Next Article in Journal
Two New Red/Near-Infrared Ir(Ⅲ) Complexes with Reversible and Force-Induced Enhanced Mechanoluminescence
Next Article in Special Issue
Effects of Hot Extrusion Temperature Conditions on the Hardness and Electrical Conductivity of Rapidly Solidified Al-Fe Alloys
Previous Article in Journal
Polymeric Coatings for Magnesium Alloys for Biodegradable Implant Application: A Review
Previous Article in Special Issue
Effect of Versenium Hydrogensulfate on Properties of Nickel Coatings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Corrosion Resistance Enhancement of CoCrFeMnNi High-Entropy Alloy with WC Particle Reinforcements via Laser Melting Deposition

1
School of Materials Science and Engineering, Jiangsu University, Zhenjiang 212013, China
2
Institute of Materials, China Academy of Engineering Physics, Mianyang 621907, China
3
Research Institute of Huazhong University of Science and Technology in Shenzhen, Shenzhen 518057, China
4
State Key Laboratory of Materials Processing and Die & Mould Technology, School of Materials Science and Engineering, Huazhong University of Science and Technology, No. 1037 Luoyu Road, Wuhan 430074, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(13), 4701; https://doi.org/10.3390/ma16134701
Submission received: 16 May 2023 / Revised: 26 June 2023 / Accepted: 27 June 2023 / Published: 29 June 2023
(This article belongs to the Special Issue Physical Metallurgy of Metals and Alloys II)

Abstract

:
In the present work, a WC particle-reinforced CoCrFeMnNi high-entropy alloy (HEA) was fabricated by laser melting deposition (LMDed). The LMDed CoCrFeMnNi high-entropy alloy (CoCrFeMnNi) composite is primarily comprised of a face-centered cubic (FCC) crystal structure. However, in the case of CoCrFeMnNi with 2.5 wt.% WC, it exhibits a combination of an FCC matrix and a ceramic phase known as M23C6. The corrosion behavior of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC particle in 0.5 M H2SO4 was comparatively investigated. Compared with CoCrFeMnNi, the passive film formed on the CoCrFeMnNi with 2.5 wt.% WC had a more stable and stronger protective property. The corrosion current density of the CoCrFeMnNi with 2.5 wt.% WC dropped by 149.1% compared to that of the CoCrFeMnNi, indicating that the CoCrFeMnNi with 2.5 wt.% WC had better corrosion resistance than that of the CoCrFeMnNi.

1. Introduction

The development of high-entropy alloys (HEAs) has drawn significant interest since the pioneering work in 2004 by Yeh et al. and Cantor et al. [1,2]. In contrast to the conventional method with only one dominant element in the alloy, HEAs are composed of multiple elements with near-equal atomic percentages. Due to the “four core effects” defined by Yeh, HEAs can exhibit remarkable mechanical and functional properties, such as excellent thermal stability, wear, and oxidation resistance corrosion resistance [3,4,5,6,7,8,9,10].
Traditional manufacturing processes, such as vacuum arc melting, mechanical alloying, and powder metallurgy, have been extensively used to fabricate HEAs [11,12,13,14,15]. However, these HEAs usually have restricted shapes and coarse grains, which limits the wide application of HEAs. In recent years, several studies on additively manufactured (AM) HEAs have been carried out [16,17,18,19,20]. Additive manufacturing technology offers a rapid and efficient approach to fabricate alloys with gradient or complex shapes, making it a valuable tool in advancing the development of high-entropy alloys. This technique possesses several advantageous features, including unrestricted forming size and structure, free-forming capabilities, net shaping, and precise manufacturing. In contrast to conventional preparation methods, AM provides better control over structural uniformity and enables the production of ultra-fine grains, leading to enhanced overall mechanical properties of WC-containing HEAs. As a result, AM holds significant potential in ensuring the structural integrity and enhancing the comprehensive mechanical performance of WC-containing HEAs, contributing to their advancement and application. As one of the most widely used AM methods, laser melting deposition (LMD) can fabricate metal parts in complex shapes with high precision and excellent performance [21,22,23,24,25,26].
The CoCrFeMnNi high-entropy alloy is a well-researched material known for its outstanding mechanical properties, particularly at cryogenic temperatures. This is attributed to the activation of diverse deformation mechanisms, including dislocation and twin-mediated processes, within its single-face FCC structure. Notably, studies have revealed that additive manufacturing techniques can further enhance the CoCrFeMnNi HEA by producing finer grain sizes and increasing its overall strength, surpassing the properties of conventionally manufactured counterparts. In order to advance the comprehensive properties of high-entropy alloys, there is a continual need for novel modification methods. In recent times, the incorporation of ceramic particles such as carbides and nitrides into HEAs have emerged as a promising approach to enhance their properties. This strategy enables the customization of structures and facilitates the synergistic combination of mechanical and chemical attributes in HEAs. By incorporating ceramic particles, researchers aim to achieve optimized performance and further unlock the potential of these materials [4,27,28,29,30,31,32].
In this study, a CoCrFeMnNi high-entropy alloy with 2.5 wt.% WC particles was fabricated using laser melting deposition. A comparative analysis was conducted to examine the electrochemical corrosion behavior of the CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC particles in a 0.5 M H2SO4 environment. The aim was to gain insights into the corrosion resistance enhancement. The investigation included microstructural characterization and examination of the composition uniformity of the alloy. By comprehending these factors, the underlying reasons behind the improved corrosion resistance could be identified in this work.

2. Materials and Methods

2.1. Sample Preparation

Laser melting deposition (LMD) was utilized to fabricate the CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC high-entropy alloy particles. The average particle sizes of pre-alloyed CrMnFeCoNi powder and WC powder were analyzed by Microtrac S3500 (Microtrac, Largo, FL, USA) laser particle size analyzer, which were approximately 120 μm and 10 μm, respectively. The two powders are mixed according to the designed proportion. The mixed powder is heated to 80 °C and dried for 2 h and then cooled to room temperature in a vacuum chamber before use. The mixed powder is transported to the laser molten pool through a closed loop powder supply unit, and deposited continuously on 316 L stainless steel substrate by a reciprocating multi-layer scanning under 1000 W laser power and 500 mm/min scanning speed. During deposition, the atmosphere was under the protection of argon, and the oxygen content in the room was below 20 ppm. The mixed powder is delivered to the chamber through the coaxial nozzle at an argon flow rate of 15–18 L/min and a feed rate of 7–9 g/min. The width of a single deposition track was approximately 35 mm, the thickness was approximately 4 mm. After each layer was deposited, the laser head would rise to a certain height until the deposition height reaches 40 mm.

2.2. Microstructural Characterization

Electric spark corrosion was used to cut thin-walled samples with at least 4 mm away from the substrate. Samples were polished with sandpaper from 200 #, 400 #, 600 #, to 3000 #, followed by mechanical polishing until there were no obvious scratches under a 400× optical microscope. X-ray diffraction (XRD) analysis of the particles’ crystalline structure was conducted using a Japan Nigaku D/max-RB X-ray diffraction spectrometer equipped with Cu-Kα radiation. The scanning angle for the analysis ranged from 15° to 90°. The size and microstructure of samples are characterized by S-4800 SEM and JEM 200CX transmission electron microscope (TEM) with 200 kV operating voltage. Aqua regia was used for corrosion before SEM testing. For TEM testing, the samples were grinded to below 100 µm, and then punched a hole of Φ3 mm. The ion thinning method was used to thin them until met the requirements of TEM testing. JEM 200CX transmission electron microscope (TEM) with 200 kV operating voltage. The composition of the etched passivation film was characterized using FEI EscaLab 250Xi X-ray photoelectron spectroscopy (XPS).

2.3. Electrochemical Measurements

The electrochemical workstation was used to characterize the corrosion resistance of the alloy. The corrosion medium was 0.5 M H2SO4 solution. The test surface size of 10 mm × 10 mm, the back side of the test surface is connected to the copper wire using conductive adhesive, and the other surfaces are wrapped and sealed with epoxy resin to avoid contacting with the test solution. During the test, a three-electrode system is selected, with the platinum plate electrode as the auxiliary electrode. The connecting wire of the sample is encapsulated with rubber resin, and the surface exposed for the test is used as the working electrode. The potentiodynamic polarization curve and AC impedance curve of the alloy was obtained. The scanning rate was 3 mv/s during the action potential polarization test. When conducting the AC impedance test, the test frequency range is 10−2–106 Hz, and the amplitude is 5 mv.

3. Results and Discussion

3.1. Microstructure before Corrosion

Figure 1a illustrates the X-ray diffraction (XRD) analysis results of the LMDed fabricated CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC high-entropy alloy samples. The two alloys show a single-phase FCC structure, and the peak of WC is not found in the XRD results of CoCrFeMnNi with 2.5 wt.% WC samples. Figure 1b is an enlarged view of the (111) peak. It can be seen from the figure that the main peak of the CoCrFeMnNi with 2.5 wt.% WC sample shifted significantly to the left side. According to the Bragg equation (2dsin θ = λ), the observed difference in lattice constants between the LMDed CoCrFeMnNi sample with 2.5 wt.% WC (a0 = 3.6032 Å) and the CoCrFeMnNi sample (a0 = 3.6007 Å) can be attributed to the incorporation of WC particles into the HEA matrix. The addition of WC can introduce lattice strain and result in a slight expansion of the lattice. The presence of WC particles may cause lattice distortion and contribute to the observed increase in lattice constant. It is worth noting that the difference in lattice constant may also be influenced by factors such as processing conditions, composition variations, and the distribution of WC particles within the HEA matrix [22].
Figure 2 shows the SEM images and EDS mapping of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC. It can be seen from the SEM image that is a mixed structure of dendritic and cytosolic crystals, while CoCrFeMnNi with 2.5 wt.% WC are mostly columnar crystals. Compared with CoCrFeMnNi, CoCrFeMnNi with 2.5 wt.% WC shows a finer grain size. As can be seen from the EDS mapping in Figure 2c, Mn and Fe elements in the CoCrFeMnNi are separated, which is consistent with the results of Cantor and Salishchev et al. [2,33]. The red border in the Figure 2 shows the enriched part of Mn, while the white border shows the poor part of Mn. The segregation degree of Fe is lower than that of Mn, the difference of element content between rich Fe region and poor Fe region is smaller, and the size of enrichment region is smaller than Mn. Element segregation also exists in the CoCrFeMnNi with 2.5 wt.% WC samples prepared by laser melting deposition. As can be seen from the EDS mapping in Figure 2d, Cr; Mn and Fe elements in the CoCrFeMnNi with 2.5 wt.% WC are separated. In the laser melting deposition process, the extremely fast heating speed brings about a great temperature gradient from the substrate to the cladding layer. In addition, due to the uneven distribution of laser radiation energy, the convection of molten pool is caused during cladding. Both may lead to composition segregation. The results of EDS point scan tests on part A and B are shown in Table 1. The CoCrFeMnNi with 2.5 wt.% WC samples have more significant component segregation, which may be due to the decomposition of WC and the formation of solid solutions.
In order to further analyze the effect of WC addition on the microstructure of the alloy, TEM and HAADF-STEM technology have been applied, as shown in Figure 3. According to Figure 3a, some nanoscale particles can be found in HEA matrix, the corresponding selected area diffraction pattern, as illustrated in Figure 3a, revealed that the precipitations were M23C6 carbides. The HAADF-STEM technology was used to characterize the particle in the red frame, the results are shown in Figure 3b. The particle is mainly composed of Cr, Mn, and W elements. This indicates that the particles in the matrix are precipitates related to the decomposition of WC and the formation of solid solutions. After the decomposition of WC, both W and C elements are solidly dissolved into the matrix, and some W elements also form nano precipitates with other elements. The EDS point scan results in Table 1 show that the alloy composition at B location contains high Cr and C elements, with a high probability of Cr carbides. However, the results of HAADF in Figure 3 show that M23C6-type carbides rich in Cr, Mn, and W appear in LMDed HEA with 2.5 wt.% WC alloy, which corresponds to the EDS results. From Figure 2(b2), it can be observed that the particle (M23C6, M is Cr, Mn, W) mainly appears at the grain boundary, this may lead to the formation of Cr poor zone along the grain boundary.

3.2. Electrochemical Corrosion Properties

The potentiodynamic polarization curve test is an effective method to evaluate the corrosion behavior of materials. The polarization curves of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC in 0.5 M H2SO4 are shown in Figure 4a. The characteristic parameters used to describe corrosion properties can be obtained from Figure 4a. In this work, Taffel’s extrapolation method is used to calculate the results of the polarization curves. The fitting electrochemical parameters are shown in Table 2. The corrosion potential (Ecoor) represents the corrosion potential of a material in the open circuit condition, and Icorr represents the corrosion current density. It is clear from both the fitting parameters and the potentiodynamic polarization curves of the alloys that CoCrFeMnNi with 2.5 wt.% WC has higher Ecorr and lower Icorr, which represent higher corrosion resistance. As can be seen from Figure 4a, both samples exhibit strong “activation-passivation” behavior, with wide primary passivation intervals and secondary passivation phenomena. In order to investigate the passivation process, AC electrochemical impedance tests were carried out, and the results were shown in the form of Nyquist diagram in Figure 4b. The Nyquist plots of samples are semicircular capacitance arcs. Generally, the larger the curvature radius of the arc, the stronger the corrosion resistance. It can be seen from the figure that the semi-arc of CoCrFeMnNi with 2.5 wt.% WC has a larger curvature radius, indicating its better corrosion resistance, which is consistent with the results of the polarization curve.
As Warburg impedance in the low-frequency part of Nyquist diagram appears, R(Q(R(QR)))(W) models are used to fit the results, and the equivalent electrical circuit diagram is shown in Figure 5. The fitting results are shown in Table 3, where Rs stands for solution resistance, Rb stands for film resistance, Rt stands for charge transfer resistance, and Ws represents the Warburg diffusion impedance, which is a very slow process. Before it affects the corrosion of the alloy, the alloy undergoes severe corrosion due to other reasons, and the corrosion resistance of the alloy is generally not determined by Ws. Therefore, in AC impedance testing, the Ws results are generally not discussed, constant phase element CPE stands for non-ideal capacitance caused by the non-uniform electrode, and its impedance value is given by the following formula:
Z = 1 T j ω n
T is the scale factor, j is the imaginary number unit, ω is the angular frequency, and n is the phase shift, which is between 0 and 1. When n = 0, CPE behaves as a pure resistance. When n = 1, CPE is equivalent to a pure capacitor [34]. Generally, when the chi-square value is between 10−3 and 10−4, it indicates that the fitting results are reliable. In this paper, the chi-square values of all the results are between 10−3 and 10−4. Rt value is positively correlated with corrosion resistance. The larger the Rt value, the more difficult the charge transfer and the better the corrosion resistance. It can be seen from Table 3 that the Rt value of CoCrFeMnNi with 2.5 wt.% WC sample is 859.1 Ω·cm2, about six-fold that of CoCrFeMnNi sample.
Figure 5. AC impedance spectrum fitting circuit diagram of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC.
Figure 5. AC impedance spectrum fitting circuit diagram of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC.
Materials 16 04701 g005
Table 3. Equivalent circuit fitting parameters of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC in 0.5 M H2SO4 solution.
Table 3. Equivalent circuit fitting parameters of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC in 0.5 M H2SO4 solution.
CoCrFeMnNiCoCrFeMnNi with 2.5 wt.% WC
Rs (Ω·cm2)1.1430.460
Rb (Ω·cm2)0.4110.806
Rt (Ω·cm2)138.8859.1
T1 (F·cm−2)4.1042 × 10−62.179 × 10−6
n10.93040.986
T2 (F·cm−2)4.442 × 10−54.081 × 10−5
n20.87230.928
Chi-square0.829 × 10−37.199 × 10−3

3.3. Microstructure after Corrosion

The morphology of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC after the potentiodynamic polarization curve test in 0.5 M H2SO4 solution is shown in Figure 6. As can be seen in Figure 6(a1), strong corrosion occurred on the surface of the CoCrFeMnNi sample, and the dendrite structure was clearly visible. This was mainly due to the segregation of Mn and Fe elements during slow solidification, which resulted in the difference in composition between the first-solidification and post-solidification regions, and corrosion galvanic cells were formed under applied voltage. However, from the SEM images of Figure 6(a2), only a slight surface relief and partial corrosion pits were observed, indicating a large area of uniform corrosion. On the other hand, CoCrFeMnNi with 2.5 wt.% WC sample underwent relatively large corrosion along the grain boundary, which is because the C atom generated during the decomposition of WC tended to produce nanometer carbide with Cr element at the grain boundary [19,35], resulting in the formation of Cr-poor zone at the grain boundary, resulting in relatively severe corrosion at the grain boundary [36], as can be seen in Figure 6(b1).
To elucidate the improvement of corrosion resistance of CoCrFeMnNi with 2.5 wt.% WC alloy, XPS was carried out to study the composition and valence state of the passivation film formed after the corrosion. The high-resolution spectra of O 1 s, Fe 2p3/2, Cr 2p3/2, Ni 2p3/2, Co 2p3/2, Mn 2p3/2, and W 4f are shown in Figure 7a–f and Figure 8a–g, and the element distribution in the passivated film is shown in Figure 7g and Figure 8h. The spectrum of Co 2p3/2 in the passivated film is composed of Co0 and Coox3+/2+; Coox3+/2+ is related to its oxides CoO and Co3O4. Although CoO is easily dissolved in acidic solution, it may be formed in air, and oxide can also be formed in the anode polarization process [37,38]. The spectrum of Cr 2p3/2 consists of Cr0, Crox3+ and Crhy3+, among which Crox3+ is related to Cr2O3, FeCr2O4 or NiCr2O4, while Crhy3+ is related to Cr(OH)3 [39,40,41], in which Crox3+ accounts for 53.08%. The proportion of Crhy3+ is 33.47%, indicating that the main valence state of Cr element in the passivation film is Crox3+. While Cr2O3 and Cr(OH)3 are considered to be the key point to the quality of the passivation film [42], Cr2O3 and Cr(OH)3 accounting for 86.55% in total, and Cr element accounts for 21.87% in the whole passivation film, which is much bigger than other elements. The spectrum of Fe 2p3/2 is complicated, and there are many possible substances with overlapping binding energy, which is very difficult to distinguish. As shown in Figure 7c and Figure 8c, the spectrum is divided into constituent peaks representing Fe, Feox2+/3+, Feox3+, and Fehy3+. After anode polarization in 0.5 M H2SO4 solution, FeO is difficult to exist [43] and is not easy to form in dry air; therefore, the compounds related to Feox2+/3+ may be Fe3O4 and FeCr2O4. Feox3+ is related to Fe2O3, and due to the spectral overlap of Fe2O3 and its complex composition, NiFe2O4 may be related with Feox3+. Fehy3+ comes from Fe(OH)3 or FeOOH [37,38,39,40]. From the figure, the relative intensities and the peak areas of Feox2+/3+, Feox3+, and Fehy3+ are similar. The Mn 2p3/2 spectrum consists of Mn0, Mn2+, Mn3+ and Mn4+. Mn2+ is connected with MnO; Mn3+ may be Mn2O3 or MnOOH, and Mn4+ is related to MnO2. The spectrum of Ni 2p3/2 consists of Ni0 and Niox2+. Niox2+ is related to NiO, NiFe2O4, or NiCr2O4 [37,44]. Comparing the peak intensity, it is found that the valence state of Ni is mainly Ni0. As shown in Figure 8f, unlike the CrMnFeCoNi alloy, O1s spectrum of the CoCrFeMnNi with 2.5 wt.% WC alloy is composed of O2−, OH, and H2O, which correspond to metal oxides and hydroxides in the passivation film. H2O may be the binding water formed in the passivation film [45,46]; Luo et al. reported the same results—that the binding water can be an effective substance to capture dissolved metal ions, and a new film forms to resist further corrosion [47]. It can be found, by analyzing the peak intensity, that O in the passivated film comes from a large number of metal hydroxides, which corresponds to the spectrum of other elements. The spectra of W 4f are mainly W0 and W6+, and W6+ is related to WO3 [48,49].
In summary, the valence states of each element in CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC are similar, except for oxygen element. Only CoCrFeMnNi with 2.5 wt.% WC alloy is composed of H2O. The corrosion resistance of the metal is highly dependent on the composition and structure of the passivation film formed in the solution, among which Cr is considered to be the main reason for the corrosion resistance of stainless steel. As can be seen from the element distribution in Figure 8h, the passivation film is mainly composed of the oxides/hydroxides of Cr(Cr2O3, Cr(OH)3), whose content reaches 21.87% equivalent to 304 L stainless steel. The content of Mn in the passivated film hardly decreases compared with the nominal composition of the alloy, which is consistent with the literature [50]. Carbon has an impact on the formation of bound water in the passivation film. If there is bound water in the passivation film, it will have a significant impact on the stability of the passivation film. The existence of bound water in the passivation film has a strong self-healing ability on the passivation film, and the bound water in the film will capture dissolved metal ions, and the new film will form to prevent further corrosion [46,47]. Adding a small amount of carbon to FeCoCrNiMn will increase the content of bound water in the passivation film, so that the corrosion resistance of the passivation film is improved. In the passivation film formed by the CoCrFeMnNi with 2.5 wt.% WC in 0.5 M H2SO4 solution, the content of bound water is 4.49%, which will improve the corrosion resistance of the passivation film. In addition, the addition of W can inhibit the dissolution of metal in acidic electrolyte and also improve the corrosion resistance of the alloy.

4. Conclusions

In this study, LMDed CoCrFeMnNi with 2.5 wt.% WC HEA particle were fabricated, and a comparative study on the electrochemical corrosion behavior of the CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC particle in 0.5 M H2SO4 was carried out. The influence of WC on the corrosion resistance of CoCrFeMnNi was investigated, and the conclusions are as follows:
(1)
The microstructure of CoCrFeMnNi with WC particle prepared by laser melting deposition is composed of columnar crystals and equiaxed crystals. During the preparation process, WC particles were decomposed, and elements C and W were incorporated into the CoCrFeMnNi matrix, resulting in strong lattice distortion;
(2)
The electrochemical measurement results show that CoCrFeMnNi with 2.5 wt.% WC have a smaller corrosion current density of 1.594 × 10−5 A·cm−2 and larger corrosion potential −0.285 VAg/AgCl and higher charge transfer 859.1 Ω·cm2, showing better corrosion resistance than CoCrFeMnNi;
(3)
The morphology after corrosion shows that the CoCrFeMnNi has a large area of uniform corrosion, while the CoCrFeMnNi with 2.5 wt.% WC corrodes along the grain boundary; furthermore, the XPS results of the passive film show that the content of Cr2O3 and Cr(OH)3 are high, which is helpful to improve the stability of the passive film, and additionally, that the decomposition of WC is not a bad thing. The incorporation of C atoms causes the combined water to appear in the passive film, which makes the passive film have a self-repairing function and improves its corrosion resistance. In addition, the addition of W can inhibit the dissolution of metal in acidic electrolyte and also improve the corrosion resistance of the alloy.

Author Contributions

Z.P. concepted the work, supervised the research progress, and wrote the manuscript. Z.F. conducted the sample preparation and performed the microstructural analysis. M.R.A. and C.R. contributed to the corrosion test and performed the electrochemical analysis. J.L. and P.G. revised the manuscript and contributed to the discussion. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (No. 52271147, and No. 12261160364) and Guangdong Basic and Applied Basic Research Foundation (Grant number 2023A1515012158).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this work are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yeh, J.W.; Chen, S.K.; Lin, S.J.; Gan, J.Y.; Chin, T.S.; Shun, T.T.; Tsau, C.H.; Chang, S.Y. Nanostructured high-entropy alloys with multiple principal elements: Novel alloy design concepts and outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  2. Cantor, B.; Chang, I.T.H.; Knight, P.C.; Vincent, A. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375, 213–218. [Google Scholar] [CrossRef]
  3. Miracle, D.B.; Senkov, O.N. A critical review of high entropy alloys and related concepts. Acta Mater. 2017, 122, 448–511. [Google Scholar] [CrossRef] [Green Version]
  4. Otto, F.; Dlouhý, A.; Somsen, C.; Bei, H.; Eggeler, G.; George, E.P. The influences of temperature and microstructure on the tensile properties of a CoCrFeMnNi high-entropy alloy. Acta Mater. 2013, 61, 5743–5755. [Google Scholar] [CrossRef] [Green Version]
  5. Luan, H.; Zhang, X.; Ding, H.; Zhang, F.; Luan, J.; Jiao, Z.; Yang, Y.C.; Bu, H.; Wang, R.; Gu, J.; et al. High-entropy induced a glass-to-glass transition in a metallic glass. Nat. Commun. 2022, 13, 2183. [Google Scholar] [CrossRef]
  6. Peng, Z.; Sun, J.; Luan, H.W.; Chen, N.; Yao, K.F. Effect of Mo on the high temperature oxidation behavior of Al19Fe20-xCo20-xNi41Mo2x high entropy alloys. Intermetallics 2023, 155, 107845. [Google Scholar] [CrossRef]
  7. Lei, Z.F.; Liu, X.J.; Wu, Y.; Wang, H.; Jiang, S.; Wang, S.D.; Hui, X.D.; Wu, Y.D.; Gault, B.; Kontis, P.; et al. Enhanced strength and ductility in a high-entropy alloy via ordered oxygen complexes. Nature 2018, 563, 546–550. [Google Scholar] [CrossRef]
  8. Lu, Z.P.; Wang, H.; Chen, M.W.; Baker, I.; Yeh, J.W.; Liu, C.T.; Nieh, T.G. An assessment on the future development of high-entropy alloys: Summary from a recent workshop. Intermetallics 2015, 66, 67–76. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, Y.; Zuo, T.T.; Tang, Z.; Gao, M.C.; Dahmen, K.A.; Liaw, P.K.; Lu, Z.P. Microstructures and properties of high-entropy alloys. Prog. Mater Sci. 2014, 61, 1–93. [Google Scholar] [CrossRef]
  10. Peng, Z.; Li, B.W.; Luo, Z.B.; Chen, X.F.; Tang, Y.; Yang, G.N.; Gong, P. A Lightweight AlCrTiV0.5Cux High-Entropy Alloy with Excellent Corrosion Resistance. Materials 2023, 16, 2922. [Google Scholar] [CrossRef] [PubMed]
  11. Chen, X.F.; Wang, Q.; Cheng, Z.Y.; Zhu, M.L.; Zhou, H.; Jiang, P.; Zhou, L.L.; Xue, Q.; Yuan, F.P.; Zhu, J.; et al. Direct observation of chemical short-range order in a medium-entropy alloy. Nature 2021, 592, 712–716. [Google Scholar] [CrossRef]
  12. George, E.P.; Raabe, D.; Ritchie, R.O. High-entropy alloys. Nat. Rev. Mater. 2019, 4, 515–534. [Google Scholar] [CrossRef]
  13. Peng, Z.; Luo, Z.B.; Li, B.W.; Li, J.F.; Luan, H.W.; Gu, J.L.; Wu, Y.; Yao, K.F. Microstructure and mechanical properties of lightweight AlCrTiV0.5Cux high-entropy alloys. Rare Met. 2022, 41, 2016–2020. [Google Scholar] [CrossRef]
  14. Jia, Y.F.; Jia, Y.D.; Wu, S.W.; Ma, X.D.; Wang, G. Novel Ultralight-Weight Complex Concentrated Alloys with High Strength. Materials 2019, 12, 1136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Senkov, O.N.; Senková, S.; Miracle, D.B.; Woodward, C. Mechanical properties of low-density, refractory multi-principal element alloys of the Cr–Nb–Ti–V–Zr system. Mater. Sci. Eng. A 2013, 565, 51–62. [Google Scholar] [CrossRef]
  16. Ren, J.; Zhang, Y.; Zhao, D.X.; Chen, Y.; Guan, S.; Liu, Y.F.; Liu, L.; Peng, S.Y.; Kong, F.; Poplawsky, J.D.; et al. Strong yet ductile nanolamellar high-entropy alloys by additive manufacturing. Nature 2022, 608, 62–68. [Google Scholar] [CrossRef] [PubMed]
  17. Xiang, S.; Luan, H.W.; Wu, J.; Yao, K.F.; Li, J.F.; Liu, X.; Tian, Y.Z.; Mao, W.L.; Bai, H.; Le, G.M.; et al. Microstructures and mechanical properties of CrMnFeCoNi high entropy alloys fabricated using laser metal deposition technique. J. Alloys Compd. 2019, 773, 387–392. [Google Scholar] [CrossRef]
  18. Xiang, S.; Li, J.; Luan, H.W.; Amar, A.; Lu, S.Y.; Li, K.; Zhang, L.; Liu, X.; Le, G.M.; Wang, X.Y.; et al. Effects of process parameters on microstructures and tensile properties of laser melting deposited CrMnFeCoNi high entropy alloys. Mater. Sci. Eng. A 2019, 743, 412–417. [Google Scholar] [CrossRef]
  19. Li, J.F.; Xiang, S.; Luan, H.W.; Amar, A.; Liu, X.; Lu, S.Y.; Zeng, Y.Y.; Le, G.M.; Wang, X.Y.; Qu, F.S.; et al. Additive manufacturing of high-strength CrMnFeCoNi high-entropy alloys-based composites with WC addition. J. Mater. Sci. Technol. 2019, 35, 2430–2434. [Google Scholar] [CrossRef]
  20. Chew, Y.; Bi, G.J.; Zhu, Z.G.; Ng, F.L.; Weng, F.; Liu, S.B.; Nai, S.M.L.; Lee, B.Y. Microstructure and enhanced strength of laser aided additive manufactured CoCrFeNiMn high entropy alloy. Mater. Sci. Eng. A 2019, 744, 137–144. [Google Scholar] [CrossRef]
  21. Wang, Y.M.; Voisin, T.; McKeown, J.T.; Ye, J.; Calta, N.P.; Li, Z.; Zeng, Z.; Zhang, Y.; Chen, W.; Roehling, T.T.; et al. Additively manufactured hierarchical stainless steels with high strength and ductility. Nat. Mater. 2018, 17, 63–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Melia, M.A.; Carroll, J.D.; Whetten, S.R.; Esmaeely, S.N.; Locke, J.; White, E.; Anderson, I.; Chandross, M.; Michael, J.R.; Argibay, N.; et al. Mechanical and Corrosion Properties of Additively Manufactured CoCrFeMnNi High Entropy Alloy. Addit. Manuf. 2019, 29, 100833. [Google Scholar] [CrossRef]
  23. Jia, Y.J.; Chen, H.N.; Liang, X.D. Microstructure and wear resistance of CoCrNbNiW high-entropy alloy coating prepared by laser melting deposition. Rare Met. 2019, 38, 1153–1159. [Google Scholar] [CrossRef]
  24. Moghaddam, A.O.; Shaburova, N.A.; Samodurova, M.N.; Abdollahzadeh, A.; Trofimov, E.A. Additive manufacturing of high entropy alloys: A practical review. J. Mater. Sci. Technol. 2021, 77, 131–162. [Google Scholar] [CrossRef]
  25. Sun, Z.; Tan, X.P.; Descoins, M.; Mangelinck, D.; Tor, S.B.; Lim, C.S. Revealing hot tearing mechanism for an additively manufactured high-entropy alloy via selective laser melting. Scr. Mater. 2019, 168, 129–133. [Google Scholar] [CrossRef]
  26. Guo, J.; Liu, C.H.; Wang, D.X.; Xu, L.F.; Song, K.K.; Gao, M. Structure and Wear Resistance of TiC-Reinforced Al1.8CrCuFeNi2 High-Entropy Alloy Coating Using Laser Cladding. Materials 2023, 16, 3422. [Google Scholar] [CrossRef]
  27. Chen, P.; Li, S.; Zhou, Y.H.; Yan, M.; Attallah, M.M. Fabricating CoCrFeMnNi high entropy alloy via selective laser melting in-situ alloying. J. Mater. Sci. Technol. 2020, 43, 40–43. [Google Scholar] [CrossRef]
  28. Kim, Y.K.; Cho, J.; Lee, K.A. Selective laser melted equiatomic CoCrFeMnNi high-entropy alloy: Microstructure, anisotropic mechanical response, and multiple strengthening mechanism. J. Alloys Compd. 2019, 805, 680–691. [Google Scholar] [CrossRef]
  29. Zhang, Z.J.; Mao, M.M.; Wang, J.G.; Gludovatz, B.; Zhang, Z.; Mao, S.X.; George, E.P.; Yu, Q.; Ritchie, R.O. Nanoscale origins of the damage tolerance of the high-entropy alloy CrMnFeCoNi. Nat. Commun. 2015, 6, 10143. [Google Scholar] [CrossRef] [Green Version]
  30. Xu, Z.L.; Zhang, H.; Du, X.J.; He, Y.Z.; Luo, H.; Song, G.S.; Mao, L.; Zhou, T.W.; Wang, L.L. Corrosion resistance enhancement of CoCrFeMnNi high-entropy alloy fabricated by additive manufacturing. Corros. Sci. 2020, 177, 108954. [Google Scholar] [CrossRef]
  31. Yehia, H.M. Microstructure, physical, and mechanical properties of the Cu/ (WC-TiC-Co) nano-composites by the electroless coating and powder metallurgy technique. J. Compos. Mater. 2019, 53, 1963–1971. [Google Scholar] [CrossRef]
  32. Hassan, M.A.; Yehia, H.M.; Mohamed, A.S.A.; El-Nikhaily, A.E.; Elkady, O.A. Effect of Copper Addition on the AlCoCrFeNi High Entropy Alloys Properties via the Electroless Plating and Powder Metallurgy Technique. Crystals 2021, 11, 540. [Google Scholar] [CrossRef]
  33. Salishchev, G.; Tikhonovsky, M.A.; Shaysultanov, D.; Stepanov, N.D.; Kuznetsov, A.V.; Kolodiy, I.V.; Tortika, A.S.; Senkov, O.N. Effect of Mn and V on structure and mechanical properties of high-entropy alloys based on CoCrFeNi system. J. Alloys Compd. 2014, 591, 11–21. [Google Scholar] [CrossRef]
  34. Kissi, M.; Bouklah, M.; Hammouti, B.; Benkaddour, M. Establishment of equivalent circuits from electrochemical impedance spectroscopy study of corrosion inhibition of steel by pyrazine in sulphuric acidic solution. Appl. Surf. Sci. 2006, 252, 4190–4197. [Google Scholar] [CrossRef]
  35. Park, J.M.; Choe, J.; Kim, J.G.; Bae, J.W.; Moon, J.; Yang, S.; Kim, K.T.; Yu, J.H.; Kim, H.S. Superior tensile properties of 1%C-CoCrFeMnNi high-entropy alloy additively manufactured by selective laser melting. Mater. Res. Lett. 2019, 8, 1–7. [Google Scholar] [CrossRef] [Green Version]
  36. Samoilova, O.; Pratskova, S.; Shaburova, N.; Ostovari Moghaddam, A.; Trofimov, E. Corrosion Resistance of Al0.5CoCrFeNiCuxAgy (x = 0.25, 0.5; y = 0, 0.1) High-Entropy Alloys in 0.5M H2SO4 Solution. Materials 2023, 16, 3585. [Google Scholar] [CrossRef] [PubMed]
  37. Luo, H.; Li, Z.M.; Mingers, A.M.; Raabe, D. Corrosion behavior of an equiatomic CoCrFeMnNi high-entropy alloy compared with 304 stainless steel in sulfuric acid solution. Corros. Sci. 2018, 134, 131–139. [Google Scholar] [CrossRef]
  38. Gardin, E.; Zanna, S.; Seyeux, A.; Allion-Maurer, A.; Marcus, P. XPS and ToF-SIMS characterization of the surface oxides on lean duplex stainless steel – Global and local approaches. Corros. Sci. 2019, 155, 121–133. [Google Scholar] [CrossRef]
  39. Luo, H.; Zou, S.W.; Chen, Y.H.; Li, Z.; Du, C.W.; Li, X.G. Influence of carbon on the corrosion behaviour of interstitial equiatomic CoCrFeMnNi high-entropy alloys in a chlorinated concrete solution. Corros. Sci. 2020, 163, 108287. [Google Scholar] [CrossRef]
  40. Sun, Y.P.; Wang, Z.; Yang, H.J.; Lan, A.D.; Qiao, J.W. Effects of the element La on the corrosion properties of CrMnFeNi high entropy alloys. J. Alloys Compd. 2020, 842, 155825. [Google Scholar] [CrossRef]
  41. Pourbaix, M. Atlas of Electrochemical Equilibria in Aqueous Solutions; National Association of Corrosion: Houston, TX, USA, 1974. [Google Scholar]
  42. Haupt, S.; Strehblow, H.H. Combined Surface Analytical and Electrochemical Study of the Formation of Passive Layers on Fe/Cr Alloys in 0.5 M H2SO4. ChemInform 1995, 26, 43–54. [Google Scholar] [CrossRef]
  43. Mansour, C.; Lefèvre, G.; Pavageau, E.M.; Catalette, H.; Fédoroff, M.; Zanna, S. Sorption of sulfate ions onto magnetite. J. Colloid Interface Sci. 2009, 331, 77–82. [Google Scholar] [CrossRef] [PubMed]
  44. Okamoto, G.O.; Shibata, T. Desorption of Tritiated Bound-water from the Passive Film Formed on Stainless Steels. Nature 1965, 206, 1350. [Google Scholar] [CrossRef]
  45. Jiang, Y.; Yang, J.F.; Liu, R.; Wang, X.P.; Fang, Q.F. Oxidation and corrosion resistance of WC coated tungsten fabricated by SPS carburization. J. Nucl. Mater. 2014, 450, 75–80. [Google Scholar] [CrossRef]
  46. Noji, N.; Kashiwagura, K.; Akao, N.; Soma, S.; Hara, N.; Sugimoto, K. Corrosion resistance of tungsten and tungsten alloys for spallation target in stagnant and flowing water. J. Jpn. Inst. Met. 2002, 66, 1107–1115. [Google Scholar] [CrossRef] [Green Version]
  47. Hsu, K.M.; Chen, S.H.; Lin, C.S. Microstructure and corrosion behavior of FeCrNiCoMnx (x = 1.0, 0.6, 0.3, 0) high entropy alloys in 0.5 M H2SO4. Corros. Sci. 2021, 190, 109694. [Google Scholar] [CrossRef]
  48. Wang, C.M.; Yu, Y.; Zhang, H.; Xu, L.X.; Ma, X.Y.; Wang, F.F.; Song, B.Y. Microstructure and corrosion properties of laser remelted CrFeCoNi and CrMnFeCoNi high entropy alloys coatings. J. Mater. Res. Technol. 2021, 15, 5187–5196. [Google Scholar] [CrossRef]
  49. Zhao, Q.H.; Liu, W.; Li, S.Z.; Zhang, B.L.; Zhu, Y.C.; Lu, M.X. Effects of W and Mo Additions on Wet-dry Acid Corrosion Behavior of Low-Alloy Steels Under Different O2 Concentrations. Acta Metall. Sin. (Engl. Lett.) 2016, 29, 951–962. [Google Scholar] [CrossRef] [Green Version]
  50. Ma, M.Y.; He, C.L.; Chen, L.Q.; Wei, L.L.; Misra, R.D.K. Effect of W and Ce additions on the electrochemical corrosion behaviour of 444-type ferritic stainless steel. Corros. Eng. Sci. Technol. 2018, 53, 199–205. [Google Scholar] [CrossRef]
Figure 1. (a) X-ray diffraction patterns of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC samples, (b) partial enlargement of the XRD pattern in (a).
Figure 1. (a) X-ray diffraction patterns of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC samples, (b) partial enlargement of the XRD pattern in (a).
Materials 16 04701 g001
Figure 2. Microstructural characterization of samples (a1) SEM of CoCrFeMnNi, (a2) partial enlargement view of a1, (b1) SEM of CoCrFeMnNi with 2.5 wt.% WC, (b2) partial enlargement view of b1 (c) EDS mapping of CoCrFeMnNi, (d) EDS mapping of CoCrFeMnNi with 2.5 wt.% WC.
Figure 2. Microstructural characterization of samples (a1) SEM of CoCrFeMnNi, (a2) partial enlargement view of a1, (b1) SEM of CoCrFeMnNi with 2.5 wt.% WC, (b2) partial enlargement view of b1 (c) EDS mapping of CoCrFeMnNi, (d) EDS mapping of CoCrFeMnNi with 2.5 wt.% WC.
Materials 16 04701 g002
Figure 3. Characterization of precipitate in CoCrFeMnNi with 2.5 wt.% WC sample. (a) TEM image; (b) HAADF-STEM image and corresponding element maps.
Figure 3. Characterization of precipitate in CoCrFeMnNi with 2.5 wt.% WC sample. (a) TEM image; (b) HAADF-STEM image and corresponding element maps.
Materials 16 04701 g003
Figure 4. (a) Potentiodynamic polarization curves and electrochemical impedance spectroscopy of (b) CoCrFeMnNi and (c) CoCrFeMnNi with 2.5 wt.% WC in 0.5 M H2SO4 solution.
Figure 4. (a) Potentiodynamic polarization curves and electrochemical impedance spectroscopy of (b) CoCrFeMnNi and (c) CoCrFeMnNi with 2.5 wt.% WC in 0.5 M H2SO4 solution.
Materials 16 04701 g004
Figure 6. Surface morphology after dynamic potential polarization curve test in 0.5 M H2SO4 solution. (a1) CoCrFeMnNi; (a2) partial enlargement view of a1, (b1) CoCrFeMnNi with 2.5 wt.% WC, (b2) partial enlargement view of b1.
Figure 6. Surface morphology after dynamic potential polarization curve test in 0.5 M H2SO4 solution. (a1) CoCrFeMnNi; (a2) partial enlargement view of a1, (b1) CoCrFeMnNi with 2.5 wt.% WC, (b2) partial enlargement view of b1.
Materials 16 04701 g006
Figure 7. High resolution XPS spectra of the passive films formed on the CoCrFeMnNi after dynamic potential polarization curve test in 0.5 M H2SO4 solution (a) Co, (b) Cr, (c) Fe, (d) Mn, (e) Ni, (f) O, (g) elemental fractions in the passive film of HEA obtained by XPS analysis.
Figure 7. High resolution XPS spectra of the passive films formed on the CoCrFeMnNi after dynamic potential polarization curve test in 0.5 M H2SO4 solution (a) Co, (b) Cr, (c) Fe, (d) Mn, (e) Ni, (f) O, (g) elemental fractions in the passive film of HEA obtained by XPS analysis.
Materials 16 04701 g007
Figure 8. High resolution XPS spectra of the passive films formed on the CoCrFeMnNi with 2.5 wt.% WC after dynamic potential polarization curve test in 0.5 M H2SO4 solution (a) Co, (b) Cr, (c) Fe, (d) Mn, (e) Ni, (f) O, (g) W; (h) elemental fractions in the passive film of HEA obtained by XPS analysis.
Figure 8. High resolution XPS spectra of the passive films formed on the CoCrFeMnNi with 2.5 wt.% WC after dynamic potential polarization curve test in 0.5 M H2SO4 solution (a) Co, (b) Cr, (c) Fe, (d) Mn, (e) Ni, (f) O, (g) W; (h) elemental fractions in the passive film of HEA obtained by XPS analysis.
Materials 16 04701 g008
Table 1. Composition of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC high-entropy alloys A and B regions in Figure 2(a2,b2).
Table 1. Composition of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC high-entropy alloys A and B regions in Figure 2(a2,b2).
CrMnFeCoNiCW
LMDed HEAA20.4023.4520.6418.2617.25
B23.8719.9420.5518.3717.27
LMDed HEA with 2.5 wt.% WCA21.5921.5519.0718.0018.940.550.30
B26.6819.0715.0016.2015.656.870.53
Table 2. Electrochemical corrosion parameters of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC obtained from potentiodynamic polarization curve measured in 0.5 M H2SO4 solution.
Table 2. Electrochemical corrosion parameters of CoCrFeMnNi and CoCrFeMnNi with 2.5 wt.% WC obtained from potentiodynamic polarization curve measured in 0.5 M H2SO4 solution.
SampleCoCrFeMnNiCoCrFeMnNi with 2.5 wt.% WC
Ecorr(VAg/AgCl)−0.298−0.285
icorr(A/cm−2)3.038 × 10−51.594 × 10−5
Epp(VAg/AgCl)−0.14−0.107
ipass(A/cm2)6.275 × 10−52.519 × 10−5
Eb(VAg/AgCl)0.9610.926
Esp(VAg/AgCl)1.1961.147
ΔE(VAg/AgCl)1.1011.033
Epp: primary passivation potential ipass: Passivation current density Eb: breakdown potential. Esp: secondary passivation potential ΔE: Eb − Epp, length of passive zone.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Peng, Z.; Fan, Z.; Abdullah, M.R.; Ren, C.; Li, J.; Gong, P. Corrosion Resistance Enhancement of CoCrFeMnNi High-Entropy Alloy with WC Particle Reinforcements via Laser Melting Deposition. Materials 2023, 16, 4701. https://doi.org/10.3390/ma16134701

AMA Style

Peng Z, Fan Z, Abdullah MR, Ren C, Li J, Gong P. Corrosion Resistance Enhancement of CoCrFeMnNi High-Entropy Alloy with WC Particle Reinforcements via Laser Melting Deposition. Materials. 2023; 16(13):4701. https://doi.org/10.3390/ma16134701

Chicago/Turabian Style

Peng, Zhen, Zize Fan, Muhammad Raies Abdullah, Congcong Ren, Jinfeng Li, and Pan Gong. 2023. "Corrosion Resistance Enhancement of CoCrFeMnNi High-Entropy Alloy with WC Particle Reinforcements via Laser Melting Deposition" Materials 16, no. 13: 4701. https://doi.org/10.3390/ma16134701

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