Next Article in Journal
Studies on the Coating Formation and Structure Property for Plasma Electrolytic Oxidation of AZ31 Magnesium Alloy
Previous Article in Journal
Modified Asphalt Prepared by Coating Rubber Powder with Waste Cooking Oil: Performance Evaluation and Mechanism Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research on the Synergistic Evolution Law of Microstructure and Properties of Deformed Austenitic Stainless Steel

1
Nanxun Innovation Institute, Zhejiang University of Water Resources and Electric Power, Hangzhou 310018, China
2
College of Mechanical and Electrical Engineering, China Jiliang University, Hangzhou 310018, China
*
Author to whom correspondence should be addressed.
Coatings 2025, 15(7), 845; https://doi.org/10.3390/coatings15070845
Submission received: 21 June 2025 / Revised: 10 July 2025 / Accepted: 17 July 2025 / Published: 18 July 2025
(This article belongs to the Section Surface Characterization, Deposition and Modification)

Abstract

Austenitic stainless steel inevitably undergoes deformation during application, and it is necessary to study the properties of deformed steel. This article investigates the evolution of microstructure, mechanical properties, and corrosion resistance of plastic-deformed 304 steel, the evolution law of structure and properties of steel is revealed. As a result, it was found that with the increase in deformation, the grains of 304 steel were destroyed, and many small subgrains were generated internally, resulting in a significant decrease in grain size. At the same time, the content of martensitic transformation in stainless steel increased significantly. The characteristics of the surface passivation film of stainless steel also change during the deformation process. Meanwhile, with the increase in deformation, the nanohardness and wear resistance of 304 steel gradually increase, but its corrosion resistance gradually decreases. Analysis suggests that microstructural changes such as grain size and phase transformation in stainless steel lead to an improvement in its mechanical properties, while the generation of defects during deformation and changes in surface passivation film characteristics result in a deterioration of its corrosion resistance. This study can provide a reference for the forming and performance optimization of metals and has high theoretical significance and practical value.

Graphical Abstract

1. Introduction

Stainless steel materials exhibit high strength, corrosion resistance, easy cleaning, and excellent comprehensive properties, making them widely used in numerous fields such as petrochemicals, pharmaceuticals, food processing, and transportation [1,2]. With the improvement of living standards, stainless steel products in production and daily life are gradually shifting from “high-speed” to “high-quality” development. Current stainless steel materials can no longer meet the increasingly demanding performance requirements, making more in-depth research on stainless steel materials extremely necessary. Stainless steel materials inevitably undergo deformation during the manufacturing of products. Different deformation conditions alter the material’s microstructure, thereby affecting its macroscopic properties [3]. Materials undergo varying degrees of deformation during product forming; therefore, revealing the evolution of microstructure and properties of stainless steel materials with deformation level is essential.
Currently, researchers have conducted some studies on the effects of deformation on stainless steel materials. Research has found that 304 L metastable austenitic stainless steel exhibits a linear temperature-dependent relationship between tensile strength and elongation during plastic deformation at temperatures above Md (the highest temperature at which martensitic transformation can be induced during plastic deformation); below Md, these parameters increase sharply. When the number of martensites in the necking area becomes significant, a noticeable phase transformation-induced plasticity effect is observed, leading to an enhancement of the maximum point and strength ductility balance in the work hardening diagram [4]. Chen et al. [5] uses molecular dynamics simulations to investigate the effect of deformation on the microstructure evolution and mechanical properties of bismuth containing austenitic stainless steel and further explores its impact on subsequent tensile behavior. Research has shown that with the increase in drawing deformation, the internal grains of the material elongate and refine, and the dislocation density significantly increases. When the degree of deformation is large, bismuth nanoparticles can not only effectively hinder the movement of dislocations but also promote the aggregation and entanglement of dislocations. As a result, the local strength of the material is enhanced, leading to work hardening and an increase in axial tensile force. In addition, studies have shown that when strain-induced α′-martensite is present in the material, austenitic stainless steel undergoes strong strain hardening, and the stress–strain behavior is related to the increase in yield strength and tensile strength with decreasing temperature. Analysis shows that the kinetics of metastable austenitic stainless steel is influenced by temperature and the content of α’-martensite [6]. Furthermore, the differential strain-induced martensite in austenitic stainless steel can also affect the material properties. Chen et al. [7] found that as the grain size of austenite increases from the ultrafine grain range to the coarse grain range, the nucleation site of α’-martensite will shift from the grain boundary to the interior of deformed grains, and the nucleation of ε-martensite will also occur. The nucleation sites in austenite are due to the morphology of α’-martensite, which increases from the ultrafine grain range to the coarse grain range and transforms from blocky to thin film-like. Compared with the blocky α‘-martensite formed at grain boundaries, the thin film-like α’-martensite formed at the intersection of shear bands will result in a higher work hardening rate in stainless steel. It can be seen that deformation has a significant impact on the microstructure of stainless steel, which in turn affects its properties.
This paper takes the commonly used 304 stainless steel in production and daily life as the research object, studying the evolution of microstructure and macroscopic properties of stainless steel under different deformation levels during tensile deformation. Simultaneously, it analyzes the influence of stainless steel microstructure on mechanical properties and corrosion resistance, as well as the correlation between microstructure and properties, revealing the evolution mechanism of microstructure and properties in differently deformed stainless steel materials, providing a theoretical reference for the performance optimization of stainless steel materials. This research can provide a reference value for the forming and performance optimization industry of metallic materials, possessing high practical application significance and value.

2. Materials and Methods

2.1. Material Composition and Plastic Deformation Process

The 304 stainless steel used in the experiments was the material used for company production. The material composition is shown in Table 1.
First, the 304 stainless steel was subjected to tensile tests using an INSTRON universal testing machine equipped with an extensometer. Slow strain rate tensile specimens were prepared using the stainless steel material, with dimensions as shown in Figure 1. The material was deformed at the same tensile rate of 10−6 s−1, with tensile deformation levels of 5%, 10%, 20%, and 40%, respectively. The tensile test was repeated three times. Figure 2 shows the tensile curve of 304 stainless steel, and the photographs of the specimen after tensile testing was included in the illustration in Figure 2. After the tensile test, the structure and properties of stainless steel with different deformation degrees were observed.
From the photographs of the specimen after tensile testing, the stainless steel broke at the middle position, and the stress–strain curve displayed that 304 stainless steel exhibits typical elastic–plastic deformation, with a yield strength of approximately 307 MPa, tensile strength of approximately 797 MPa, and elongation of approximately 81%.

2.2. Microstructure Characterization and Macroscopic Performance Testing

The microstructure, element content, sample surface, and oxide film of 304 steel were analyzed and studied employing an optical electron microscope (OM) (LEICA, Wetzlar, Germany), scanning electron microscopy (SEM) (ZEISS, Baden-Württemberg, Germany), electron backscatter diffraction (EBSD) (ZEISS, Baden-Württemberg, Germany), X-ray photoelectron spectroscopy (XPS) (ThermoFisher, Waltham, MA, USA), and ferrite equivalent meter (Fischer, Baden-Württemberg, Germany). The acceleration voltage for EBSD testing is 25 kV, the scanning speed is 635 Hz, and each step length is 0.5 μM. The electron energy of the XPS test electron source is 2000 eV, with an energy half width of 0.4 eV. The minimum beam spot diameter is about 1 mm, and the maximum output current is about 65 μA.
Mechanical property tests were conducted on the stretched material, including a nanoindentation test and friction-wear test. The nanoindentation experiment was conducted using Agilent G200 equipment (Agilent, Santa Clara, CA, USA). When conducting nanohardness testing, the surface of the sample is first polished to obtain a smooth sample, followed by a large number of nanoindentation tests, and finally, the average nanohardness value is calculated. The nanohardness results are calculated using the formula H = Pmax/Ac (Pmax: Maximum compression load, Ac: real contact projection area between the indenter and the material) [8]. Each sample underwent at least five indentation tests. When conducting friction-wear tests, samples of the same size are prepared using materials with different deformation amounts and then tested separately using an RBT-1 friction and wear testing machine (Huahui, Lanzhou, China). The force of each test was 100 N, and the speed was 100 r/min. Each sample underwent at least three tests. After the experiment, analyze the recorded data and observe the surface wear of the tested sample using scanning electron microscopy to analyze the effect of different deformation variables on the wear resistance of 304 stainless steel material.
Corrosion tests were conducted on the stretched material, including the salt spray corrosion test and electrochemical corrosion test. During the salt spray test, block samples of the same size were prepared using stainless steel samples with different deformation amounts. After cleaning all samples, they were placed in a salt spray box simultaneously. The salt spray box equipment was equipped with a temperature control system and can set specific temperatures. The salt spray test solution was a 50 g/L NaCl solution at a temperature of 35 ± 2 °C for 7 days, and the experimental results were recorded. Three samples with the same deformation degree were prepared. At room temperature, the appearance of the tested sample was inspected, and the results of the salt spray test were evaluated. The influence of deformation on the salt spray corrosion resistance of stainless steel based on experimental results was analyzed. In addition, electrochemical workstations were used to conduct dynamic polarization tests on materials with different deformation treatments to study the pitting behavior of stainless steel. During the experiment, the back of the obtained stainless steel sample was welded to an insulated copper wire, and epoxy resin yellow feldspar was embedded in a plastic tube to make a sample with a test area of 1 cm2. Then, the sample was gradually polished with SiC sandpaper of different particle sizes to obtain a smooth surface, and the sample surface was cleaned with ultrasonic waves before the experiment. The experiment was conducted using an IviumStat electrochemical workstation, with a stainless steel sample as the working electrode, platinum plate as the counter electrode, and 232 saturated calomel electrode (SCE) as the reference electrode (0.244 V relative to the standard hydrogen electrode at 25 °C). The temperature of the experiment was kept at 25 ± 1 °C. Before the experiment, the sample was placed in the solution for about 10 min to make the experiment more stable. Each experiment should be repeated at least five times to ensure its reliability. The dynamic polarization experiment was conducted in a 0.9% NaCl electrolyte solution. The experiment scanned from (−1.0) VSCE to (+1.0) VSCE at a rate of (+0.1) mV/s to obtain the dynamic polarization curve. After the test is completed, observe the dynamic polarization curve and analyze the effect of different deformation amounts on the corrosion resistance of stainless steel.

3. Results

3.1. Microstructural Characterization

Figure 3a–c shows the metallographic images of the surface morphology of 304 stainless steel with different deformation degrees. Observation of Figure 3 reveals that the undeformed 304 stainless steel consists of irregular grains. At a small deformation degree (20%), the change is minor. The grain size distribution of the undeformed and 20% deformed specimens is similar, with most grains ranging from 70 to 110 μm. However, at a large deformation degree (40%), the surface structure of the stainless steel becomes very chaotic. Grains are deformed along the deformation direction, simultaneously destroyed, generating many fine grain clusters internally, and the grain size significantly decreases, mostly ranging from 10 to 30 μm. Different deformation degrees have significant differences in their impact on the grain structure of 304 stainless steel.
Figure 3(a1–c1) shows the EBSD inverse pole figure maps of 304 stainless steel with different deformation degrees. It is evident that the microstructure of 304 stainless steel undergoes significant changes with increasing deformation. From the figure, it can be seen that the solution-treated 304 stainless steel without deformation exhibits numerous irregular austenite grains. With increasing deformation, the specimen deformed by 20% has polygonal grains similar to the undeformed specimen, but the internal orientations of the grains gradually become chaotic. When the deformation reaches 40%, the grain orientations are highly chaotic, the original grains are severely destroyed, and they are significantly elongated along the deformation direction. The test results are consistent with the metallographic test results. Additionally, the ferrite scope was used to measure the content of transformation-induced martensite in the deformed austenite, and the results are shown in Figure 4. The results show that after 10%, 20%, and 40% deformation, the transformation-induced martensite content in 304 stainless steel is approximately 7.4%, 15.6%, and 68.2%, respectively. It can be seen that at a small deformation degree, martensitic transformation barely occurs in 304 stainless steel. As the deformation degree increases, the martensite content gradually increases. At a large deformation degree, the martensite content exceeds the austenite content, occupying the majority. The content of austenite and martensite in stainless steel will significantly affect its properties.

3.2. Surface Passive Film Characteristics

Typically, the characteristics of the passive film on the surface of austenitic stainless steel affect its corrosion resistance [9,10,11]. To analyze the corrosion behavior of stainless steel, XPS was used to analyze the characteristics of the passive films on the surfaces of 304 stainless steel specimens with different deformation conditions after anodic voltage polarization in NaCl medium. First, the thickness differences in the passivation films of the samples were relatively small, all ranging from 5.8 to 6.5 nm. Furthermore, the composition of the passivation film was analyzed through XPS test results. During XPS data processing, the background was first subtracted using Shirley’s method. Then, based on binding energy, the spectra of important components in the passive film were deconvoluted, and the peaks were separated into different components [12]. The test results show that peaks of Fe, Cr, Ni, and O were identified in the XPS tests, as shown in Figure 5.
Figure 6 shows the XPS spectra of Fe 2p3/2, Cr 2p3/2, and O 1s within the passive films on the surfaces of specimens with different deformation degrees. Through fitting, in Figure 6(a1,b1,c1), the Cr 2p3/2 spectra are separated into three components: Crmet (573.8 eV), Cr2O3 (576.2 eV), and CrOH (577.0 eV). Empirically, the chromium oxide component in the passive film on stainless steel significantly affects film stability. Higher chromium oxide content leads to better corrosion resistance of the passive film [13,14]. Observation of the figure shows that the height of the Cr2O3 peak gradually decreases with increasing deformation. Furthermore, the content of Cr2O3 in the passive films on stainless steel surfaces with different deformation degrees was calculated using software. The calculation results show that the Cr2O3 content in the passive films of 0%, 20%, and 40% deformed stainless steel specimens is 33.29%, 32.08%, and 30.11%, respectively. It is evident that the Cr2O3 content in the passive film gradually decreases with increasing deformation. The Fe 2p3/2 spectra are separated into four component peaks at 706.1 eV, 707.9 eV, 710.1 eV, and 711.6 eV, corresponding to Femet (metallic state), Fe3O4, FeO, and FeOOH components, as shown in Figure 6(a2,b2,c2). According to the fitting results, the intensity of the Femet peak is significantly higher than that of Fe3O4, FeO, and FeOOH. Additionally, studies show that higher oxide content in the passive film leads to better corrosion resistance [15,16]. This study analyzed the O 1s composition in the passive films on stainless steel samples with different deformation degrees, and the results are shown in Figure 6(a3,b3,c3). The O 1s spectra can be divided into three components, with corresponding peaks for O2− (531.1 eV), OH (531.9 eV), and H2O (533.6 eV). Clearly, the intensity of the O2− peak associated with Cr and Fe elements is much higher than that of the OH peak. Moreover, the results show that the intensity of the O2− peak changes with deformation. The above results indicate that different deformation degrees affect the oxide content, particularly chromium oxides, in the passive film on the surface of stainless steel, thereby altering the corrosion resistance of the passive film and consequently affecting the corrosion behavior of the stainless steel.

3.3. Mechanical Properties Testing

3.3.1. Nanoindentation Testing

Figure 7 shows the nanoindentation load–displacement curves for 304 stainless steel with different deformation degrees, with an inset showing the indentation image. It can be seen that as the load applied to specimens with different deformation degrees increases to 1 mN and during the hold at maximum load, the maximum depth that can be indented is significantly different. Observation reveals that as the deformation degree increases, the load–displacement curve shifts significantly to the left, and the maximum indentation depth noticeably decreases, indicating that the nanohardness of the specimen gradually increases. Table 2 is the statistical table of nanoindentation experimental results for 304 stainless steel with different deformation degrees. According to the experimental results, the nanohardness of 304 stainless steel gradually increases with increasing deformation degree. The results demonstrate that the nanohardness of 304 stainless steel gradually increases with increasing tensile deformation. The results are consistent with the load–displacement graph.

3.3.2. Friction-Wear Testing

Figure 8 shows the surface morphology images of 304 stainless steel after friction-wear tests with different deformation degrees. It can be seen from the figure that the worn surfaces of stainless steel all have shallow plowing grooves, with very small wear debris and slight material detachment around the grooves. The wear scar width of the specimens changes with the deformation degree. When the deformation degrees of 304 stainless steel are 0%, 5%, 10%, 20%, and 40%, the wear scars of the stainless steel specimens gradually narrow to 1.37, 1.23, 1.16, 1.05, and 0.75 mm, respectively. With increasing deformation, the wear scars of the 304 stainless steel material gradually narrow, indicating that as the deformation degree increases, the wear of the specimen gradually decreases, and the wear resistance becomes more excellent. This is because as the deformation degrees increase, the microstructure of the stainless steel material hardens, leading to increased adhesion between the friction pair and the specimen during the friction test. In the subsequent wear process, material transfer between the friction pair and the specimen becomes more difficult, resulting in reduced wear scars on the stainless steel material [17].

3.4. Corrosion Resistance Testing

3.4.1. Salt Spray Corrosion Test

Salt spray corrosion tests were conducted on 304 stainless steel with different tensile deformation degrees, and their surfaces were observed. Figure 9 shows the surface morphology images of 304 stainless steel with different tensile deformation degrees after the salt spray corrosion test. As seen from the figure, with increasing tensile deformation, the surface corrosion of stainless steel intensifies. Pitting and locally fragmented forms of corrosion products appear on the surface of stainless steel specimens. Pitting corrosion products are formed due to the pitting corrosion of stainless steel. As corrosion time prolongs, the corrosion products gradually expand from their initial pitting form and subsequently accumulate and overlap [18]. Table 3 shows the statistical results of mass changes before and after the salt spray corrosion test for 304 stainless steel with different deformation degrees. From the statistical results, it can be seen that with increasing deformation degree, the average mass loss of 304 stainless steel after corrosion gradually increases, indicating that the corrosion resistance of 304 stainless steel gradually decreases with increasing deformation. The results are consistent with the surface corrosion observations.

3.4.2. Electrochemical Corrosion Testing

To further investigate the changes in the corrosion performance of stainless steel material after deformation, potentiodynamic polarization tests were performed on specimens treated with different deformation degrees. In the potentiodynamic polarization result curves, the corrosion potential, Ec, and corrosion current density, ic, reflect the corrosion resistance of the material to some extent. The pitting potential, Ep, reflects the material’s resistance to pitting corrosion. A smaller ic value indicates a lower corrosion rate and better corrosion resistance. A larger Ec value indicates better corrosion resistance. A larger Ep value indicates better corrosion resistance [19,20].
Figure 10 shows the dynamic polarization test results for 304 stainless steel specimens under different deformation conditions. The dynamic polarization curve results show that before and after tensile deformation, the shape of the dynamic polarization curve for 304 stainless steel changes little, but the parameter values change significantly. Table 4 lists the key parameters of the dynamic polarization curves. The results show that after the deformation of 304 stainless steel, the Ep and Ec values of the specimens decrease, and the ic value significantly increases, indicating that the corrosion resistance of 304 stainless steel deteriorates after tensile deformation. Furthermore, with increasing deformation degree, the Ep and Ec values gradually decrease, while the ic value gradually increases, indicating that the corrosion resistance of 304 stainless steel gradually decreases with increasing deformation degree.

4. Discussion

According to the above research results, at small deformation degrees, the grains of 304 stainless steel change little. However, as the deformation degree increases, the grain structure of stainless steel is destroyed, and the internal grain orientations gradually become chaotic. With increasing deformation, martensite forms in the stainless steel, and the content of transformation-induced martensite gradually increases. Additionally, deformation affects the compositional content of the passive film on the stainless steel surface. With increasing deformation degree, the oxide content in the passive film decreases. Changes in microstructure further affect its macroscopic properties. With increasing deformation degree in 304 stainless steel, its nanohardness and wear resistance improve, while its corrosion resistance decreases.
After plastic deformation, the grain structure in austenitic stainless steel is destroyed (Figure 3) and, simultaneously, transformation-induced martensite is generated (Figure 4). The martensite crystal structure is bcc, with a lower packing density than the fcc austenite and different atomic arrangements. Lattice distortion stress fields caused by carbon and substitutional atoms in martensite are stronger, offering greater resistance to dislocation movement. An increase in martensite content makes deformation during hardness measurement more difficult, increasing the hardness value [21,22]. The martensite phase itself possesses high hardness, and its formation also directly enhances the overall hardness of the material [23]. Additionally, plastic deformation causes a significant increase in dislocation density. High-density dislocations entangle and intersect, forming complex dislocation networks, dislocation tangles, and dislocation cell walls. These dislocation structures constitute strong obstacles. To continue plastic deformation (i.e., to move dislocations), greater stress must be applied to overcome these obstacles (cutting dislocation lines, overcoming stress fields, bypassing obstacles, etc.) [24,25]. The hardness and wear resistance of a material essentially reflects its ability to resist localized plastic deformation. After the plastic deformation of stainless steel, numerous dislocations hinder deformation, manifesting macroscopically as an increase in the material’s hardness and wear resistance. After the plastic deformation of 304 stainless steel, with increasing deformation degree, the distortion of its lattice structure, the increase in martensitic transformation, and the accumulation of dislocations enhance its hardness and wear resistance.
Regarding corrosion performance, plastic deformation of stainless steel increases surface roughness and generates micro-cracks or crevices, which can tear off or scrape away the surface passive film, exposing the unpassivated fresh metal substrate. This provides a direct starting point for corrosion reactions. Furthermore, during the plastic deformation of 304 austenitic stainless steel, austenite transforms into martensite. During martensite formation, the diffusion of solute elements (particularly chromium) may be incomplete or non-uniform. Martensite may preferentially precipitate carbides, leading to the formation of chromium-depleted zones around it. Chromium is a key element for passive film formation. The presence of chromium-depleted zones makes it difficult to form or easily destroy the passive film in those regions [26], as shown by the passive film test results in Figure 6. Moreover, the newly formed martensite phase differs in composition, structure, and stress state from the surrounding austenite matrix, leading to a small potential difference between them, forming a corrosion microcell that accelerates the dissolution of the martensite regions [27]. Dislocations are line defects in crystals. The region around the dislocation line has lattice distortion, possessing higher internal energy and reactivity. The high-density dislocation network in plastically deformed stainless steel also provides preferential paths for anodic dissolution during corrosion, reducing its corrosion resistance [28]. The schematic diagram of the deformation, microstructure, and property evolution of 304 stainless steel was shown in Figure 11. Plastic deformation of stainless steel materials under different conditions affects their microstructure, which in turn influences their macroscopic mechanical and corrosion properties. This research can provide a theoretical reference value for the expanded and widespread application of stainless steel.

5. Conclusions

This paper subjected 304 stainless steel to tensile deformation and investigated the microstructure, mechanical properties, and corrosion resistance of 304 stainless steel with different deformation degrees. The following conclusions were drawn through analysis:
(1)
With increasing deformation degree in 304 stainless steel, grains are destroyed, generating numerous subgrains internally, grain size significantly decreases, crystal orientations become chaotic, and the content of transformation-induced martensite increases. Additionally, with increasing deformation, the oxide content in the passive film on the stainless steel surface decreases.
(2)
With increasing deformation degree, the nanohardness of 304 stainless steel gradually increases, and its wear resistance gradually improves.
(3)
With increasing deformation degree, the resistance of 304 stainless steel to salt spray corrosion and pitting corrosion gradually decreases.
(4)
The destruction of grains, martensitic transformation, dislocation accumulation, and changes in surface passive film characteristics caused by plastic deformation lead to the increase in nanohardness and wear resistance of 304 austenitic stainless steel but cause the deterioration of its corrosion resistance.

Author Contributions

Methodology, H.T. and Y.C.; Investigation, H.T., Y.C., Z.T., and M.D.; Validation, Z.L. and M.D.; Resources, Z.L., Y.C., and M.D.; Writing—original draft, H.T., Z.L., and H.X.; Writing—review and editing, H.T., H.X., and Z.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Nanxun Scholars Program for Young Scholars of ZJWEU (RC2022021035) and Fundamental Commonweal Research of Zhejiang Province (grant No. LGG22E050034).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data that support the findings of this study are included within the article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. The authors declare that the research did not involve human participants or animals.

References

  1. Kariksiz, M.; Kirkik, D. Antibacterial evaluation of antibiotic-coated titanium and stainless steel implants in orthopaedic application: A dip-coating approach. J. Orthop. Surg. Res. 2025, 20, 37. [Google Scholar] [CrossRef] [PubMed]
  2. Gaddam, S.; Haridas, R.S.; Tammana, D.; Sanabria, C.; Lammi, C.J.; Berman, D.; Mishra, R.S. Double-sided friction stir welding of Nitronic-40 stainless steel for application in tokamak devices. Mater. Sci. Technol. 2023, 159, 170–183. [Google Scholar] [CrossRef]
  3. Chen, X.; Xiao, H.; Shao, P.; Huang, S.; Shi, Y.; Liu, K. Microstructure evolution and deformation behavior of high strength titanium clad steel plate in the thermal compression. Steel Res. Int. 2024, 95, 2300834. [Google Scholar] [CrossRef]
  4. Zergani, H.M.R. Unraveling the effect of deformation temperature on the mechanical behavior and transformation-induced plasticity of the sus304l stainless steel. Steel Res. Int. 2020, 91, 2000114. [Google Scholar] [CrossRef]
  5. Chen, Z.; Wang, F.; Li, G.; Fan, Y.; Li, P.; Liu, M.; Jiang, K. Effect of deformation on microstructure evolution and mechanical properties of bismuth-containing austenitic stainless steels: A molecular dynamics study. Phys. Met. Metallogr. 2025, 126, 146–158. [Google Scholar] [CrossRef]
  6. Fatimah, S.; Bahanan, S.; Kang, J.-H.; Widiantara, I.P.; Ko, Y.G. Deep penetration of shear deformation in ferritic stainless steel via differential speed rolling considering contact condition. Appl. Sci. 2025, 15, 155. [Google Scholar] [CrossRef]
  7. Sohrabi, M.J.; Mirzadeh, H.; Sadeghpour, S.; Mahmudi, R. Dependency of work-hardening behavior of a metastable austenitic stainless steel on the nucleation site of deformation-induced martensite. Mater. Sci. Eng. A 2023, 868, 7. [Google Scholar] [CrossRef]
  8. Vidhyasankari, N.; John, R.R.; Senthilmurugan, P.R.; Vishnupriya, V. Comparative evaluation on surface nanohardness, surface microhardness, surface roughness, and wettability of plant-based organic nanoparticle reinforced polyetheretherketone as an implant material-An in vitro study. J. Indian Prosthodont. Soc. 2024, 24, 245–251. [Google Scholar] [CrossRef] [PubMed]
  9. Zhang, Y.; Sun, J.; Lei, H.; Jiao, F.; Niu, W.; Xu, X.; Lin, X. Influence of citric acid passivation treatment on the marine atmospheric corrosion resistance of L80-13Cr stainless steel. Discov. Appl. Sci. 2025, 7, 226. [Google Scholar] [CrossRef]
  10. Wang, L.; Liu, Z.; Chen, X.; Li, R.; Xie, Y. Effect of pre-filming potential on the electrochemical characteristics of 2205 duplex stainless steel in NaCl solution. 2024 International Conference on Corrosion Protection and Application. J. Phys. Conf. Ser. 2024, 2821, 012021. [Google Scholar] [CrossRef]
  11. Choudhary, S.; Kelly, R.G.; Birbilis, N. On the origin of passive film breakdown and metastable pitting for stainless steel 316L. Sci. C. 2024, 230, 13. [Google Scholar] [CrossRef]
  12. Pshyk, O.V.; Patidar, J.; Cancellieri, C.; Siol, S. Auger parameter analysis for TiN and AlN thin films via combined in-situ XPS and HAXPES Empa. arXiv 2025, arXiv:2505.20145. [Google Scholar]
  13. Zhao, S.; Lin, D.; Qin, L.; Sun, M.; Liao, K.; Du, X.; Chen, Y. Passivation film characteristics of 2205 duplex stainless steel in reboiler tubes under the combined effects of dissolved oxygen and applied stress. Arab. J. Sci. Eng. 2025, 50, 5007–5019. [Google Scholar] [CrossRef]
  14. Gao, T.; Liu, Z.; Yang, L.; Tang, Y.; Lu, D.; Yu, Y.; Yang, H.; Qiao, Y. Passivation behavior of different building planes of selective laser melting 316l stainless steel in 3.5% NaCl solution. Steel Res. Int. 2023, 94, 2200759. [Google Scholar] [CrossRef]
  15. Zhao, K.; Li, X.-Q.; Wang, L.-W.; Yang, Q.-R.; Cheng, L.-J.; Cui, Z.-Y. Passivation behavior of 2507 super duplex stainless steel in hot concentrated seawater: Influence of temperature and seawater concentration. J. Mater. Sci. Technol. 2022, 35, 326–340. [Google Scholar] [CrossRef]
  16. Lv, S.; Li, K. Semiconducting Behaviour and corrosion resistance of passive film on corrosion-resistant steel rebars. Materials 2022, 15, 7644. [Google Scholar] [CrossRef] [PubMed]
  17. Ralls, A.M.; Leong, K.; Liu, S.; Wang, X.; Jiang, Y.; Jiang, Y. Unraveling the friction and wear mechanisms of surface nanostructured stainless-steel. Wear 2024, 538–539, 205185. [Google Scholar] [CrossRef]
  18. Weng, C.; Tang, J.; Yu, W.; Tang, Z.; Qin, R. Effect of aging treatment on long-term salt spray corrosion resistance of 15-5pH stainless steel. Metall. Mater. Trans. A 2025, 56, 1363–1377. [Google Scholar] [CrossRef]
  19. Yang, L.; Gong, F.; Zhang, X.; Xia, J.; Liu, X. Effect of cutting parameters on pitting corrosion resistance of S32760 duplex stainless steel. Mater. Corros. Werkst. Und Korros. 2024, 75, 11. [Google Scholar] [CrossRef]
  20. Zhu, S.; Wang, B.; Yan, B. Effect of arc remelting on microstructure and pitting corrosion resistance of 441 ferritic stainless steel. Kov. Mater. 2023, 61, 115–120. [Google Scholar] [CrossRef]
  21. Zhai, D.; Zou, Z.; Jin, M.; Cui, Z.; Wang, T. Constitutive model of trip-type duplex stainless steel under symmetrical strain cyclic loading considering the effect of martensitic transformation. Metall. Mater. Trans. Part A 2025, 56, 587–599. [Google Scholar] [CrossRef]
  22. Han, P. Effect of silicon on the martensitic nucleation and transformation of 301 stainless steel under various cold-rolling deformations. Metals 2024, 14, 827. [Google Scholar] [CrossRef]
  23. Li, B.; Zhang, H.; Wu, S.; Xie, L.; Shen, F.; Xin, J.; Huang, C.; Wang, W.; Li, L. Mechanical properties and martensitic transformation behavior of 316ln stainless steel under cryogenic deformation. Steel Res. Int. 2024, 95, 9. [Google Scholar] [CrossRef]
  24. Mo, j.; Li, j.; Xu, M.; Zhang, W.; Niu, G.; Feng, G.; Pang, Q. Strengthening mechanism of antibacterial stainless steel based on heterogeneous nano/ultrafine austenite with high dislocation density. JOM 2025, 77, 2698–2705. [Google Scholar] [CrossRef]
  25. Li, J.; Jiang, W.; Zhang, Y.; Liu, L.; Yu, Y.; Luan, J.; Jiao, Z.; Liu, C.T.; Zhang, Z. Formation of core-shell nanoprecipitates and their effects on work hardening in an ultrahigh-strength stainless steel. Int. J. Plast. 2025, 184, 104184. [Google Scholar] [CrossRef]
  26. Wang, B.; Li, Y.; Li, H.; Zhao, G.; Song, Y.; Xu, H. Effect of recrystallization degree on properties of passive film of super ferritic stainless steel S44660. Corros. Rev. 2024, 42, 455–470. [Google Scholar] [CrossRef]
  27. Assumpo, R.F.; Georges, M.; Guo, X.; Farias, F.W.C.; Santos, D.B.; Frankel, G.S.; Sicupira, D.C. Reverse martensitic transformation on the corrosion behavior of a 2304 lean duplex stainless steel. Corros. Sci. 2025, 245, 112690. [Google Scholar] [CrossRef]
  28. Majid, A.; Iswanto, P.T.; Irfanuddin, F. Effect of cold rolling and shot peening on physical and mechanical properties of 316L stainless steel. AIP Conf. Proc. 2025, 3120, 020010. [Google Scholar]
Figure 1. Tensile specimen size (mm).
Figure 1. Tensile specimen size (mm).
Coatings 15 00845 g001
Figure 2. Stress–strain curve of 304 stainless steel.
Figure 2. Stress–strain curve of 304 stainless steel.
Coatings 15 00845 g002
Figure 3. Metallography of 304 stainless steel with different deformation degrees ((a): 0%; (b): 20%; (c): 40%) and EBSD reverse pole diagram ((a1): 0%; (b1): 20%; (c1): 40%).
Figure 3. Metallography of 304 stainless steel with different deformation degrees ((a): 0%; (b): 20%; (c): 40%) and EBSD reverse pole diagram ((a1): 0%; (b1): 20%; (c1): 40%).
Coatings 15 00845 g003
Figure 4. Martensitic content of 304 stainless steel with different deformation degrees.
Figure 4. Martensitic content of 304 stainless steel with different deformation degrees.
Coatings 15 00845 g004
Figure 5. XPS study of passivation film formed on the surface of 304 stainless steel with different deformation in NaCl medium.
Figure 5. XPS study of passivation film formed on the surface of 304 stainless steel with different deformation in NaCl medium.
Coatings 15 00845 g005
Figure 6. Composition of passivation film on the surface of 304 stainless steel with different deformation degrees (a1a3): deformation 0%; (b1b3): deformation of 20%; (c1c3): deformation of 40%; (a1,b1,c1): Cr 2p3/2; (a2,b2,c2): Fe 2p3/2; and (a3,b3,c3): O 1s).
Figure 6. Composition of passivation film on the surface of 304 stainless steel with different deformation degrees (a1a3): deformation 0%; (b1b3): deformation of 20%; (c1c3): deformation of 40%; (a1,b1,c1): Cr 2p3/2; (a2,b2,c2): Fe 2p3/2; and (a3,b3,c3): O 1s).
Coatings 15 00845 g006
Figure 7. Nanoindentation load–displacement curves for 304 stainless steel with different deformation degrees.
Figure 7. Nanoindentation load–displacement curves for 304 stainless steel with different deformation degrees.
Coatings 15 00845 g007
Figure 8. Wear morphology of 304 stainless steel with different deformation degrees. (a) deformation 0%; (b) deformation 5%; (c) deformation 10%; (d) deformation 20%; (e) deformation 40%.
Figure 8. Wear morphology of 304 stainless steel with different deformation degrees. (a) deformation 0%; (b) deformation 5%; (c) deformation 10%; (d) deformation 20%; (e) deformation 40%.
Coatings 15 00845 g008
Figure 9. Surface morphology of 304 stainless steel with different deformation degrees after salt spray corrosion test. (a) deformation 5%; (b) deformation 10%; (c) deformation 20%; (d) deformation 40%.
Figure 9. Surface morphology of 304 stainless steel with different deformation degrees after salt spray corrosion test. (a) deformation 5%; (b) deformation 10%; (c) deformation 20%; (d) deformation 40%.
Coatings 15 00845 g009
Figure 10. Dynamic polarization curves of 304 stainless steel with different deformation degrees.
Figure 10. Dynamic polarization curves of 304 stainless steel with different deformation degrees.
Coatings 15 00845 g010
Figure 11. Schematic diagram of the deformation, microstructure and property evolution of 304 stainless steel.
Figure 11. Schematic diagram of the deformation, microstructure and property evolution of 304 stainless steel.
Coatings 15 00845 g011
Table 1. 304 stainless steel composition (%).
Table 1. 304 stainless steel composition (%).
MaterialCSiMnPSCrNi
3040.070.331.130.0390.02218.098.06
Table 2. Nanohardness values of 304 stainless steel with different deformation degrees.
Table 2. Nanohardness values of 304 stainless steel with different deformation degrees.
MaterialDeformation Degree (%)Nanohardness Value (GPa)
3040%4.0 ± 0.1–5.0 ± 0.2
5%4.0 ± 0.1–5.5 ± 0.2
10%4.5 ± 0.1–5.5 ± 0.2
20%5.0 ± 0.2–7.5 ± 0.3
40%6.0 ± 0.3–7.5 ± 0.3
Table 3. Salt spray corrosion test results of 304 stainless steel with different deformation degrees.
Table 3. Salt spray corrosion test results of 304 stainless steel with different deformation degrees.
MaterialDeformation Degree (%)Average Mass Loss/gAverage Corrosion Rate/(g/m2·h)
3040%0.0021 ± 0.00020.0125 ± 0.001
5%0.0023 ± 0.00020.0140 ± 0.001
10%0.0027 ± 0.00030.0161 ± 0.002
20%0.0029 ± 0.00030.0173 ± 0.002
40%0.0032 ± 0.00030.0190 ± 0.002
Table 4. Dynamic polarization test results of 304 stainless steel with different deformation degrees in 3.5% NaCl solution.
Table 4. Dynamic polarization test results of 304 stainless steel with different deformation degrees in 3.5% NaCl solution.
Deformation DegreeEp
(mVSCE)
Ec
(mVSCE)
ic
(nAcm−2)
0%398.7 ± 2.2−131.4 ± 2.523.04 ± 0.14
5%360.4 ± 1.9−156.6 ± 2.122.01 ± 0.11
10%315.1 ± 1.7−179.1 ± 2.329.55 ± 0.08
20%252.9 ± 1.2−194.1 ± 1.721.74 ± 0.04
40%221.7 ± 2.3−225.6 ± 1.552.58 ± 0.06
Ep: Pitting potential; Ec: Corrosion potential; ic: Corrosion current density.
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

Tao, H.; Cai, Y.; Li, Z.; Xiu, H.; Tong, Z.; Ding, M. Research on the Synergistic Evolution Law of Microstructure and Properties of Deformed Austenitic Stainless Steel. Coatings 2025, 15, 845. https://doi.org/10.3390/coatings15070845

AMA Style

Tao H, Cai Y, Li Z, Xiu H, Tong Z, Ding M. Research on the Synergistic Evolution Law of Microstructure and Properties of Deformed Austenitic Stainless Steel. Coatings. 2025; 15(7):845. https://doi.org/10.3390/coatings15070845

Chicago/Turabian Style

Tao, Huimin, Yafang Cai, Zi Li, Haiteng Xiu, Zeqi Tong, and Mingming Ding. 2025. "Research on the Synergistic Evolution Law of Microstructure and Properties of Deformed Austenitic Stainless Steel" Coatings 15, no. 7: 845. https://doi.org/10.3390/coatings15070845

APA Style

Tao, H., Cai, Y., Li, Z., Xiu, H., Tong, Z., & Ding, M. (2025). Research on the Synergistic Evolution Law of Microstructure and Properties of Deformed Austenitic Stainless Steel. Coatings, 15(7), 845. https://doi.org/10.3390/coatings15070845

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