corrosion of API P110 steel.[20] However, the equation of the sigmoidal curve in the present investigation (N = 0.4t2/65792) is different to that described in the previous investigation (N = 923(1 − exp(−0.043t2))[20]), where N is the pit density and t is the time. This is due to the difference in the chemical composition of the samples and different corrosive solutions.

4 Discussion

Chansena et al.[13] have performed the potentiodynamic polarisation measurements of electrodeposited FeCo alloys with varying compositions (e.g., Fe20Co80, Fe40Co60, Fe60Co40, and Fe80Co20) in aerated deionised water and aerated acidic solution. In aerated deionised water, Ecorr is − 0.482 V vs. SHE for Fe40Co60 and − 0.446 V vs. SHE for Fe60Co40.[13] From the current investigation, in 5 wt pct NaCl solution, Ecorr is − 0.475 V vs. SHE for Fe50Co50. So, Fe50Co50 alloy (in 5 wt pct NaCl solution) is electrochemically less active than the Fe40Co60 alloy. However, Fe50Co50 is electrochemically more active than Fe60Co40 (in aerated deionised water). The OCP of the investigated alloy is between the OCP of the alloys presented by Chansena et al.[13] Thus, the measured OCP is following this study. The corrosion behaviour and OCP depend not only on the alloy composition but on the test solution, as well. The corrosion rate of Fe50Co50 in 5 wt pct NaCl solution is 0.183 mm/year (calculated from Icorr). For comparison in aerated deionised water, the corrosion rate of Fe40Co60 is 0.0048 mm/year, whereas, for Fe60Co40, the corrosion rate is 0.0051 mm/year.[13] Thus, the addition of 5 wt pct NaCl leads to a huge increase in the corrosion rate for iron-cobalt alloys.
In an aerated acidic solution (pH 3) the corrosion rate increases to 2.4 mm/year (2.11 A/m2) for Fe40Co60 and 3.1 mm/year (2.76 A/m2) for Fe60Co40.[13] Thus, the aerated acidic solution makes the FeCo alloys electrochemically more active, leading to a high corrosion rate. Another important observation is, that the corrosion rate increases by increasing iron content in FeCo alloys [i.e., in aerated acidic solution (pH 3) the corrosion rate of Fe80Co20 (7.7 mm/year, 6.67 A/m2) is higher than for Fe20Co80 (1.1 mm/year)].[13] As Fe (OCP =  − 0.447 V vs. SHE) is electrochemically more active than Co (OCP =  − 0.277 V vs. SHE), an increase in the iron content makes the alloy electrochemically more active leading to a higher corrosion rate. For the current investigation, a neutral pH solution is selected leading to a corrosion rate (0.0848 A/m2) significantly lower than an aerated acidic solution. Under this investigation, the corrosion behaviour of FeCo alloys is studied in general. So, a neutral pH solution is selected to study the corrosion behaviour of the FeCo alloy.

4.1 Correlation of Surface Topography and Grain Orientation

During corrosion, atoms from the surface are removed and dissolved as ions in the surrounding solution. This process requires the breaking of bonds with the atoms beneath. Thus an energy barrier between the reactants (surface atoms) and the products (ions) must be overcome.[21,22] Broken bonds are always present on the surface. The more broken bonds are present, the higher the surface energy and surface energy pre-factor. Since, the dissolution of atoms is easier, when more broken bonds are present, a high surface energy pre-factor indicates preferred dissolution.
In the present study, the local surface dissolution (conversion of Fe and Co to aqueous ions) is detected leading to the formation of a rough surface (Figures 3(a) and 3(b)). This local surface dissolution is strongly related to the density of the broken atomic bonds at the grain surface.[23] As the broken atomic bond density at the grain surface is dependent on the surface energy of the grains, which in turn depends on the grain orientation ({hkl}\(\langle {\text{uvw}}\rangle\)), the surface dissolution rate is affected by the grain orientation. The surface energy due to the presence of uncompleted bonds on the sample surface is correlated to the Miller indices of the surface plane normal {hkl}, bond energy and bond length by Eq. [6] in Reference 23. The surface energy of grains with the orientation ({hkl}\(\langle {\text{uvw}}\rangle\)) depends on the (i) surface energy pre-factor1\(\frac{2\left|\mathrm{h}\right|+\mathrm{k}}{\sqrt{{\mathrm{h}}^{2}+{\mathrm{k}}^{2}+{\mathrm{l}}^{2}}}\), (ii) the bond energy, and (iii) the bond length.[23] As the bond energy and length are not expected to change with the immersion time, here, the surface energy pre-factor of the grains is plotted on the EBSD map (Figure 8(a)). Intuitively, grains with high surface energy pre-factor values (yellow colored, Figure 8(a)) are expected to undergo faster surface dissolution as compared to grains with low surface energy pre-factor values (blue colored, Figure 8(a)). To determine the effect of the grain orientation on the surface dissolution, the deviation of the plane normal to the \(\langle {1}00\rangle\) direction is plotted on the EBSD map of one sample after 72 hours immersion time (Figure 8(b)). Here, the yellow-colored grains exhibit a very high deviation (≈ 50 deg) of the plane normal from the \(\langle {1}00\rangle\) direction, whereas blue-colored grains have a low deviation of the plane normal from the \(\langle {1}00\rangle\) direction.
Cropped SEM image, surface energy pre-factor of grain 1, and plane normal deviation from the \(\langle {1}00\rangle\) direction of grain 1 are plotted in Figure 9. Here, from the SEM image, the surface dissolution of grain 1 is lower (less material is removed) compared to the surrounding grains. This indicates a lower corrosion rate. The surface energy pre-factor of grain 1 is plotted on the EBSD map which presents a high value (2.17). Hence, the surface energy of grain 1 is expected to be high relative to other grains. Nevertheless, the corrosion rate of grain 1 is low, indicating that the surface energy is not the only factor that influences the corrosion rate. Grains with high surface energy adsorb a high amount of solvated ions.[24,25] Due to the impediment of the surface sites by ions, contact between the surface and the corrosive solution is prevented.[26] This can lead to a decrease in the corrosion rate for surfaces with high surface energies. A similar observation has been made for 316 stainless steel (316L),[23] where {111} plane normal grains exhibit the highest corrosion rates despite having low surface energy. However, from the orientation of grain 1, the mean orientation is calculated (2 \(\overline{9 }\) 0)[\(\overline{3 } \overline{1 } \overline{10 }\)]. The angular deviation of the plane normal (2\(\overline{9 }\)0) from the \(\langle {1}00\rangle\) direction is plotted on the EBSD map. Grain 1’s plane normal has a low deviation from the \(\langle {1}00\rangle\) direction. This proves that grains with a plane normal oriented close to the \(\langle {1}00\rangle\) direction exhibit low surface dissolution (low corrosion rate). This conclusion is verified for three \(\langle {1}00\rangle\) oriented grains. One example of an \(\langle {1}00\rangle\) oriented grain is presented in Figure 9.
The cropped SEM image, surface energy pre-factor and plane normal deviation from the \(\langle {11}0\rangle\)\(\langle {11}0\rangle\) direction of grain 2 is presented in Figure 10. Here, from the SEM image, the local surface elevation of grain 2 is lower than that of the surrounding grains. This indicates a high surface dissolution of grain 2 (higher corrosion rate). For grain 2, the surface energy pre-factor is high (2.5) indicating a higher number of broken surface bonds. So, here the opposite effect occurs as compared to grain 1 in which high surface energy pre-factor leads to a lower corrosion rate. The surface energy pre-factor reasoning may not apply to grain 2 orientated along \(\langle {11}0\rangle\) direction. From the orientation data of grain 2, the mean orientation is calculated (\(\overline{10 } \overline{1 } 11\))[6 11 7]. The angular deviation of the plane normal (\(\overline{10 } \overline{1 } 11\)) from the \(\langle {11}0\rangle\) direction is plotted on the EBSD map of grain 2. Grain 2’s plane normal has a low deviation from the \(\langle {11}0\rangle\) direction. This shows that grains with a plane normal close to the \(\langle {11}0\rangle\) direction have higher surface dissolution (higher corrosion rate). This conclusion is verified for three \(\langle {11}0\rangle\) oriented grains (grain 2, grain 7 and grain 8 in Figures 4(c) and 5(c)). In Figure 10, an example of one \(\langle {11}0\rangle\) oriented grain 2 (Figures 4(c) and 5(c)) is depicted.
The cropped SEM image, surface energy pre-factor and plane normal deviation from the \(\langle {111}\rangle\) direction of grain 3 is displayed in Figure 11. As can be seen from the SEM image, the local surface elevation of grain 3 is lower than the surrounding grains. This indicates a high surface dissolution of grain 3 (higher corrosion rate). The surface energy pre-factor of grain 3 is 2.1 (slightly lower compared to grain 1 and grain 2) indicating a high number of surface-broken bonds. So, an opposite effect occurs as compared to grain 1 in which high surface energy pre-factor leads to a lower corrosion rate. Thus, surface energy pre-factor reasoning shall not apply to grain 3 oriented along the \(\langle {111}\rangle\) direction. The mean orientation of grain 3 from the orientation data is (4\(\overline{2 }\overline{3 }\))[111]. The angular deviation of the plane normal (15°) from the \(\langle {111}\rangle\) direction is shown on the EBSD map of grain 3. Grain 3 with a plane normal close to the \(\langle {111}\rangle\) direction has higher surface dissolution (higher corrosion rate). This conclusion is checked from three \(\langle {111}\rangle\) oriented grains. In Figure 11, an example of one \(\langle {111}\rangle\) oriented grain is depicted.

4.2 Effect of Grain Orientation on Surface Dissolution

The EBSD maps of grain 4 (marked in Figures 4(c) and 5(c)) after 72 and 168 hours immersion time are presented in Figures 12(a) and (b). A slight change in the shape of grain 4 is observed for the immersion time between 72 and 168 hours due to surface dissolution. Nonetheless, no change in the size of grain 4 is detected with increasing immersion time (Figure 12(a)). The orientations of grain 4 on an IPF are shown in Figure 12(b).
Consider grain 5 and grain 6 from the EBSD map after 168 hours dissolution (Figure 5(c)). The EBSD IQ and EBSD IPF maps are shown in Figures 13(a) and (b). The orientations of grain 5 and grain 6 are plotted on the inverse pole figure plot (Figure 13(c)). The orientation of grain 5 is different to grain 6 (rotation of 40.6 about [\(\overline{1 }\)02] axis). The average EBSD IQ of grain 5 (69176) is lower than the average EBSD IQ of grain 6 (76188). Grains oriented along certain directions show increased dissolution and consequently, the surface quality of these grains decreases further compared to other-oriented grains. Consequently, these grain orientations are less represented in the pole figure due to poor identification of these areas. Thus, it can be stated that certain directions are strongly affected by corrosion.
At last, a comparison of different grain orientations on the corrosion behaviour of different alloy compositions is undertaken. The effect of the grain orientation on the surface dissolution behaviour depends on the alloy composition and the corrosive solution. As an example, the FeCo alloy shows a higher surface dissolution with a plane normal orientated close to the \(\langle {11}0\rangle\) direction and \(\langle {111}\rangle\) direction. However, 316L steel shows the highest surface dissolution for only the plane normal close to the \(\langle {111}\rangle\) direction immersed in 10 wt pct oxalic acids for 45 sesonds with an imposed constant current density of 1 A/m2.[23] A Ni-22Cr–13Mo–3W–3Fe alloy, when exposed to 3 M HCl + 1 m NaCl solution, shows the plane normal close to the \(\langle {11}0\rangle\) direction having the highest surface dissolution.[27] Whereas, the Ni–22Cr–13Mo–3W–3Fe alloy when exposed to 1 M HCl + 1 m NaCl solution, it was found that the plane normal close to the \(\langle {11}0\rangle\) direction has the highest surface dissolution.[27] Thus, the grain orientation is decisive for degradation behaviour. On the opposite, a less decisive impact of the surface energy prefactor is observed, in the present study.

5 Conclusions

Based on the quasi-in situ immersion test of the additively manufactured FeCo sample up to 168 hours, the following conclusions can be drawn:
1.
Upon increasing immersion time, a mixture of surface dissolution and pitting corrosion is detected on the sample surface. Adsorption of oxygen on the sample surface is spotted which may be due to a thin layer of corrosion products on the sample surface after immersion. Using potentiodynamic polarization measurements and the Tafel extrapolation method, the corrosion potential (Ecorr) referred to SHE is − 0.475 V and the corrosion current density (icorr) is 0.0848 A/m2.
 
2.
The formation of rectangular and circular pits is noticed on the sample surface. Both types of pits have ultra-fine dimensions (size less than 0.5 µm) and grow with increasing immersion time.
 
3.
Quasi-in situ observations of the same area revealed that with increasing immersion time (from 72 to 168 hours) the fraction of LAGBs and HAGBs does not change significantly.
 
4.
Quasi-in situ experiments highlight that grains oriented along certain crystallographic directions show increased dissolution resulting in the lower surface quality of these grains compared to other-oriented grains. Consequently, these orientations are represented less in the inverse pole figure due to decreased identification of these grains.
 
5.
Grains with a plane normal oriented close to the \(\langle {1}00\rangle\) direction reveal lower surface dissolution and high corrosion resistance whereas grains with a plane normal oriented close to the \(\langle {11}0\rangle\) or \(\langle {111}\rangle\) directions manifest a higher surface dissolution and thus, lower corrosion resistance. So, by adapting the LPBF processing parameters favourable grain orientations can be created making the FeCo alloy more corrosion-resistant. A targeted adaption of orientation by modification of the LBM process is possible. Thus, the corrosion resistance can be improved by the adaption of the LBM process and resulting grain orientation.
 

Acknowledgments

The authors would like to thank Thomas Janzen for performing the EBSD measurements.

Data Availability

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also forms part of an ongoing study.

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Fußnoten
1
The method of calculation of surface energy pre-factor is explained in Reference 23.
 
Literatur
3.
Zurück zum Zitat A.B. Kustas, C.M. Fancher, S.R. Whetten, D.J. Dagel, J.R. Michael, and D.F. Susan: Addit. Manuf., 2019, vol. 28, pp. 772–80. A.B. Kustas, C.M. Fancher, S.R. Whetten, D.J. Dagel, J.R. Michael, and D.F. Susan: Addit. Manuf., 2019, vol. 28, pp. 772–80.
4.
Zurück zum Zitat A.B. Kustas, D.F. Susan, K.L. Johnson, S.R. Whetten, M.A. Rodriguez, D.J. Dagel, J.R. Michael, D.M. Keicher, and N. Argibay: Addit. Manuf., 2018, vol. 21, pp. 41–52. A.B. Kustas, D.F. Susan, K.L. Johnson, S.R. Whetten, M.A. Rodriguez, D.J. Dagel, J.R. Michael, D.M. Keicher, and N. Argibay: Addit. Manuf., 2018, vol. 21, pp. 41–52.
5.
Zurück zum Zitat T.F. Babuska, K.L. Johnson, T. Verdonik, S.R. Subia, B.A. Krick, D.F. Susan, and A.B. Kustas: Addit. Manuf., 2020, vol. 34, p. 101187. T.F. Babuska, K.L. Johnson, T. Verdonik, S.R. Subia, B.A. Krick, D.F. Susan, and A.B. Kustas: Addit. Manuf., 2020, vol. 34, p. 101187.
6.
Zurück zum Zitat T.F. Babuska, M.A. Wilson, K.L. Johnson, S.R. Whetten, J.F. Curry, J.M. Rodelas, C. Atkinson, P. Lu, M. Chandross, B.A. Krick, J.R. Michael, N. Argibay, D.F. Susan, and A.B. Kustas: Acta Mater., 2019, vol. 180, pp. 149–57.CrossRef T.F. Babuska, M.A. Wilson, K.L. Johnson, S.R. Whetten, J.F. Curry, J.M. Rodelas, C. Atkinson, P. Lu, M. Chandross, B.A. Krick, J.R. Michael, N. Argibay, D.F. Susan, and A.B. Kustas: Acta Mater., 2019, vol. 180, pp. 149–57.CrossRef
7.
Zurück zum Zitat D.W. Clegg and R.A. Buckley: Met. Sci. J., 1973, vol. 7, pp. 48–54.CrossRef D.W. Clegg and R.A. Buckley: Met. Sci. J., 1973, vol. 7, pp. 48–54.CrossRef
8.
Zurück zum Zitat V. Chaudhary, N.M. Sai Kiran Kumar Yadav, S.A. Mantri, S. Dasari, A. Jagetia, R.V. Ramanujan, and R. Banerjee: J. Alloys Compd., 2020, vol. 823, p. 153817.CrossRef V. Chaudhary, N.M. Sai Kiran Kumar Yadav, S.A. Mantri, S. Dasari, A. Jagetia, R.V. Ramanujan, and R. Banerjee: J. Alloys Compd., 2020, vol. 823, p. 153817.CrossRef
9.
Zurück zum Zitat J.E. May, C.A.C. Souza, C.L. Morelli, N.A. Mariano, and S.E. Kuri: J. Alloys Compd., 2005, vol. 390, pp. 106–11.CrossRef J.E. May, C.A.C. Souza, C.L. Morelli, N.A. Mariano, and S.E. Kuri: J. Alloys Compd., 2005, vol. 390, pp. 106–11.CrossRef
10.
Zurück zum Zitat J.E. May, M.F. De Oliveira, and S.E. Kuri: J. Alloys Compd., 2004, vol. 369, pp. 121–24.CrossRef J.E. May, M.F. De Oliveira, and S.E. Kuri: J. Alloys Compd., 2004, vol. 369, pp. 121–24.CrossRef
11.
Zurück zum Zitat N. Yogal and C. Lehrmann: Int. J. Electr. Comput. Electron. Commun. Eng., 2014, vol. 8. N. Yogal and C. Lehrmann: Int. J. Electr. Comput. Electron. Commun. Eng., 2014, vol. 8.
12.
Zurück zum Zitat B.Q. Han: in CORROSION 2018, Phoenix, Arizona, USA, April 2018, OnePetro, 2018, p. NACE-2018-11296. B.Q. Han: in CORROSION 2018, Phoenix, Arizona, USA, April 2018, OnePetro, 2018, p. NACE-2018-11296.
13.
Zurück zum Zitat A. Chansena and S. Sutthiruangwong: J. Magn. Magn. Mater., 2017, vol. 429, pp. 251–56.CrossRef A. Chansena and S. Sutthiruangwong: J. Magn. Magn. Mater., 2017, vol. 429, pp. 251–56.CrossRef
14.
Zurück zum Zitat S. Pramanik, L. Tasche, K.-P. Hoyer, and M. Schaper: Addit. Manuf., 2021, vol. 46, p. 102087. S. Pramanik, L. Tasche, K.-P. Hoyer, and M. Schaper: Addit. Manuf., 2021, vol. 46, p. 102087.
15.
Zurück zum Zitat R. Hielscher and H. Schaeben: J. Appl. Crystallogr., 2008, vol. 41, pp. 1024–37.CrossRef R. Hielscher and H. Schaeben: J. Appl. Crystallogr., 2008, vol. 41, pp. 1024–37.CrossRef
16.
Zurück zum Zitat D. Rana, Y. Kim, T. Matsuura, and H.A. Arafat: J. Membr. Sci., 2011, vol. 367, pp. 110–18.CrossRef D. Rana, Y. Kim, T. Matsuura, and H.A. Arafat: J. Membr. Sci., 2011, vol. 367, pp. 110–18.CrossRef
17.
18.
Zurück zum Zitat U.K. Chatterjee, S.K. Bose, and S.K. Roy: Environmental Degradation of Metals: Corrosion Technology Series/14, CRC Press, Boca Raton, 2001.CrossRef U.K. Chatterjee, S.K. Bose, and S.K. Roy: Environmental Degradation of Metals: Corrosion Technology Series/14, CRC Press, Boca Raton, 2001.CrossRef
19.
Zurück zum Zitat J.B. GaetanoPalumbo, D. Dunikowski, R. Wirecka, T. Mazur, U. Lelek-Borkowska, and K. Wawer: Materials (Basel), 2021, vol. 14, p. 5084.CrossRef J.B. GaetanoPalumbo, D. Dunikowski, R. Wirecka, T. Mazur, U. Lelek-Borkowska, and K. Wawer: Materials (Basel), 2021, vol. 14, p. 5084.CrossRef
20.
Zurück zum Zitat P. Guo, E.C. La Plante, B. Wang, X. Chen, M. Balonis, M. Bauchy, and G. Sant: Sci. Rep., 2018, vol. 8, pp. 1–12. P. Guo, E.C. La Plante, B. Wang, X. Chen, M. Balonis, M. Bauchy, and G. Sant: Sci. Rep., 2018, vol. 8, pp. 1–12.
21.
Zurück zum Zitat W. Peng, W. Zeng, Q. Wang, and H. Yu: Mater. Des., 2013, vol. 51, pp. 95–104.CrossRef W. Peng, W. Zeng, Q. Wang, and H. Yu: Mater. Des., 2013, vol. 51, pp. 95–104.CrossRef
22.
Zurück zum Zitat A. He, G. Xie, H. Zhang, and X. Wang: Mater. Des., 2013, vol. 52, pp. 677–85.CrossRef A. He, G. Xie, H. Zhang, and X. Wang: Mater. Des., 2013, vol. 52, pp. 677–85.CrossRef
23.
Zurück zum Zitat S. Dong, X. Chen, E.C. La Plante, M. Gussev, K. Leonard, and G. Sant: Mater. Des., 2020, vol. 191, p. 108583.CrossRef S. Dong, X. Chen, E.C. La Plante, M. Gussev, K. Leonard, and G. Sant: Mater. Des., 2020, vol. 191, p. 108583.CrossRef
24.
Zurück zum Zitat Z.Y. Zhou, N. Tian, J.T. Li, I. Broadwell, and S.G. Sun: Chem. Soc. Rev., 2011, vol. 40, pp. 4167–85.CrossRef Z.Y. Zhou, N. Tian, J.T. Li, I. Broadwell, and S.G. Sun: Chem. Soc. Rev., 2011, vol. 40, pp. 4167–85.CrossRef
25.
Zurück zum Zitat L. Xu, H.W. Liang, Y. Yang, and S.H. Yu: Chem. Rev., 2018, vol. 118, pp. 3209–50.CrossRef L. Xu, H.W. Liang, Y. Yang, and S.H. Yu: Chem. Rev., 2018, vol. 118, pp. 3209–50.CrossRef
27.
Zurück zum Zitat J.J. Gray, B.S. El Dasher, and C.A. Orme: Surf. Sci., 2006, vol. 600, pp. 2488–94.CrossRef J.J. Gray, B.S. El Dasher, and C.A. Orme: Surf. Sci., 2006, vol. 600, pp. 2488–94.CrossRef
Metadaten
Titel
Quasi-In Situ Localized Corrosion of an Additively Manufactured FeCo Alloy in 5 Wt Pct NaCl Solution
verfasst von
Sudipta Pramanik
Jan Krüger
Mirko Schaper
Kay-Peter Hoyer
Publikationsdatum
10.09.2023
Verlag
Springer US
Erschienen in
Metallurgical and Materials Transactions A / Ausgabe 11/2023
Print ISSN: 1073-5623
Elektronische ISSN: 1543-1940
DOI
https://doi.org/10.1007/s11661-023-07186-7

Weitere Artikel der Ausgabe 11/2023

Metallurgical and Materials Transactions A 11/2023 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.