Startseite Eddy current methodology in the non-direct measurement of martensite during plastic deformation of SS316L
Artikel Open Access

Eddy current methodology in the non-direct measurement of martensite during plastic deformation of SS316L

  • Dominik Kukla EMAIL logo , Adam Kondej , Sylwester Jończyk , Piotr Lasota , Jakub Tabin , Daniela Schob , Robert Roszak , Jakub Kawałko , Andrzej Zagórski und Mateusz Kopec
Veröffentlicht/Copyright: 24. Juni 2025
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

This study examines the use of various eddy current induction techniques to evaluate the stability of austenite in SS316L steel subjected to plastic deformation. This deformation, which occurs locally in austenitic steel structures under operational loads, leads to a martensitic transformation. This transformation affects both the mechanical and magnetic properties of the steel. The martensitic phase content, being ferromagnetic, can be quantitatively assessed using a ferritoscope and other magnetic induction methods. The research explores techniques based on the analysis of impedance signal changes obtained using the NORTEC defetoscope and the WIROTEST device developed by the author’s team. By examining the phase angle, ET signal amplitude, and resonance frequency changes in the eddy current excitation system, the study aims to quantitatively assess the martensitic phase content in samples subjected to plastic deformation. These results were verified through comparison with data from a ferritoscope and X-ray diffraction analysis. Additionally, the eddy current technique facilitates surface screening of the specimen, making it possible to identify cracks and locate the martensitic transformation front in areas of stress concentration.

1 Introduction

Austenitic steels are widely used in various industrial applications due to their high mechanical strength, corrosion resistance, and the stability of their non-magnetic properties, particularly in electrical and energy sectors. In these applications, maintaining the stability of paramagnetic austenite is crucial, as the formation of a ferromagnetic martensitic phase can occur under operational loads, both mechanical and thermal. The stability of austenite is primarily influenced by its chemical composition and the content of stabilizing alloying elements.

Austenitic steels are susceptible to martensitic transformation under plastic deformation conditions [1], and this process is more intensive at low temperatures [2]. It has been reported that in the case of 304 steel, ε-martensite (hexagonal close-packed [HCP], paramagnetic) forms at the beginning of deformation and reaches a peak value at about 5% of tensile strain. After exceeding this value, the volume fraction of this phase decreases to almost zero at 20% of plastic strain. On the other hand, the volume fraction of α′-martensite (body-centered cubic [BCC], ferromagnetic) increased steadily with deformation. At higher strain values, the α′ phase was the only martensite phase present in the material volume [3]. The transformation is accompanied by a change in the magnetic properties of the steel, which allows the use of magnetic techniques to assess its effects [4].

Similar results were reported for the AISI 316 steel deformed during rolling and subsequent tension test [5]. These properties are of great importance in the case of parts manufactured using additive technologies, because during the printing process, ferrite is precipitated at the grain boundaries of austenitic steel, the share of which reaches 6% [6,7]. Since the ferrite volume fraction in the structure of austenitic steel can adversely affect fatigue strength [8,9] it is advisable to control its content. The formation of martensite can be detected and evaluated by X-ray diffraction (XRD) [10] neuron diffraction [11], electron backscatter diffraction (EBSD) [12], magnetic methods including the eddy current method [13], metallography, and other techniques [14]. Monitoring the ferrite content released during martensitic transformation is crucial for evaluating the corrosion resistance of steel. Corrosion studies have demonstrated that martensitic transformation, induced by cold forming processes and cyclic loading conditions, significantly affects the corrosion resistance of AISI 316 austenitic stainless steel [15,16]. The formation of martensite is highly dependent on factors such as chemical composition, austenitic grain size [17], deformation temperature [18], and other variables. One should highlight that eddy current method is particularly sensitive to changes in the electrical resistivity and magnetic permeability of materials, which are influenced by variations in chemical composition, hardness, mechanical strength, and the degree of cold work. Consequently, this technique provides an indirect measurement of the overall physical and chemical properties of a material. It has been shown to be sensitive enough to detect phase transformations in stainless steels [19].

Plastic deformation of conventional and additive manufacture (AM) specimens may significantly influence their magnetic properties due to deformation-induced martensitic transformation [20]. The metastable ASSs exhibit tendency toward diffusion-free phase transformation during plastic deformation, especially when temperatures tend to 0 K [21]. In this process, initial austenite with a face-centered cubic (γ) structure transforms into martensite ε with a HCP structure and subsequently into martensite α′ with a BCC structure. It is worth noting that the transformation can also occur directly and, during tension at room temperature, mainly martensite α′ is observed [21,22].

Therefore, this article presents an investigation into the stability of austenite in 316 steel samples, produced from both rolled sheets and via AM, under static tensile loading conditions.

2 Methodology

The tests were conducted on two types of flat specimens: conventional (R) 316L and AM 316L.

The conventional specimens were fabricated using electrical discharge machining from a commercial 316L stainless steel sheet (2 mm thickness) with the tensile axis parallel to the rolling direction (RD). The semi-products were supplied in cold-rolled and annealed conditions (Figure 1). In such a form, the materials are typically used for structural components working in cryogenic conditions, down to the temperature of superfluid helium, 1.9.

Figure 1 
               Ultrafuse 316L filament, created by BASF for producing stainless steel parts through 3D printing, was employed (as an example of the possibilities of FFF technology).
Figure 1

Ultrafuse 316L filament, created by BASF for producing stainless steel parts through 3D printing, was employed (as an example of the possibilities of FFF technology).

The AM process utilizes Ultrafuse 316L, a metal–polymer composite filament specifically engineered for fused filament fabrication (FFF). This filament comprises over 80% metal powder, bound by a unique polymer matrix that is fully removed during debinding and sintering. The result is a high-strength, corrosion-resistant part composed entirely of 316L stainless steel. The printing process was conducted using an Ultimaker S5 printer with the following optimized parameters: a glass print platform coated with Magigoo PRO metal adhesive, a 0.4 mm hardened steel nozzle, platform temperature of 110°C, and printing temperature of 230°C. No print cooling was applied. The layer thickness was set to 0.15 mm, with a print speed of 25 mm/s and a fill density of 100%.

Debinding followed the Badische Anilin-und Soda-Fabrik (BASF) process, using a Nabertherm NRA 40/02-CDB debinding furnace at 120°C with HNO3 at a concentration of >98%. Nitric acid was fed at a typical rate of 30 L/h, along with nitrogen purging gas at 500 L/h, ensuring safe and consistent processing. The debinding phase concludes upon reaching a minimal mass loss of 10.5%.

The final stage, sintering, was conducted in a 100% pure, dry hydrogen atmosphere. The sintering cycle involved a controlled temperature ramp as follows:

  • From room temperature to 600°C at a rate of 5 K/min, followed by a 1 h hold at 600°C.

  • From 600 to 1,380°C at a rate of 5 K/min, followed by a 3 h hold at 1,380°C.

  • Gradual furnace cooling to complete the process.

The initial microstructure (Figure 3) of the as-received materials in all cases consists of equiaxial grains with recrystallization twins present in some of the larger grains. The average grain size in the conventional 316L steels, equaling 19.2 µm, while FFF 316L has average grain diameter of 37 µm. All steels have a heterogeneous distribution of grain sizes.

Figure 2 
               AM 316L specimen: (a) printed part and (b) final part after sintering.
Figure 2

AM 316L specimen: (a) printed part and (b) final part after sintering.

Figure 3 
               As-received microstructures of the investigated materials: inverse pole figure maps in the top row and grain size distributions (equivalent circle diameter) in the bottom row.
Figure 3

As-received microstructures of the investigated materials: inverse pole figure maps in the top row and grain size distributions (equivalent circle diameter) in the bottom row.

Each specimen was divided into 14 areas, 10 of which were in the gauge section (Figure 4). In each area, the ferrite content and electromagnetic parameters were measured.

Figure 4 
               Specimen cut from sheet metal R 316L (a) and printed AM 316L (b).
Figure 4

Specimen cut from sheet metal R 316L (a) and printed AM 316L (b).

The following tests were conducted to measure the ferrite content formed during the plastic deformation of specimens:

  • Manual measurement of the ferromagnetic martensitic phase content, using a DMP30 ferritoscope.

  • Manual measurement of impedance parameters (phase angle and amplitude) using the lift off technique and NORTEC 600 eddy current flaw detector.

  • Automatic measurement of the amplitude and resonance frequency of the eddy current signal using developed setup based on the WIROTEST device made by WIT Łukasiewicz (Figure 5).

  • Profile of the ferrite phase content using XRD and XRDynamic 500 diffractometer.

Ferrite phase content, phase angle, amplitude, and resonance frequency were measured in all specimens using the above-mentioned techniques before and after static tensile tests performed up to 8% in plastic deformation. Subsequently, one series of specimens was subjected to five loading–unloading tensile cycles, to obtain 1% deformation in the first cycle, 2% in the second, 4% in the third, 6% in fourth, and 8% of plastic deformation in the last, fifth cycle. After subsequent loading cycles, the ferrite content and impedance parameters were measured in each measuring area. Measurements with the WIROTEST setup (Figure 5) and XRD tests were performed after loading. The results obtained with these techniques refer only to the state before and after plastic deformation to a certain level.

Figure 5 
               Automated system for measuring the amplitude and frequency of the eddy current signal by WIROTEST.
Figure 5

Automated system for measuring the amplitude and frequency of the eddy current signal by WIROTEST.

The feritscope, widely used due to its speed and non-destructive nature, provides reliable results when calibrated appropriately [23]. The feritscope measures ferromagnetic microstructures in materials, primarily ferrite but also α′ martensite, which has a ferromagnetic BCC structure. It cannot measure ε martensite, as this paramagnetic, HCP phase does not generate a detectable magnetic signal. ε martensite is transitional, reducing with plastic strain and converting to α′ martensite at higher strains [23,24].

The feritscope detects voltage proportional to magnetic properties, but differences between ferrite and martensite magnetic permeability necessitate conversion of the readings. Talonen et al. [23] established a widely cited proportionality coefficient of 1.7 for martensite content, supported by other researchers [24].

Talonen et al. [23] demonstrated a linear relationship between martensite content and feritscope readings up to 55%, with a bilinear approximation beyond this. The formula for martensite content (%) based on feritscope readings is

M ( % ) = 1.7 Fr, Fr 50 0.5357 Fr + 58.2143 , Fr 50,78 .

The martensite content was measured using the feritscope Fisher FMP 30, an instrument that detects all magnetic phases, including ferrite. The martensite content for ASS can be calculated from the feritscope reading according to the formula presented by Talonen et al.

Thus, in the manuscript the ferrite content measured by means of commercial NORTEC 600 flaw detector was verified by feritoscope and WIROTEST measurement, as well as by XRD analysis.

3 Results

At the beginning of experimental procedure, three specimens of 316L steel (produced via rolling – R and AM) were subjected to static tensile loading at 77 K, inducing a similar plastic strain in AM and R. For these specimens, the results are reported for the states before and after deformation (Figures 7 and 8). For one pair of samples (AM and R), ferrite content measurements and commercial flaw detection data were collected after each cycle of tensile loading (Figures 9 and 10). In the case of AM 316L some initial ferrite grains are present at austenite grain boundaries as a result of high temperature sintering process, while the R316L sample presents typical single-phase microstructure (Figure 3).

Figure 6 
               Profiles of ferrite content measured along the symmetry axis for the rolled (R) and printed (AM) specimens after plastic deformation at 77 K, obtained by XRD method.
Figure 6

Profiles of ferrite content measured along the symmetry axis for the rolled (R) and printed (AM) specimens after plastic deformation at 77 K, obtained by XRD method.

Figure 7 
               Profiles of ferrite content (a), phase angle (b), and impedance amplitude (c) changes measured for the rolled (R) and printed (AM) specimens, before and after deformation at 77 K.
Figure 7

Profiles of ferrite content (a), phase angle (b), and impedance amplitude (c) changes measured for the rolled (R) and printed (AM) specimens, before and after deformation at 77 K.

Figure 8 
               Resonance frequency (a) and eddy current voltage amplitude (b) obtained by WIROTEST in the scanning mode. Ferrite content profiles for five states of plastic deformation, without elastic stresses (measurement at force 0) (c).
Figure 8

Resonance frequency (a) and eddy current voltage amplitude (b) obtained by WIROTEST in the scanning mode. Ferrite content profiles for five states of plastic deformation, without elastic stresses (measurement at force 0) (c).

Figure 9 
               Ferrite content profiles (a) and eddy current phase angle profiles for five states of plastic deformation measured during unloading (b) and loading (c).
Figure 9

Ferrite content profiles (a) and eddy current phase angle profiles for five states of plastic deformation measured during unloading (b) and loading (c).

Figure 6 presents the ferrite phase content measured using the XRD method for two specimens manufactured by different techniques. The graphs illustrate the profile of ferrite content percentages at successive measurement points along the sample axis. To achieve a substantial amount of ferrite, experiments were carried out at 77 K (liquid nitrogen temperature). At this temperature, the deformation-induced phase transformation in austenitic steels occurs with much greater intensity compared to similar deformation conducted at room temperature [21]. This allowed for a comparison of ferrite distribution between conventionally manufactured and additively manufactured 316L steel. One can observe increased ferrite phase content in the AM specimen, both in the non-plastically deformed area (grip part) and along the entire length of the gauge area (Figure 6).

Figure 10 
               Comparison of ferrite content profiles for two types of SS316L after deformation obtained by XRD (line) and feritoscope (points).
Figure 10

Comparison of ferrite content profiles for two types of SS316L after deformation obtained by XRD (line) and feritoscope (points).

The result shown in Figure 8a shows the ferrite contents in previously specified areas of the specimens measured by using ferritoscope. Uniaxial tensile tests were carried out at 77 K. Similar to the XRD measurements, an increased ferrite phase content was observed in the printed specimens, both, in the initial state and after deformation. The similar trend was observed during phase angle measurements (Figure 7b) and impedance amplitude measurements (Figure 7c) executed by applying an eddy current probe at the established measurement points. These are the results of the sintering process of AM specimens (Figure 2).

Subsequently, measurements of the resonance frequency and the voltage amplitude of the eddy current signal induced by the WIROTEST setup developed by the authors were presented. Figure 8a shows the frequency values obtained in automatic mode for two rolled and two AM specimens after plastic deformation. Since the WIROTEST setup also measures the voltage amplitude of the induced eddy currents, Figure 9b presents the results of these measurements for two pairs of specimens after plastic deformation. The eddy current amplitude changes along the axis of plastically deformed specimens indicating the possibility of detecting deformed areas and related sensitivity to the manufacturing method and the corresponded ferrite content. Furthermore, in the case of 316R, a local increase of the amplitude value was observed. Plastic front propagation is observed in 316L, but only at 77 K. This peak is related to the presence of the martensitic transformation front [22]. The results obtained for five static load cycles are presented only for the rolled 316L steel specimen. They include measurements of the ferrite content and the phase angle value after each cycle, in the unloaded stage (force 0) and at the maximum load applied (force max). The aim of such interrupted test was to assess the effect of elastic deformation on the measurable parameters. Figure 8c shows the increase in ferrite content along the specimen axis in subsequent cycles, i.e., as a function of increasing strain, for the unloaded variant. Figure 9a presents the changes in ferrite content measured at the maximum force of subsequent load cycles. With the increase of the plastic deformation level, differences in ferrite content are visible for the unloaded and maximum load conditions (Figure 9b and c). It is so called Villari effect [25,26].

4 Discussion

All applied research techniques demonstrate sensitivity to changes in the content of ferrite which result from manufacturing processes and deformation-induced phase transformation. The results of tests conducted under cyclic loading indicate the potential for identifying elastic deformation in both phases: austenite as well as in ferrite (Figures 8 and 9). This implies that flaw detector measurements, commonly used for detecting potential cracks, can also be utilized to identify and monitor the evolution of deformation-induced phase transformations in austenitic stainless steels. In the case of impedance amplitude measurements using the WIROTEST setup, two specimens exhibited a peak value near the end of the measurement section (points 10 and 11). This is likely associated with the location of the martensitic transformation front. However, this effect was not consistently observed. The results obtained from techniques based on magnetic induction demonstrate sensitivity to changes in magnetic permeability, which are linked to the presence of the ferromagnetic martensitic phase.

One should highlight the suitability of eddy current (EC) methodology in ferrite content identification. The comparison between feritscope and XRD techniques presented in Figure 10 proved the high efficiency of presented methodology.

5 Conclusions

Eddy current methodology is a promising non-destructive technique for indirectly measuring martensite formation in 316L during plastic deformation, though it comes with certain limitations. 316L, an austenitic stainless steel, can undergo deformation-induced phase transformation from austenite to martensite during, which it alters its magnetic properties. Since eddy currents are sensitive to changes in the material’s electrical conductivity and magnetic permeability, they can indirectly detect the presence of martensite, which is ferromagnetic, unlike the non-magnetic austenitic phase. This makes EC a suitable method for tracking martensitic transformation in real-time during mechanical loading without interfering with the deformation process. The high sensitivity of EC to local changes in phase composition is an advantage, allowing the detection of even small volume fraction of martensite. Additionally, EC is fast, scalable, and can be applied to complex geometries, making it versatile for in situ industrial applications. However, its sensitivity to other factors, such as temperature variations, strain-induced microstructural changes unrelated to martensite formation, and surface condition, may complicate accurate phase quantification. Calibration with complementary techniques like XRD or magnetic Barkhausen noise analysis might be required to enhance the accuracy of martensite content determination. Thus, while EC is suitable for detecting trends in martensitic transformation during plastic deformation, precise quantification and isolation of martensite may necessitate careful control of experimental conditions and the integration of additional methodologies.

Acknowledgments

Thanks to Mirosław Wyszkowski, Andrzej Chojnacki for help with low-temperature strength tests.

  1. Funding information: This work has been supported by the National Science Centre through Grant No UMO-2023/51/D/ST8/02370 and Polish National Agency for Academic Exchange through Grant No BPN/BDE/2023/1/00022/U/00001/ZU/00001. The research used devices developed as part of the project entitled “Development of technology for high-pressure gas hardening of satellite gears of the epicyclic aircraft gearbox of the FDGS engine, made of Pyrowear 53 steel and operating under long-term and cyclically variable operating loads” financed under the TECHMASTRATEG II competition by the National Centre for Research and Development (No.: TECHMATSTRATEG2/406725/1/NCBR/2020).

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results and approved the final version of the manuscript. D.K.: development of the research concept, performance of phase angle measurements, preparation of the manuscript; A.K.: measurements in the frequency range and signal amplitude; P.L.: performance of diffractometric tests and development of results; J.T.: development of the cryo-testing device and measurements of ferrite content: D.S.: EBSD tests; R.R., J.K.: preparation of specimen; A.Z.: development of the strength testing program; M.K.: development of research results and text editing.

  3. Conflict of interest: Authors state no conflict of interest.

  4. Data availability statement: The datasets generated during the current study are available from the corresponding author on reasonable request.

References

[1] Tavares SSM, Gunderov D, Stolyarov V, Neto JM. Phase transformation induced by severe plastic deformation in the AISI 304L stainless steel. Mater Sci Eng A. Oct 2003;358(1–2):32–6. 10.1016/S0921-5093(03)00263-6.Suche in Google Scholar

[2] Datta K, Delhez R, Bronsveld PM, Beyer J, Geijselaers HJM, Post J. A low-temperature study to examine the role of ε-martensite during strain-induced transformations in metastable austenitic stainless steels. Acta Mater. Jun 2009;57(11):3321–6. 10.1016/j.actamat.2009, 10.1016/S0925-8388(00)00874-4.Suche in Google Scholar

[3] Ferreri NC, Pokharel R, Livescu V, Brown DW, Knezevic M. Effects of heat treatment and build orientation on the evolution of ϵ and α′ martensite and strength during compressive loading of additively manufactured 304L stainless steel. Acta Mater. 2020;195:59–70. ISSN 1359-6454, 10.1016/j.actamat.2020.04.036.Suche in Google Scholar

[4] Tavares SSM, Fruchart D, Miraglia S. Magnetic study of the reversion of martensite α′ in a 304 stainless steel. J Alloy Compd. Jul 2000;307(1–2):311–7.10.1016/S0925-8388(00)00874-4Suche in Google Scholar

[5] Bruschi S, Simonetto E, Pigato M, Ghiotti A, Bertolini R. Analysis of the AISI 316 stainless steel sheet response to sub-zero deformation temperatures. Manuf Lett. Aug 2023;35:208–14. 10.1016/j.mfglet.2023.08.023.Suche in Google Scholar

[6] Gundgire T, Jokiaho T, Santa-aho S, Rautio T, Järvenpää A, Vippola M. Comparative study of additively manufactured and reference 316 L stainless steel samples – effect of severe shot peening on microstructure and residual stresses. Mater Charact. Sep 2022;191:112162. 10.1016/j.matchar.2022.112162.Suche in Google Scholar

[7] Liu H, Wei Y, Tan CKI, Ardi DT, Tan DCC, Lee CJJ. XRD and EBSD studies of severe shot peening induced martensite transformation and grain refinements in austenitic stainless steel. Mater Charact. Oct 2020;168:110574. 10.1016/j.matchar.2020.110574.Suche in Google Scholar

[8] Li Y, Luo Y, Li J, Song D, Xu B, Chen X. Ferrite formation and its effect on deformation mechanism of wire arc additive manufactured 308 L stainless steel. J Nucl Mater. Jul 2021;550:152933. 10.1016/j.jnucmat.2021.152933.Suche in Google Scholar

[9] Zhong Y, Rännar LE, Liu L, Koptyug A, Wikman S, Olsen J, et al. Additive manufacturing of 316L stainless steel by electron beam melting for nuclear fusion applications. J Nucl Mater. Apr 2017;486:234–45. 10.1016/j.jnucmat.2016.12.042.Suche in Google Scholar

[10] Gauss C, Souza Filho IR, Sandim MJR, Suzuki PA, Ramirez AJ, Sandim HRZ. In situ synchrotron X-ray evaluation of strain-induced martensite in AISI 201 austenitic stainless steel during tensile testing. Mater Sci Eng A. Jan 2016;651:507–16. 10.1016/j.msea.2015.10.110.Suche in Google Scholar

[11] Haušild P, Davydov V, Drahokoupil J, Landa M, Pilvin P. Characterization of strain-induced martensitic transformation in a metastable austenitic stainless steel. Mater Des. Apr 2010;31(4):1821–7. 10.1016/j.matdes.2009.11.008.Suche in Google Scholar

[12] Li J, Fang C, Liu Y, Huang Z, Wang S, Mao Q, et al. Deformation mechanisms of 304L stainless steel with heterogeneous lamella structure. Mater Sci Eng A. Jan 2019;742:409–13. 10.1016/j.msea.2018.11.047.Suche in Google Scholar

[13] Lois A, Ruch M. Assessment of martensite content in austenitic stainless steel specimens by eddy current testing. Insight Non Destr Test Cond Monit. 2006;48:26–9.10.1784/insi.2006.48.1.26Suche in Google Scholar

[14] Sohrabi MJ, Naghizadeh M, Mirzadeh H. Deformation-induced martensite in austenitic stainless steels: a review. Arch Civ Mech Eng. 2020;20:124. 10.1007/s43452-020-00130-1.Suche in Google Scholar

[15] Solomon N, Solomon I. Effect of deformation-induced phase transformation on AISI 316 stainless steel corrosion resistance. Eng Failure Anal. 2017;79:865–75. Elsevier Ltd, 10.1016/j.engfailanal.2017.05.031.Suche in Google Scholar

[16] Solomon N, Solomon I. Deformation induced martensite in AISI 316 stainless steel. Rev Met. Mar 2010;46(2):121–8. 10.3989/revmetalm.0920.Suche in Google Scholar

[17] Takaki S, Fukunaga K, Syarif J, Tsuchiyama T. Effect of grain refinement on thermal stability of metastable austenitic steel. Mater Trans. 2004;45(7):2245–51. 10.2320/matertrans.45.2245.Suche in Google Scholar

[18] Stalder M, Vogel S, Bourke MAM, Maldonado JG, Thoma DJ, Yuan VW. Retransformation (α′ → γ) kinetics of strain induced martensite in 304 stainless steel. Mater Sci Eng A. 2000;280:270–81.10.1016/S0921-5093(99)00706-6Suche in Google Scholar

[19] Normando PG, Moura EP, Souza JA, Tavares SSM, Padovese LR. Ultrasound, eddy current and magnetic Barkhausen noise as tools for sigma phase detection on a UNS S31803 duplex stainless steel. Mater Sci Eng A. May 2010;527(12):2886–91. 10.1016/j.msea.2010.01.017.Suche in Google Scholar

[20] Nalepka K, Skoczeń B, Ciepielowska M, Schmidt R, Tabin J, Schmidt E, et al. Phase transformation in 316l austenitic steel induced by fracture at cryogenic temperatures: experiment and modelling. Materials. 2021;14:127.10.3390/ma14010127Suche in Google Scholar PubMed PubMed Central

[21] Spencer K, Véron M, Yu-Zhang K, Embury JD. The strain induced martensite transformation in austenitic stainless steels: Part 1 – influence of temperature and strain history. Mater Sci Technol. 2009;25:7–17.10.1179/174328408X293603Suche in Google Scholar

[22] Tabin J, Nalepka J, Kawałko K, Brodecki A, Bała P, Kowalewski Z. Plastic flow instability in 304 austenitic stainless steels at room temperature. Metall Mater Trans A. 2023;54:4606–11.10.1007/s11661-023-07223-5Suche in Google Scholar

[23] Talonen J, Aspegren P, Hänninen H. Comparison of different methods for measuring strain induced α-martensite content in austenitic steels. Mater Sci Technol. 2004;20:1506–12.10.1179/026708304X4367Suche in Google Scholar

[24] Somani MC, Juntunen P, Karjalainen LP, Misra RDK, Kyröläinen A. Enhanced mechanical properties through reversion in metastable austenitic stainless steels. Metall Mater Trans A. 2009;40:729–44.10.1007/s11661-008-9723-ySuche in Google Scholar

[25] Glage A, Weidner A, Biermann H. Effect of austenite stability on the low cycle fatigue behavior and microstructure of high alloyed metastable austenitic cast TRIP steels. Procedia Eng. 2010;2:2085–94.10.1016/j.proeng.2010.03.224Suche in Google Scholar

[26] Domenjoud M, Daniel L. Effects of plastic strain and reloading stress on the magneto-mechanical behavior of electrical steels: experiments and modeling. Mech Mater. 2023;176:104510. 10.1016/j.mechmat.2022.104510.Suche in Google Scholar

Received: 2024-10-11
Revised: 2025-04-23
Accepted: 2025-04-27
Published Online: 2025-06-24

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Artikel in diesem Heft

  1. Research Article
  2. Modification of polymers to synthesize thermo-salt-resistant stabilizers of drilling fluids
  3. Study of the electronic stopping power of proton in different materials according to the Bohr and Bethe theories
  4. AI-driven UAV system for autonomous vehicle tracking and license plate recognition
  5. Enhancement of the output power of a small horizontal axis wind turbine based on the optimization approach
  6. Design of a vertically stacked double Luneburg lens-based beam-scanning antenna at 60 GHz
  7. Synergistic effect of nano-silica, steel slag, and waste glass on the microstructure, electrical resistivity, and strength of ultra-high-performance concrete
  8. Expert evaluation of attachments (caps) for orthopaedic equipment dedicated to pedestrian road users
  9. Performance and rheological characteristics of hot mix asphalt modified with melamine nanopowder polymer
  10. Second-order design of GNSS networks with different constraints using particle swarm optimization and genetic algorithms
  11. Impact of including a slab effect into a 2D RC frame on the seismic fragility assessment: A comparative study
  12. Analytical and numerical analysis of heat transfer from radial extended surface
  13. Comprehensive investigation of corrosion resistance of magnesium–titanium, aluminum, and aluminum–vanadium alloys in dilute electrolytes under zero-applied potential conditions
  14. Performance analysis of a novel design of an engine piston for a single cylinder
  15. Modeling performance of different sustainable self-compacting concrete pavement types utilizing various sample geometries
  16. The behavior of minors and road safety – case study of Poland
  17. The role of universities in efforts to increase the added value of recycled bucket tooth products through product design methods
  18. Adopting activated carbons on the PET depolymerization for purifying r-TPA
  19. Urban transportation challenges: Analysis and the mitigation strategies for road accidents, noise pollution and environmental impacts
  20. Enhancing the wear resistance and coefficient of friction of composite marine journal bearings utilizing nano-WC particles
  21. Sustainable bio-nanocomposite from lignocellulose nanofibers and HDPE for knee biomechanics: A tribological and mechanical properties study
  22. Effects of staggered transverse zigzag baffles and Al2O3–Cu hybrid nanofluid flow in a channel on thermofluid flow characteristics
  23. Mathematical modelling of Darcy–Forchheimer MHD Williamson nanofluid flow above a stretching/shrinking surface with slip conditions
  24. Energy efficiency and length modification of stilling basins with variable Baffle and chute block designs: A case study of the Fewa hydroelectric project
  25. Renewable-integrated power conversion architecture for urban heavy rail systems using bidirectional VSC and MPPT-controlled PV arrays as an auxiliary power source
  26. Review Articles
  27. A modified adhesion evaluation method between asphalt and aggregate based on a pull off test and image processing
  28. Architectural practice process and artificial intelligence – an evolving practice
  29. Special Issue: 51st KKBN - Part II
  30. The influence of storing mineral wool on its thermal conductivity in an open space
  31. Use of nondestructive test methods to determine the thickness and compressive strength of unilaterally accessible concrete components of building
  32. Use of modeling, BIM technology, and virtual reality in nondestructive testing and inventory, using the example of the Trzonolinowiec
  33. Tunable terahertz metasurface based on a modified Jerusalem cross for thin dielectric film evaluation
  34. Integration of SEM and acoustic emission methods in non-destructive evaluation of fiber–cement boards exposed to high temperatures
  35. Non-destructive method of characterizing nitrided layers in the 42CrMo4 steel using the amplitude-frequency technique of eddy currents
  36. Evaluation of braze welded joints using the ultrasonic method
  37. Analysis of the potential use of the passive magnetic method for detecting defects in welded joints made of X2CrNiMo17-12-2 steel
  38. Analysis of the possibility of applying a residual magnetic field for lack of fusion detection in welded joints of S235JR steel
  39. Eddy current methodology in the non-direct measurement of martensite during plastic deformation of SS316L
  40. Methodology for diagnosing hydraulic oil in production machines with the additional use of microfiltration
  41. Special Issue: IETAS 2024 - Part II
  42. Enhancing communication with elderly and stroke patients based on sign-gesture translation via audio-visual avatars
  43. Optimizing wireless charging for electric vehicles via a novel coil design and artificial intelligence techniques
  44. Evaluation of moisture damage for warm mix asphalt (WMA) containing reclaimed asphalt pavement (RAP)
  45. Comparative CFD case study on forced convection: Analysis of constant vs variable air properties in channel flow
  46. Evaluating sustainable indicators for urban street network: Al-Najaf network as a case study
  47. Node failure in self-organized sensor networks
  48. Comprehensive assessment of side friction impacts on urban traffic flow: A case study of Hilla City, Iraq
  49. Design a system to transfer alternating electric current using six channels of laser as an embedding and transmitting source
  50. Security and surveillance application in 3D modeling of a smart city: Kirkuk city as a case study
  51. Modified biochar derived from sewage sludge for purification of lead-contaminated water
  52. Special Issue: AESMT-7 - Part II
  53. Experimental study on behavior of hybrid columns by using SIFCON under eccentric load
Heruntergeladen am 8.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/eng-2025-0118/html
Button zum nach oben scrollen