Next Article in Journal
Synergistic Zn-Cd Bimetallic Engineering in ZIFs for High-Chloride 2e ORR to H2O2 in Simulated Neutral Seawater
Previous Article in Journal
Modification Mechanism of Glass Fibers on Ordinary Portland Cement and Sulphoaluminate Cement Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Cryogenic Treatment and Tempering Duration on the Microstructure and Mechanical Properties of Martensitic Stainless Steel 13Cr-2Ni-2Mo

Department of Materials Science and Engineering, National Taiwan University, No. 1, Sec. 4, Roosevelt Road, Taipei 10617, Taiwan
*
Author to whom correspondence should be addressed.
Materials 2025, 18(8), 1784; https://doi.org/10.3390/ma18081784
Submission received: 10 March 2025 / Revised: 31 March 2025 / Accepted: 1 April 2025 / Published: 14 April 2025

Abstract

:
Martensitic stainless steel (MSS) is widely used in several parts of automobiles where high strength, hardness, and corrosion resistance are required. However, the metastability of retained austenite can transform into martensite under severe deformation, adversely affecting material properties. Cryogenic treatments (CTs) have been extensively employed in iron-based alloys for fastener application due to their advantageous effect. This study explores the heat treatment processes applied to 13Cr-2Ni-2Mo martensitic stainless steel (MSS), including austenitizing, cryogenic treatment, and tempering cycles. Cryogenic treatment at (−150 °C) for varying durations, followed by tempering at 200 °C for 2 h, and the impact of post-cryogenic tempering at 200 °C for different tempering duration on the microstructure and mechanical properties were evaluated. Experimental results indicate that the sample quenched at 1040 °C for 2 h (CHT) contains lath martensite, retained austenite, δ-ferrite, and undissolved carbide precipitation. Compared to as-quenched samples, hardness decreased by 5.04%, 7.24%, and 7.32% after tempering for 2 h, 5 h, and 10 h, respectively. Extending cryogenic durations to 2 h, 12 h, and 20 h promoted nucleation of a mixture of M3C and M23C6 small globular carbides (SGCs) and grain refinement but resulted in hardness reductions of 5.04%, 5.32%, and 8.36%, respectively. The reduction in hardness is primarily attributed to a decrease in solid solution strengthening and promoted carbide coarsening.

1. Introduction

Martensitic stainless steels (MSSs) are extensively used in automobiles, turbine blades, surgical instruments, bearings, and fasteners due to their excellent strength, hardness, wear resistance, and corrosion resistance [1]. A recent advancement in automotive manufacturing involves the cold forming of high-strength steel and martensitic stainless steel components, reducing material waste and improving mechanical performance. This process allows the production of complex geometries with enhanced strength and ductility, making it suitable for critical applications such as chassis reinforcements and crash-resistant structures. Main properties such as formability must be optimized, alongside maintaining sufficient corrosion resistance and fatigue durability for extended service life [2].
To fulfill these requirements, microstructure and mechanical properties need to be precisely managed. In MSSs, the presence of certain phases, such as retained austenite and delta ferrite, during heat treatment and quenching may have an adverse impact on the mechanical properties. During the annealing of certain martensitic stainless steels, reverted austenite can achieve partial stabilization and persist at room temperature. This stabilization primarily arises from chemical factors, such as the partitioning of austenite-stabilizing elements, including its lamellar morphology. Retained austenite (RA) is metastable at room temperature and can transform under heavy load or stress during component operation. This transformation can lead to dimensional changes, negatively impacting the durability of tools and components during operation. Thus, it is crucial to eliminate RA from most stainless steels before use to ensure optimal functional performance [3,4,5]. The impact of δ-ferrite on martensitic-based steel remains unclear; however, the consensus among researchers is that its presence may deteriorate the mechanical properties. Specifically, δ-ferrite can adversely influence fatigue fracture as cracks are prone to nucleate in the regions containing δ-ferrite [6].
Cryogenic treatment (CT) is widely recognized for its effectiveness in reducing the fraction of retained austenite and stabilizing phases at room temperature. During CT, the material is held at a temperature significantly lower than the martensite finish (Mf) temperature for a specified duration, followed by a reheating process to room temperature. This treatment promotes the conversion of retained austenite into martensite and induces the formation of finely dispersed secondary carbide precipitates within nucleation sites [4,5,6,7,8,9,10,11].
This study focuses on the development of novel stainless steel 13Cr-2Ni-2Mo due to its superior mechanical properties compared to conventional martensitic stainless steel such as AISI 410 and AISI 420. This study will use martensitic stainless steel for experiments and design a series of heat treatments, including cryogenic treatments, to measure its hardness and observe microstructural differences. The objective is to understand the influence of cryogenic treatments and tempering duration towards the distribution of carbide and other phases.

2. Materials and Methods

2.1. Materials and Heat Treatments

In this study, experiments were conducted using a total of 7 samples of 13Cr-2Ni-2Mo martensitic stainless steel rods. The test rod used in this study was subjected to rod rolling and two annealing processes at above Ac1 temperature to relieve internal stresses, resulting in a final diameter of 3.5 mm for the rod used in this experiment. The composition of the 13Cr-2Ni-2Mo martensitic stainless steel rod was confirmed using a Spark Optical Emission Spectrometer (Spark OES) with four different points with the average elemental analysis results presented in Table 1.
For simplicity, this paper designates conventional heat treatment as CHT, which excludes cryogenic treatment, cryogenic treatment as CT, and tempering as T. The as-received material underwent solution treatment at 1040 °C for 2 h, followed by air-cooling as a conventional heat treatment (CHT) sample. The as-quenched sample was subsequently tempered at 200 °C for 2 h (CHT/T2) immediately. Both samples served as reference conditions for comparison against other experimental groups. To evaluate the effect of tempering duration on cryogenically treated CHT samples, various tempering times of 2 h, 5 h, and 10 h were applied after immersing the CHT sample in liquid nitrogen at −196 °C immediately after quenching. In addition, to study the influence of cryogenic treatment duration, samples were subjected to cryogenic treatment for 2, 12, and 20 h after CHT, followed by tempering at 200 °C for 2 h. The experimental design is illustrated in Figure 1 and Table 2.

2.2. Microstructure and Mechanical Properties Analysis

Samples were etched on glyceregia solution (3 parts of HCl, 2 parts of glycerol and 1 part of HNO3) for 1 min. A Leica DM2500M optical microscope (Leica Microsystems, Wetzlar, Germany) and a Canon EOD 5D MARK II camera (Future-Tech, Tokyo, Japan) were used for this experiment. To investigate microscale secondary phase particles, secondary electron images were captured at 20 kV accelerating voltage using a FEI NOVA scanning electron microscope. In this experiment, the FEI Tecnai G2 F20 200 kV FEG-TEM (Thermo Fisher Scientific, Hillsboro, OR, USA) was used for microstructure analysis. Bright-field imaging and diffraction patterns were employed to observe the morphology and analyze the crystal structure of martensite, austenite, and precipitates. The incorporated Energy-Dispersive Spectrometer (EDS) with an X-MAX 80 detector facilitated chemical composition analysis to observe the element of the precipitate. X-ray diffraction measurements were conducted using a D2 PHASER X-ray Powder Diffractometer (Bruker Corporation, Billerica, MA, USA). The XRD analysis utilized Cu-Kα radiation with a wavelength of 1.54060 Å with a 300 W high-voltage power supply. A JeoL JSM-7800F Prime field emission gun scanning electron microscope equipped (JEOL Ltd, Tokyo, Japan) with a NordlysNano EBSD detector (Oxford Instruments, Wycombe, UK) with a working voltage of 20 kV was used to receive backscattered electrons was used for the measurement of material crystal orientations, providing information such as microstructural composition, texture, orientation differences, phase proportions, grain size, and morphology.
XRD results were used to investigate the proportions of retained austenite and martensite. Proportional retained austenite was calculated using the following equation:
V γ = 1 V C 1 + R γ h k l R h k l I h k l I γ h k l
where Vγ is the volume percent of retained austenite, VC is the volume percent of carbide, which were calculated from SEM microstructure using ImageJ software (version 1.54p, National Institutes of Health (NIH), Bethesda, MD, USA), and Iα′ and Iγ are the integrated intensities measured for a single preselected martensite and austenite peak line, respectively. Rα′ and Rγ are theoretical intensity values for the same hkl planes [12,13]. Six diffraction peaks of (211)α′, (200)α′, (110)α′, (111)γ, (200)γ, and (220)γ were selected to measure the integrated intensity.
After applying conventional heat treatment and cryogenic treatment, hardness measurements were performed using an FM-810 Vickers hardness tester (Future-Tech, Tokyo, Japan) at a load of 1000 g and a dwell time of 10 s. Fifteen randomly selected points were tested on the material surface and the average hardness value. In this experiment, the MTS 810 universal testing machine (MTS Systems Corporation, Eden Prairie, MN, USA) was employed to conduct the three-point bend test at a downward speed of 0.48 mm/min. The three-point bending test was chosen over the tensile test to evaluate plastic deformation in martensitic stainless steel with minimal sample preparation. The testing method followed ASTM E290-22 [14] with a span of 15 mm and a sample length of 30 mm, which was calculated using the standard span formula as in the equation below:
C = 2 r + 3 t ± t 2
with C as the distance between lower supports, r as the radius of the pin end, t as the specimen thickness, d as the round specimen diameter, and was the specimen width [14]. The indenter and sample rod diameters are 4.90 mm and 3.50 mm, respectively. The applied force starts at zero and fracture marks the endpoint. The resulting force–displacement graph is used to assess the material tensile strength and ductility.

3. Results and Discussions

3.1. Microstructure of As-Quenched and Conventionally Quenched-Tempered Samples

The as-received microstructure contains distributed spherical carbide inside the grain and along the grain boundary, as shown in Figure 2a. The majority of carbides were redissolved following austenitization at 1040 °C for 2 h as illustrated in Figure 2b, indicating that this treatment parameter effectively facilitates the dissolution of a substantial quantity of alloying elements. Precipitation initiates along the grain boundaries after 2 h of tempering, as shown in Figure 2c, as these regions offer accelerated pathways for atomic diffusion compared to the inside grain due to their higher atomic disorder. Consequently, grain boundaries serve as favorable sites for carbon atom migration during the tempering process [4,15].
Optical microscope examination reveals that low amounts of δ-ferrite phase exist in sample CHT, which consists of a small discontinuous platelet-like shape with a size of around 1–50 μm, as shown in Figure 3 (red arrow). The cause is that ferrite-stabilizing elements like Cr and Mo diffuse and concentrate at elevated temperatures, leading to an increase in the Cr equivalent (Creq). This promotes the formation of delta ferrite more readily when austenitized at higher temperatures. As is well known, δ-ferrite is a relatively soft phase within martensite, which consequently decreases the overall strength of the steel [3,15].
The low tempering temperature, typically below 250 °C inhibits the widespread precipitation of carbides throughout the metal matrix and minimizes the formation of Cr-rich carbides in martensitic stainless steels. This material is designed for enhanced wear and corrosion resistance, which is crucial to carefully managing carbide formation, as an excessive presence of carbides could adversely affect the alloy’s corrosion resistance. Moreover, tempering at elevated temperatures ranging from 480 to 600 °C may result in reduced corrosion resistance owing to the onset of sensitization and reversed austenite formation [1,15,16,17].
During tempering, carbide precipitates begin to occur within the martensite matrix. These carbides tend to form at dislocations, grain boundaries, and interfaces within the martensitic microstructure. The precipitation of carbides may assist in refining the grain boundaries and substructures; thus, the grain size appears smaller in the tempered condition. The average prior austenite grain boundary (PAGB) size decreased to 42.54 ± 12.69 μm following conventional tempering, as illustrated in Figure 4d, compared to the as-quenched sample, which exhibited a PAGB size of 51.15 ± 10.17 μm (Figure 4c). KAM maps on both samples were observed to evaluate the degree of misorientation. In this paper, the threshold for general grain boundary misorientation is set at 5°. This implies that misorientations greater than 5° were excluded from the calculation of local misorientation, as they are attributed to grain boundaries rather than the accumulation of geometrically necessary dislocations (GNDs). The GND density can be extrapolated using the formula provided below [4,16,18]:
ρ G N D = 2 θ μ b
ρGND represents the geometrically necessary dislocation (GND) density at each point, θ denotes the local misorientation, μ refers to the unit length (which is 100 nm in this paper), and b is the Burgers vector. The CHT/T2 sample has a relatively higher average misorientation degree compared to CHT. The increase in misorientation degree during tempering can be attributed to a combination of transformation from retained austenite to secondary martensite while simultaneously carbon diffuses during tempering from martensite to another RA site which may introduce local changes in orientation, contributing to higher misorientation degrees. Therefore, according to Equation (3), the CHT/T2 sample exhibits a higher dislocation density compared to the as-quenched sample which contrasts with findings from other authors that indicate lower dislocation densities when tempering occurs at elevated temperatures (above 500 °C) due to stress relieve occurs simultaneously [16,17,18].
Figure 4. EBSD analysis IPF-Z images of (a) CHT and (b) CHT/T2; parent austenite grain boundary (PAGB) of (c) CHT and (d) CHT/T2; KAM maps of (e) CHT and (f) CHT/T2; average misorientation histogram of (g) CHT and (h) CHT/T2.
Figure 4. EBSD analysis IPF-Z images of (a) CHT and (b) CHT/T2; parent austenite grain boundary (PAGB) of (c) CHT and (d) CHT/T2; KAM maps of (e) CHT and (f) CHT/T2; average misorientation histogram of (g) CHT and (h) CHT/T2.
Materials 18 01784 g004
In this study, the relatively lower content of austenite-stabilizing elements, such as Ni, may result in an even lower Ms temperature. Additionally, fine carbides precipitate within the martensitic matrix and along grain boundaries during tempering. These precipitates can generate localized stress concentrations, increasing grain boundary misorientation [18,19,20,21,22,23,24,25]. Consequently, tempering at such a low temperature does not necessarily result in a lower misorientation degree compared to tempering at elevated temperatures [16].

3.2. Effects of Cryogenic Treatment

3.2.1. Microstructural Characterization

As illustrated in Figure 5, small globular carbides precipitated in the CHT sample after tempering and in samples subjected to cryogenic treatment, particularly along the grain boundaries. The carbides predominantly nucleate along the boundaries of martensitic laths and exhibit nanocluster growth within carbon-enriched regions across all samples, which is consistent with findings reported by other researchers [5,6,9]. As depicted in Figure 4a, the sample without cryogenic treatment exhibits the lowest carbide distribution.
Figure 6 shows the TEM morphologies of CHT/T2, which contain lath martensite matrix and retained austenite, while after cryogenic treatment CT2/T2, crystal defects such as twin martensite were observed. It shows that the size of martensite lath decreases significantly with the cryogenic treatment with an average width of approximately 417 nm (Figure 6b), while the conventional QT process exhibits a larger average lath martensite with a width of approximately 524 nm (Figure 6a) [19]. TEM analysis reveals that cryogenic treatment induces a higher dislocation density within the matrix, primarily attributed to the transformation of retained austenite into fresh martensite [26,27,28,29,30,31,32,33,34,35,36,37].
In addition, a smaller retained austenite width is observed in Figure 7 in the form of thin films with an approximate width of 100 nm located between martensite laths. Based on the identification of diffraction spots, these black elongated substances are retained austenite, which aligns with the K-S relationship with martensite, i.e., < 011 > γ F e < 111 > α F e and 111 γ F e 011 α F e . These retained austenite structures impede dislocation movement caused by external factors, thereby contributing to the improvement of hardness [8,23]. At cryogenic temperatures, C atoms diffuse from martensite to martensite/austenite interfaces due to lattice strain energy, hence enriching the retained austenite with C atoms. This C enrichment stabilizes the retained austenite during recovery to ambient temperature which prevents its transformation to martensite. This process is known as the thermal stabilization of austenite [3].
The influence of cryogenic treatment duration on the microstructure is depicted in Figure 8. As the treatment duration increases, the size of the lath martensite decreases, as shown in Figure 8a–c, and the prior austenite grain boundaries (PAGB) also become smaller, as shown in Figure 8d–f from 41.14 ± 9.22 μm, 27.58 ± 9.06 μm, and 24.84 ± 8.05 μm for sample CT2/T2, CT12/T2, and CT20/T2, respectively. This phenomenon can be attributed to two main factors. First, the transformation of retained austenite to martensite is enhanced with longer treatment durations, resulting in a refined martensitic structure. The newly formed martensite typically exhibits smaller sub-grains, which contributes to an overall reduction in grain size. Continuous cooling to cryogenic temperatures enhances martensite transformation energy and instability [3]. Second, extended cryogenic treatment promotes finer carbide precipitation along the grain boundaries. These precipitates can act as pinning forces at the grain boundaries, thereby restricting grain growth. The presence of carbides stabilizes the microstructure and inhibits coarsening of the grains [4,5,7,8,9,10,11].
The austenite-to-martensite transformation induces a positive volumetric effect, with the magnitude of this volumetric change increasing alongside the carbon content, which is consistently accompanied by plastic deformation in the newly formed martensite. Recent research has demonstrated that the martensitic transformation of retained austenite at low temperatures leads to plastic deformation of the newly formed martensite [3,4,5]. This deformation results in immobile carbon atoms being trapped by dislocations, forming carbon clusters that act as nucleation sites for finer carbide particles during tempering, thereby affecting the carbide distribution [5,9]. M. Villa et al. [32] studied the development of compressive strains in austenite following its transformation into martensite at low temperatures. This explains the observation of higher average misorientation degrees in the KAM results as the cryogenic treatment duration is extended [5]. Figure 8g,h,j show the local kernel average misorientation (KAM), which represents the strain concentration results for low angle boundary, and high KAM indicates a high-strain region. Following cryogenic treatment, the distribution of strain zones became more uniform. Additionally, regions of high strain were predominantly localized at the interfaces between the martensite and RA phases [4]. The KAM values of the martensite and RA phases increased after the cryogenic treatment from 0.375°, 1.693°, and 1.704° of sample CT2/T2, CT12/T2, and CT20/T2, respectively.
The findings, consistent with S. Li et al. [3], reveal a substantial number of carbides precipitated from the martensitic matrix, with the carbides being relatively uniformly distributed after a 20 h cryogenic treatment. Figure 9 illustrates the EDS spectra, where nanoscale and microscale small globular carbides (SGCs) enriched with Mo were identified. The presence of these carbides is a typical outcome observed following cryogenic treatment [5,36,37,38,39].
To examine the microstructure following cryogenic treatment, TEM analysis was performed on the CT2/T2 sample. Figure 10a shows the morphology and distribution of retained austenite and lath martensite, which include a small globular carbide. At lower tempering temperatures, the diffusion rate of chromium atoms remains slow, which leads to the preferential formation of a few instances of Fe-rich M3C (Figure 11d). However, since cryogenic treatment was applied prior to tempering, it facilitated nanoscale carbide nucleation within the matrix and followed by carbide stabilization after tempering [25,40,41,42,43,44,45]. As depicted in Figure 11, the carbides observed were predominantly a mixture of M23C6 with a size of approximately 500 nm and M3C. M23C6 SGC types were observed to be enriched with Cr and Mo, as shown in Figure 10b; this carbide most likely originated from undissolved carbide during austenitizing. The cryogenic treatment enhances the formation and stabilization of this nanoscale globular carbide M23C6 through diffusion as a reaction below [25,44,45]:
Matrix → M3C → M7C3 → M23C6

3.2.2. Mechanical Properties

As in Figure 12c, sample CHT exhibits the highest hardness among all samples (572 ± 4.99 HV). The primary strengthening mechanisms in tempered martensitic steel are influenced by fine grain strengthening, solid solution strengthening, dislocation strengthening, and precipitation strengthening. In the as-quenched sample, the highest hardness value can be attributed primarily to solid solution strengthening. Soluble atoms cause distortion in the Fe crystal lattice, creating a stress field that hinders dislocation movement due to the interaction between the stress field and the dislocations. Furthermore, since the as-quenched sample lacks carbide precipitation, precipitation strengthening plays a negligible role in its hardness and strength. Based on the KAM results, the misorientation degree, which correlates with dislocation density (and thus dislocation strengthening) also appears to have little effect on hardness [4,17,40]. The sample subjected to 2 h tempering (CHT/T2) demonstrates the lowest hardness, which is 527 ± 8.27 HV. This reduction in hardness can be attributed to the diffusion of alloying elements such as C, Cr, and Mo from the solid solution and carbide coarsening. These carbides precipitate immediately due to the higher fraction of retained austenite in conventional quenching and tempering (QT) treatments, facilitating more rapid carbon diffusion in the FCC phase. This process reduces solid solution strengthening and, as a result, causes a decrease in the overall hardness [16,41,42,43,44,45].
Cryogenic treatment before tempering was applied exhibited a higher hardness value of 543 ± 8.67 HV for the CT2/T2 sample, as observed in Figure 12c. Although there is a slight decrease in hardness for sample CT12/T2, this reduction was not significant. The combined effects of higher dislocation density, as illustrated in Figure 8l, precipitation strengthening, and a lower fraction of retained austenite, as shown in Figure 13, contribute substantially to the hardness improvement compared to conventional QT treatment. Conversely, in the sample subjected to 20 h of cryogenic treatment, the hardness deteriorates markedly due to the substantial amount of solute migrating out of the matrix to form carbides. Additionally, the chromium solute present in the martensite matrix and δ-ferrite may be rejected into the matrix, leading to the formation and coarsening of carbides at the interface between the δ-ferrite and the martensitic phase, as illustrated in Figure 9. This was supported by the rightward shift of the martensite peak on the XRD result observed in Figure 13 [7,8,9,26,36].
Figure 12a shows the mechanical properties of all samples from the three-point bend test. Surprisingly, the sample without cryogenic treatment (CHT/T2) has the highest maximum load, which is 2772 N compared with all samples. On the other hand, CHT reached only 1133 N before fracturing, primarily due to its brittleness. This phenomenon is attributed to carbon atoms being trapped within the martensitic structure, causing lattice distortion and significantly increasing brittleness. The material’s limited capacity for plastic deformation makes it highly prone to cracking under applied stress. The yield strength increments in stainless steels can be expressed using the following equation below:
σ y = σ 0 + σ s + σ p 2 + σ ρ 2 + k y d 1 / 2
In this equation, Δσ0 represents the friction stress for steel, often referred to as the Peierls–Nabarro force; Δσs accounts for solid solution hardening; Δσρ is associated with the strengthening effect of dislocations within the martensitic laths, and Δσp corresponds to precipitation hardening. The term kyd−1/2 describes the grain boundary strengthening, where ky is the Hall-Petch slope and d represents the effective grain size or the spacing of high-angle grain boundaries [25,40,46].
The CHT/T2 sample exhibited a relatively higher maximum load of 2772 N with the lowest vertical displacement at 9.1 mm, and fracture occurs during applied deformation compared to all cryogenic treatment samples, which demonstrated relatively better toughness without fracture with loads ranging from 1989 N (CT2/T5) to 2035 N (CT20/T2). Specifically, the maximum loads for the cryogenic treatment samples were recorded as follows: 2665 N (CT2/T2), 1989 N (CT2/T5), 2034 N (CT2/T10), 2001 N (CT12/T2) and 2035 N (CT20/T2). The primary reason stems from the tempering conditions referred to in Equation (5), which promote greater carbide precipitation along grain boundaries and induce more pronounced lattice distortion from dislocation accumulation during martensitic transformation and carbide formation, this process also stabilizes the retained austenite. The CHT/T2 sample, with its higher retained austenite content, demonstrates improved toughness despite lower hardness. This behavior can be attributed to the reduced solid solution strengthening effect, as confirmed by XRD results, which allows the sample to sustain a higher maximum load before fracture even with its lower hardness [47,48,49,50,51,52,53].
It is widely recognized that the aging or tempering of as-hardened steel can stabilize retained austenite (RA). During the aging or tempering process, carbon atoms initially diffuse from the martensite (M) phase to the M/RA interfaces, where they segregate and anchor the typically mobile dislocations. In the later stages, the carbon atoms begin to diffuse into the austenite phase, leading to an increase in the carbon content within the austenite [47,48,49,50,51].
In terms of quantifying residual austenite, analyses were conducted on the austenitized specimens using XRD, as shown in Figure 13, retained austenite (RA) volume fraction revealed a decreasing trend after cryogenic treatment, reducing from 9.53% in the CHT sample to 8.43% after 2 h of immersion (CT2/T2) and further to 6.51% after 20 h of immersion (CT20/T2). This downward trend is consistent with the findings of Li S. et al. [3] for stainless steel and Kang C. et al. [4] for steel in their respective experiments, which conclude that cryogenic treatment is an effective way to reduce retained austenite.

3.2.3. Microstructural Evolution with Different Tempering Duration

Cryogenic treatment is an effective way to reduce retained austenite to an acceptable level. In addition to the retained austenite reduction, a significant increased amount of additional small globular carbides in the microstructure of the cryogenically treated material [7,8,9,10,11].
A longer tempering duration exhibits a decreasing trend in Cr content as the tempering time increases, as reported by Jiang W. et al. [1]. This phenomenon can be attributed to carbide precipitation which is evident from the SEM analyses. XRD result Figure 14 illustrates the martensite peak is the predominant peak, with a smaller intensity peak corresponding to retained austenite clearly visible. However, distinguishing between the martensite and body-centered cubic (bcc) δ-ferrite peaks is challenging due to their peak overlap. The martensite peak for sample CHT/T2 displays a significant rightward shift compared to sample CHT, which indicates that conventional QT treatment develops Cr-depleted regions and a reduction in the martensite lattice parameter. This shift results from the depletion of C and Cr in the martensite matrix as it forms a solid solution [1,17,23].
RA content was reduced for the sample after 10 h tempering, which was 6.04% compared to 2 h tempering at 8.43%. The main reasons were as follows: first, carbon diffusion during prolonged tempering, carbon atoms diffuse more significantly from the retained austenite into the martensitic matrix and decrease the stability of the austenite. This depletion of carbon facilitates the transformation of austenite into martensite. Second, carbide precipitation and extended tempering increase the precipitation of carbides, particularly M23C6, which further depletes carbon from the austenite, reducing its stability and encouraging the transformation of austenite into martensite [17,47,48,49,50,51]. As shown in Figure 15, an increased amount of carbide (red arrow) precipitation was evident, particularly in sample CT2/T10, which likely contains Mo- and Cr-rich carbides within the martensitic matrix.
The hardness decreases with extended tempering durations from 543 ± 8.67 HV, 531 ± 8.73 HV, and 530 ± 7.49 HV for samples CT2/T2, CT2/T5, and CT2/T10, respectively. The hardness observed after tempering was influenced not only by the dissolution of carbon atoms but also by the recovery of the martensitic structure at a low tempering temperature of 200 °C. As tempering progresses, the size and quantity of carbides tend to increase [42,43,44,45]. Larger PAGB were observed as well as extended tempering duration refer to Figure 16d–f from 41.14 ± 9.22 μm, 43.24 ± 12.24 μm, and 55.73 ± 14.08 μm for CT2/T2, CT2/T5, and CT2/T10, respectively. A longer tempering duration can lead to a decrease in hardness. This was attributed to the generation of non-uniformly distributed coarser particles, making them less effective in acting as pinning points, thus reducing the overall hardness [5].

4. Conclusions

In this study, it was observed that both cryogenic treatment and tempering duration influence microstructure and mechanical properties with the following results:
  • The as-quenched sample (CHT) of 13Cr-2Ni-2Mo contains lath martensite, retained austenite, undissolved carbide, and δ-ferrite. Cryogenic treatment (CT) facilitates the conversion of retained austenite into martensite, resulting in increased hardness compared to conventional QT treatment (CHT/T2).
  • The predominant carbide types are nanosized M23C6 and a few instances of M3C (cementite), along with a smaller grain size distribution relative to the CHT process. Following cryogenic treatment durations of 2 h, 12 h, and 20 h, the average prior austenite grain boundary (PAGB) sizes were found to be 41.14 ± 9.22 μm, 27.58 ± 9.06 μm, and 24.84 ± 8.05 μm, respectively.
  • Variations in tempering duration can lead to an increase in PAGB size, which is associated with the coarsening of carbides and grain size. After tempering for 2 h, 5 h, and 10 h at 200 °C, the PAGB sizes were 41.14 ± 9.22 μm, 43.24 ± 12.24 μm, and 55.73 ± 14.08 μm, respectively.
  • Extended durations of cryogenic treatment can result in decreased hardness compared to the as-quenched sample, primarily due to a reduction in the solid solution strengthening effect. After cryogenic treatment for 2 h, 12 h, and 20 h followed by tempering, the hardness values relative to the as-quenched state were 543 ± 8.67 HV (−5.04%), 541 ± 6.12 HV (−5.32%), and 524 ± 6.53 HV (−8.36%), respectively.
  • The hardness of samples subjected to different tempering durations post-cryogenic treatment were 2 h, 5 h, and 10 h at 200 °C, compared to the as-quenched sample, yielded values of 543 ± 8.67 HV (−5.04%), 531 ± 8.73 HV (−7.24%), and 530 HV ± 7.49 (−7.32%), respectively.
  • 13Cr-2Ni-2Mo martensitic stainless steel, which is intended for automobile applications requiring high hardness and good toughness, and a combination of cryogenic treatment for 2 h at −150 °C, followed by tempering for 2 h at 200 °C (CT2/T2), was the optimal processing route in this study. Cryogenic treatment contributes to a reduction in retained austenite content effectively, which prevents dimensional changes caused by austenite transformation during service, thereby avoiding failure and effectively improving the service life of the experimental steel.

Author Contributions

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

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

The authors would like to express their gratitude to the instrumental assistance provided by the Instrumentation Center of the National Taiwan University for the EBSD experiment and the National Taiwan University of Science and Technology for XRD measurement.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jiang, W.; Wu, D.; Zhang, Q.; Li, M.; Liu, W. Effect of Tempering Time on the Microstructure and Properties of Martensitic Stainless Steel. Metals 2014, 14, 322. [Google Scholar] [CrossRef]
  2. de Pablos, J.L.; Sierra-Soraluce, A.; Sabirov, I.; Muratori, M.; Smith, A. Assessing the Feasibility of Cold Forming of Automotive Parts from Quenched and Partitioned Martensitic Stainless Steels. Steel Res. Int. 2024, 95, 2300280. [Google Scholar] [CrossRef]
  3. Li, S.; Xiao, M.; Ye, G.; Zhao, K.; Yang, M. Effects of deep cryogenic treatment on microstructural evolution and alloy phases precipitation of a new low carbon martensitic stainless bearing steel during aging. Mater. Sci. Eng. A 2018, 732, 167–177. [Google Scholar] [CrossRef]
  4. Kang, C.; Liu, F.; Jiang, Z.; Suo, H.; Yu, X.; Zhang, H.; Ding, S. Effect of cryogenic treatment on microstructure evolution and mechanical properties of high nitrogen plastic die steel. J. Mater. Res. Technol. 2021, 15, 5128–5140. [Google Scholar] [CrossRef]
  5. Jurči, P.; Dlouhý, I. Cryogenic Treatment of Martensitic Steels: Microstructural Fundamentals and Implications for Mechanical Properties and Wear and Corrosion Performance. Materials 2014, 17, 548. [Google Scholar] [CrossRef]
  6. Wang, P.; Lu, S.P.; Xiao, N.M.; Li, D.Z.; Li, Y.Y. Effect of delta ferrite on impact properties of low carbon 13Cr-4Ni martensitic stainless steel. Mater. Sci. Eng. A 2010, 527, 3210–3216. [Google Scholar] [CrossRef]
  7. Klug, P.J. Mechanisms and Effect of Deep Cryogenic Treatment on Steel Properties. Ph.D. Thesis, Jožef Stefan International Postgraduate School, Ljubljana, Slovenia, 2022. [Google Scholar]
  8. Li, D.; Li, Z.; Xiao, M.; Li, S.; Zhao, K.; Yang, M. Effect of Deep Cryogenic Treatment on Mechanical Property and Microstructure of a Low Carbon High Alloy Martensitic Bearing Steel during Tempering. Chin. J. Mater. Res. 2019, 33, 561–571. [Google Scholar] [CrossRef]
  9. Jovičević-Klug, P.; Jovičević-Klug, M.; Sever, T.; Feizpour, D.; Podgornik, B. Impact of steel type, composition and heat treatment parameters on effectiveness of deep cryogenic treatment. J. Mater. Res. Technol. 2021, 14, 1007–1020. [Google Scholar] [CrossRef]
  10. Xu, G.; Huang, P.; Feng, Z.; Wei, Z.; Zu, G. Effect of Deep Cryogenic Time on Martensite Multi-Level Microstructures and Mechanical Properties in AISI M35 High-Speed Steel. Materials 2022, 15, 6618. [Google Scholar] [CrossRef]
  11. Singh, G.; Pandey, K.N. Effect of cryogenic treatment on properties of materials: A review. Proc. Inst. Mech. Eng. Part E-J. Process Mech. Eng. 2022, 236, 1758–1773. [Google Scholar] [CrossRef]
  12. Jatczak, C.F. Retained Austenite and Its Measurement by X-Ray Diffraction. In Proceedings of the Automotive Engineering Congress and Exposition, Detroit, OH, USA, 25–29 February 1980, ISSN 0148-7191. [Google Scholar]
  13. He, J.; Wang, Y.; Qiu, X.; Yang, B.; Gu, J.; Chi, H.; Cheng, X. Effect of tempering time on microstructure and mechanical properties of a low carbon stainless bearing steel. Mater. Today Commun. 2025, 42, 111305. [Google Scholar] [CrossRef]
  14. ASTM E290-22; Standard Test Methods for Bend Testing of Materials for Ductility. ASTM International: West Conshohocken, PA, USA, 2022.
  15. Li, S.; Yuan, X.; Jiang, W.; Sun, H.; Li, J.; Zhao, K.; Yang, M. Effects of heat treatment influencing factors on microstructure and mechanical properties of a low-carbon martensitic stainless bearing steel. Mater. Sci. Eng. A 2014, 605, 229–235. [Google Scholar] [CrossRef]
  16. Wang, Y.; Zhang, X.; Jiang, L.; Yuan, C.; Zhang, J.; Yan, Q. Advancement of strength and toughness in ultra-low carbon martensitic stainless steel by reversed austenite. Nucl. Mater. Energy 2024, 38, 101601. [Google Scholar] [CrossRef]
  17. Chen, H.; Zeng, T.; Shi, Q.; Wang, N.; Zhang, S.; Yang, K.; Yan, W.; Wei, W. Microstructure evolution and mechanical properties during long-term tempering of a low carbon martensitic stainless bearing steel. J. Mater. Res. Technol. 2023, 25, 297–309. [Google Scholar] [CrossRef]
  18. Zhang, X.; Li, J.; Gu, J.; Liao, L.; Deng, Y. Effect of nitrogen and tempering temperature on microstructure evolution and mechanical properties of 0Cr15Ni6Mo2 martensitic stainless steel. Ironmak. Steelmak. 2022, 49, 311–321. [Google Scholar] [CrossRef]
  19. Fan, S.; Hao, H.; Meng, L.; Zhang, X. Effect of deep cryogenic treatment parameters on martensite multi-level microstructures and properties in a lath martensite/ferrite dual-phase steel. Mater. Sci. Eng. A 2021, 810, 141022. [Google Scholar] [CrossRef]
  20. Abid, A.A.H.; Muhammad, H.B. Effect of Cryogenic Treatment on the Tensile Properties of Carbon Dual Phase Steel. J. Eng. 2013, 19, 574–582. [Google Scholar] [CrossRef]
  21. Yan, X.G.; Li, D.Y. Effects of the sub-zero treatment condition on microstructure, mechanical behavior and wear resistance of W9Mo3Cr4V high speed steel. Wear 2013, 302, 854–862. [Google Scholar] [CrossRef]
  22. Savyasachi, N.; Reji, R.; Sajan, J.A.; Rafi, A.M. A Review on the Cryogenic Treatment of Stainless Steels, Tool Steels, and Carburized Steels. Int. J. Innov. Sci. Res. Technol. 2020, 5, 31–38. [Google Scholar]
  23. Li, D.H.; He, W.C.; Zhang, X.; Xiao, M.G.; Li, S.H.; Zhao, K.Y.; Yang, M.S. Effects of traditional heat treatment and a novel deep cryogenic treatment on microstructure and mechanical properties of low-carbon high-alloy martensitic bearing steel. J. Iron Steel Res. Int. 2021, 28, 370–382. [Google Scholar] [CrossRef]
  24. Villa, M.; Somers, M. Cryogenic Treatment of Steel: From Concept to Metallurgical Understanding. In Proceedings of the 24th International Feration for Heat Treatment and Surface Engineering Congress, Nice, France, 26–29 June 2017; pp. 26–29. [Google Scholar]
  25. Yang, Z.; Liu, Z.; Liang, J.; Yang, Z.; Sheng, G. Elucidating the role of secondary cryogenic treatment on mechanical properties of a martensitic ultra-high strength stainless steel. Mater. Charact. 2021, 178, 111277. [Google Scholar] [CrossRef]
  26. Prieto, G.; Ipiña, J.P.; Tuckart, W.R. Cryogenic treatments on AISI 420 stainless steel: Microstructure and mechanical properties. Mater. Sci. Eng. A 2014, 605, 236–243. [Google Scholar] [CrossRef]
  27. Prieto, G.; Tuckart, W.R.; Perez Ipiña, J.E. Influence of a Cryogenic Treatment on the Fracture Toughness of an AISI 420 Martensitic Stainless Steel. Mater. Technol. 2017, 51, 591–596. [Google Scholar] [CrossRef]
  28. Peng, J.; Zhou, B.; Li, Z.; Huo, D.; Xiong, J.; Zhang, S. Effect of tempering process on the cryogenic impact toughness of 13Cr4NiMo martensitic stainless steel. J. Mater. Res. Technol. 2023, 23, 5618–5630. [Google Scholar] [CrossRef]
  29. Antony, A.; Schmerl, N.M.; Sokolova, A.; Mahjoub, R.; Fabijanic, D.; Stanford, N.E. Quantification of the Dislocation Density, Size, and Volume Fraction of Precipitates in Deep Cryogenically Treated Martensitic Steels. Metals 2020, 10, 1561. [Google Scholar] [CrossRef]
  30. Cai, X.; Hu, X.; Zheng, L.; Li, D. Redistribution of C and N Atoms in High Nitrogen Martensitic Stainless Steel During Cryogenic Treatment. Acta Metall. Sin. Engl. Lett. 2022, 35, 591–595. [Google Scholar] [CrossRef]
  31. Singh, J.; Nath, S.K. Effects of Cyclic Heat Treatment on Microstructure and Mechanical Properties of 13%Cr-4%Ni Martensitic Stainless Steel. J. Mater. Eng. Perform. 2020, 29, 2478–2490. [Google Scholar] [CrossRef]
  32. Villa, M.; Pantleon, K.; Somers, M.A. Evolution of compressive strains in retained austenite during sub-zero Celsius martensite formation and tempering. Acta Mater. 2014, 65, 383–392. [Google Scholar] [CrossRef]
  33. Yildiz, E.; Özbek, N. Effect of cryogenic treatment and tempering temperature on mechanical and microstructural properties of AISI 431 steel. Int. J. 3D Print. Technol. Digit. Ind. 2022, 6, 74–82. [Google Scholar] [CrossRef]
  34. Sola, R.; Giovanardi, R.; Parigi, G.; Veronesi, P. A Novel Method for Fracture Toughness Evaluation of Tool Steels with Post-Tempering Cryogenic Treatment. Metals 2017, 7, 75. [Google Scholar] [CrossRef]
  35. Idayan, A.; Gnanavelbabu, A.; Rajkumar, K. Influence of Deep Cryogenic Treatment on the Mechanical Properties of AISI 440C Bearing Steel. Procedia Eng. 2014, 97, 1683–1691. [Google Scholar] [CrossRef]
  36. Das, D.; Dutta, A.K.; Ray, K.K. Sub-zero treatments of AISI D2 steel: Part I. Microstructure and hardness. Mater. Sci. Eng. A 2010, 527, 2182–2193. [Google Scholar] [CrossRef]
  37. Tyshchenko, A.I.; Theisen, W.; Oppenkowski, A.; Siebert, S.; Razumov, O.N.; Skoblik, A.P.; Sirosh, V.A.; Petrov, Y.N.; Gavriljuk, V.G. Low-temperature martensitic transformation and deep cryogenic treatment of a tool steel. Mater. Sci. Eng. A 2010, 527, 7027–7039. [Google Scholar] [CrossRef]
  38. Essam, M.A.; Shash, A.Y.; El-Fawakhry, M.K.; El-Kashif, E.; Megahed, H. Effect of Deep Cryogenic Treatment on Wear Behavior of Cold Work Tool Steel. Metals 2023, 13, 382. [Google Scholar] [CrossRef]
  39. Özbek, N.A.; Özbek, O. Investigation of The Effects of Shallow Cryogenic Treatment on the Mechanical and Microstructural Properties of 1.2436 Tool Steel. J. Mater. Mechatron. A 2022, 3, 151–162. [Google Scholar] [CrossRef]
  40. Zheng, K.; Zhong, Z.; Wang, H.; Xu, H.; Yu, F.; Wang, C.; Wu, G.; Liang, J.; Godfrey, A.; Cao, W. Obtaining Excellent Mechanical Properties in an Ultrahigh-Strength Stainless Bearing Steel via Solution Treatment. Metals 2023, 13, 1824. [Google Scholar] [CrossRef]
  41. Wang, F.Y.; Jin, J.J.; Jiang, Z.H.; Wang, X.Z.; Hu, C.W. Effect of Heat Treatment Temperature on Microstructure and Properties of New Maraging Stainless Steel. J. Mater. Eng. 2019, 47, 152–160. [Google Scholar] [CrossRef]
  42. Zhang, L.; Wang, P.; Zhang, Y.; Liu, G. Effect of Normalizing Temperature on High Strength of 13Cr4Ni Martensitic Stainless Steel. Metall. Eng. 2022, 9, 321–327. [Google Scholar] [CrossRef]
  43. Jiang, W.; Ye, D.; Li, J.; Su, J.; Zhao, K. Reverse Transformation Mechanism of Martensite to Austenite in 00Cr15Ni7Mo2WCu2 Super Martensitic Stainless Steel. Steel Res. Int. 2014, 85, 1150–1157. [Google Scholar] [CrossRef]
  44. Hsiao, P.C. The Effect of Heat Treatment on Microstructure, Mechanical Properties, and Corrosion Resistance of 13Cr-2Ni-2Mo Martensitic Stainless Steel. Master’s Thesis, Material Science and Engineering, National Taiwan University, Taipei, Taiwan, 2023. [Google Scholar]
  45. Wang, R.; Li, F.; Wu, Z.; Kang, Y.; Fan, J.; Yu, Z.; Yan, Y.; Du, S.; Eckert, J. Precipitation and Transformation of Carbides during Tempering of 7Cr14 Martensitic Stainless Steel. Steel Res. Int. 2023, 95, 2300248. [Google Scholar] [CrossRef]
  46. Yuan, X.H.; Yang, M.S.; Zhao, K.Y. Effects of Microstructure Transformation on Strengthening and Toughening for Heat-Treated Low Carbon Martensite Stainless Bearing Steel. Mater. Sci. Forum 2015, 817, 667–674. [Google Scholar] [CrossRef]
  47. Burja, J.; Šuler, B.; Češnjaj, M.; Nagode, A. Effect of Intercritical Annealing on the Microstructure and Mechanical Properties of 0.1C-13Cr-3Ni Martensitic Stainless Steel. Metals 2021, 11, 392. [Google Scholar] [CrossRef]
  48. Chenna Krishna, S.; Srinath, J.; Jha, A.K.; Pant, B.; Sharma, S.C.; George, K.M. Effect of Heat Treatment on Microstructure and Mechanical Properties of 12Cr–10Ni–0.25Ti–0.7Mo Stainless Steel. Metallogr. Microstruct. Anal. 2013, 2, 234–241. [Google Scholar] [CrossRef]
  49. Gao, Q.; Zhang, Y.; Zhang, H.; Li, H.; Qu, F.; Han, J.; Cheng, L.; Wu, B.; Lu, Y.; Ma, Y. Precipitates and Particles Coarsening of 9Cr-1.7W-0.4Mo-Co Ferritic Heat-Resistant Steel after Isothermal Aging. Sci. Rep. 2017, 7, 5859. [Google Scholar] [CrossRef]
  50. Wang, L.; Dong, C.; Yu, Q.; Man, C.; Hu, Y.; Dai, Z.; Li, X. The Correlation Between the Distribution/Size of Carbides and Electrochemical Behavior of 17Cr-1Ni Ferritic-Martensitic Stainless Steel. Metall. Mater. Trans. A 2019, 50, 388–400. [Google Scholar] [CrossRef]
  51. Fedoseeva, A.; Dolzhenko, A.; Kaibyshev, R. Thermo-Mechanical Processing as Method Decreasing Delta-Ferrite and Improving the Impact Toughness of the Novel 12% Cr Steels with Low N and High B Contents. Materials 2022, 15, 8861. [Google Scholar] [CrossRef]
  52. Li, J.; Wang, L.; Wang, H.; Zhang, P.; Guo, F.; Zhang, X. Dissolution Behavior of Delta Ferrites in Martensitic Heat-resistant Steel for Ultra Supercritical Units Blades. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2022, 37, 730–734. [Google Scholar] [CrossRef]
  53. Chen, Q.Y.; Ren, J.K.; Xie, Z.L.; Zhang, W.N.; Chen, J.; Liu, Z.Y. Correlation between reversed austenite and mechanical properties in a low Ni steel treated by ultra-fast cooling, intercritical quenching and tempering. J. Mater. Sci. 2020, 55, 1840–1853. [Google Scholar] [CrossRef]
Figure 1. Heat treatment process for (a) different cryogenic duration; (b) different tempering duration (note: cryogenic treatment in liquid nitrogen was performed immediately after air-cooling (A.C.) quenching, followed immediately by tempering).
Figure 1. Heat treatment process for (a) different cryogenic duration; (b) different tempering duration (note: cryogenic treatment in liquid nitrogen was performed immediately after air-cooling (A.C.) quenching, followed immediately by tempering).
Materials 18 01784 g001
Figure 2. SEM microstructure of sample: (a) as-received sample; (b) CHT; (c) CHT/T2 (red arrow: carbide precipitation).
Figure 2. SEM microstructure of sample: (a) as-received sample; (b) CHT; (c) CHT/T2 (red arrow: carbide precipitation).
Materials 18 01784 g002
Figure 3. Optical micrographs of sample CHT (dashed red arrow: δ-Fe as a platelet-like structure; solid red arrow: δ-Fe as isolated islands).
Figure 3. Optical micrographs of sample CHT (dashed red arrow: δ-Fe as a platelet-like structure; solid red arrow: δ-Fe as isolated islands).
Materials 18 01784 g003
Figure 5. FESEM microstructure of sample: (a) CHT/T2 without cryogenic treatment; (b) CT2/T2; (c) CT12/T2; (d) CT20/T2 (solid red arrow: carbide; dashed red arrow: tempered martensite).
Figure 5. FESEM microstructure of sample: (a) CHT/T2 without cryogenic treatment; (b) CT2/T2; (c) CT12/T2; (d) CT20/T2 (solid red arrow: carbide; dashed red arrow: tempered martensite).
Materials 18 01784 g005
Figure 6. TEM micrographs of sample: (a) CHT/T2 without cryogenic treatment; (b) with cryogenic treatment CT2/T2 (T: twin martensite).
Figure 6. TEM micrographs of sample: (a) CHT/T2 without cryogenic treatment; (b) with cryogenic treatment CT2/T2 (T: twin martensite).
Materials 18 01784 g006
Figure 7. TEM micrograph of sample CT2/T2 with cryogenic treatment: (a) bright field image; (b) dark field image; (c) SAED pattern of retained austenite (RA).
Figure 7. TEM micrograph of sample CT2/T2 with cryogenic treatment: (a) bright field image; (b) dark field image; (c) SAED pattern of retained austenite (RA).
Materials 18 01784 g007
Figure 8. EBSD analysis IPF-Z images of (a) CT2/T2, (b) CT12/T2, and (c) CT20/T2; parent austenite grain boundary (PAGB) of (d) CT2/T2, (e) CT12/T2, and (f) CT20/T2; KAM maps of (g) CT2/T2, (h) CT12/T2, and (i) CT20/T2; and average misorientation histogram of (j) CT2/T2, (k) CT12/T2, and (l) CT20/T2.
Figure 8. EBSD analysis IPF-Z images of (a) CT2/T2, (b) CT12/T2, and (c) CT20/T2; parent austenite grain boundary (PAGB) of (d) CT2/T2, (e) CT12/T2, and (f) CT20/T2; KAM maps of (g) CT2/T2, (h) CT12/T2, and (i) CT20/T2; and average misorientation histogram of (j) CT2/T2, (k) CT12/T2, and (l) CT20/T2.
Materials 18 01784 g008
Figure 9. EDS spectra of sample after cryogenic for 20 h and tempering for 2 h (CT20/T2). Point 1: matrix; point 2: Mo-rich carbide.
Figure 9. EDS spectra of sample after cryogenic for 20 h and tempering for 2 h (CT20/T2). Point 1: matrix; point 2: Mo-rich carbide.
Materials 18 01784 g009
Figure 10. (a) TEM micrograph of CT2/T2; (b) TEM-EDS point label (T: twin martensite).
Figure 10. (a) TEM micrograph of CT2/T2; (b) TEM-EDS point label (T: twin martensite).
Materials 18 01784 g010
Figure 11. TEM image results after CT at −150 °C and tempering at 200 °C (CT2/T2): (a) bright field image; (b) dark field image; (c) SAED for M23C6; (d) TEM micrograph shows few instances of needle-shape M3C (red arrow).
Figure 11. TEM image results after CT at −150 °C and tempering at 200 °C (CT2/T2): (a) bright field image; (b) dark field image; (c) SAED for M23C6; (d) TEM micrograph shows few instances of needle-shape M3C (red arrow).
Materials 18 01784 g011
Figure 12. (a) Three-point bending test results at different heat treatments; (b) larger picture for (a); (c) maximum load, hardness value, and grain size all samples.
Figure 12. (a) Three-point bending test results at different heat treatments; (b) larger picture for (a); (c) maximum load, hardness value, and grain size all samples.
Materials 18 01784 g012
Figure 13. Fraction of RA and carbide.
Figure 13. Fraction of RA and carbide.
Materials 18 01784 g013
Figure 14. XRD spectra for CHT and CT specimens.
Figure 14. XRD spectra for CHT and CT specimens.
Materials 18 01784 g014
Figure 15. FESEM microstructure of sample: (a) with cryogenic treatment CT/T2; (b) CT2/T5; (c) CT2/T10.
Figure 15. FESEM microstructure of sample: (a) with cryogenic treatment CT/T2; (b) CT2/T5; (c) CT2/T10.
Materials 18 01784 g015
Figure 16. EBSD analysis IPF-Z images of (a) CT2/T2, (b) CT2/T5, and (c) CT2/T10; parent grain boundary (PAGB) of (d) CT2/T2, (e) CT2/T5, and (f) CT2/T10; KAM maps of (g) CT2/T2, (h) CT2/T5, and (i) CT2/T10; average misorientation histogram of (j) CT2/T2, (k) CT2/T5, and (l) CT2/T10.
Figure 16. EBSD analysis IPF-Z images of (a) CT2/T2, (b) CT2/T5, and (c) CT2/T10; parent grain boundary (PAGB) of (d) CT2/T2, (e) CT2/T5, and (f) CT2/T10; KAM maps of (g) CT2/T2, (h) CT2/T5, and (i) CT2/T10; average misorientation histogram of (j) CT2/T2, (k) CT2/T5, and (l) CT2/T10.
Materials 18 01784 g016
Table 1. Chemical composition of 13Cr-2Ni-2Mo in wt%.
Table 1. Chemical composition of 13Cr-2Ni-2Mo in wt%.
MSSCCrNiMoSiMnNFe
13Cr-2Ni-2Mo0.17 ± 0.0312.9 ± 0.051.8 ± 0.071.92 ± 0.060.36 ± 0.020.08 ± 0.010.12 ± 0.04Bal.
Table 2. Specimen identification.
Table 2. Specimen identification.
IdentificationDescription
As-receivedPrior heat treatment
CHT Conventionally heat-treated (as-quenched)
CHT/T2As-quenched + Tempering at 200 °C for 2 h
CT2/T2As-quenched + cryogenically at −150 °C for 2 h + Tempering at 200 °C for 2 h
CT2/T5As-quenched + cryogenically at −150 °C for 2 h + Tempering at 200 °C for 5 h
CT2/T10As-quenched + cryogenically at −150 °C for 2 h + Tempering at 200 °C for 10 h
CT12/T2As-quenched + cryogenically at −150 °C for 12 h + Tempering at 200 °C for 2 h
CT20/T2As-quenched + cryogenically at −150 °C for 20 h + Tempering at 200 °C for 2 h
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

Fatih, M.R.R.; Chen, H.-J.; Lin, H.-C. The Effect of Cryogenic Treatment and Tempering Duration on the Microstructure and Mechanical Properties of Martensitic Stainless Steel 13Cr-2Ni-2Mo. Materials 2025, 18, 1784. https://doi.org/10.3390/ma18081784

AMA Style

Fatih MRR, Chen H-J, Lin H-C. The Effect of Cryogenic Treatment and Tempering Duration on the Microstructure and Mechanical Properties of Martensitic Stainless Steel 13Cr-2Ni-2Mo. Materials. 2025; 18(8):1784. https://doi.org/10.3390/ma18081784

Chicago/Turabian Style

Fatih, Muhammad R. R., Hou-Jen Chen, and Hsin-Chih Lin. 2025. "The Effect of Cryogenic Treatment and Tempering Duration on the Microstructure and Mechanical Properties of Martensitic Stainless Steel 13Cr-2Ni-2Mo" Materials 18, no. 8: 1784. https://doi.org/10.3390/ma18081784

APA Style

Fatih, M. R. R., Chen, H.-J., & Lin, H.-C. (2025). The Effect of Cryogenic Treatment and Tempering Duration on the Microstructure and Mechanical Properties of Martensitic Stainless Steel 13Cr-2Ni-2Mo. Materials, 18(8), 1784. https://doi.org/10.3390/ma18081784

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