Next Article in Journal
Multi-Objective Optimization Inverse Analysis for Characterization of Petroleum Geomechanical Properties During Hydraulic Fracturing
Previous Article in Journal
Correction: Cai et al. Acoustic Characterization of Leakage in Buried Natural Gas Pipelines. Processes 2025, 13, 2274
Previous Article in Special Issue
Influence of Additional Devices and Polymeric Matrix on In Situ Welding in Material Extrusion: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tensile and Fracture Properties Evaluation of Additively Manufactured Different Stainless Steels via Small Punch Testing

1
State Key Laboratory of Intelligent Coal Mining and Strata Control, Beijing 100013, China
2
Graduate School, China Coal Research Institute, Beijing 100013, China
3
College of Mechanical and Electronic Engineering, Shandong University of Science and Technology, Qingdao 266590, China
4
Beijing Tianma Intelligent Control Technology Co., Ltd., Beijing 101399, China
*
Author to whom correspondence should be addressed.
Processes 2025, 13(8), 2584; https://doi.org/10.3390/pr13082584
Submission received: 2 July 2025 / Revised: 31 July 2025 / Accepted: 13 August 2025 / Published: 15 August 2025
(This article belongs to the Special Issue Welding and Additive Manufacturing Processes)

Abstract

Laser powder bed fusion (LPBF) can fabricate hydraulic components with significant weight reduction, and in this study, small punch tests (SPTs) evaluated the tensile and fracture properties of four stainless steels (30Cr13, 316L, 15-5PH, 17-4PH), alongside metallographic, scanning electron microscope (SEM), and Electron Backscatter Diffraction (EBSD) analyses which examined their fracture modes, grain orientation, phase distribution, and grain boundary distribution. The tensile property results showed ductility rankings as 316L > 17-4PH > 15-5PH > 30Cr13, with correlations between Rp0.2 and Rm from SPT and uniaxial tensile tests for all four, while high-magnification SEM fractographs revealed ductile dimples on 15-5PH, 17-4PH, and 316L SPT specimens versus distinct cleavage fracture in 30Cr13. EBSD analysis indicated austenite content order as 316L > 17-4PH > 30Cr13 > 15-5PH, grain size order as 316L > 17-4PH > 15-5PH > 30Cr13, and high-angle grain boundaries ranking as 15-5PH > 30Cr13 > 17-4PH > 316L; additionally, notched SPT specimens inspected per EN 10371 for fracture toughness showed J-integral (JIC) values in the order 316L > 17-4PH > 15-5PH > 30Cr13, consistent with ductility and grain size results.

1. Introduction

Water-based hydraulic technology utilizes water-based fluids as the working medium, offering an environmentally friendly, safe, and resource-efficient alternative to traditional oil-based hydraulic systems in specific applications. Lightweight design serves as the primary strategy for water-based hydraulic systems to address weight restrictions, broaden application possibilities, and enhance performance. This approach not only reduces equipment weight and power demands for drive units but also enhances equipment durability, maneuverability, and load-bearing capacity while promoting energy efficiency and emission reduction. The integration of additive manufacturing (AM) technology, in conjunction with layout optimization and topological optimization design, enables hydraulic components and system structures to be more compact, smaller in volume, and lighter in weight [1]. The whole AM family is categorized with seven different sub-strategies according to their process variables and printing material types, including fused filament fabrication, vat-photopolymerization, material jetting techniques, binder jetting, laser powder bed fusion (LPBF), direct energy deposition, and sheet lamination methods. Among them, LPBF is primarily employed for the production of metal and ceramic objects, making it a prominent AM process for producing mechanical components [2]. This technique employs a high-power-density laser to melt metallic powders and join them together [3,4,5]. Many studies have focused on the material properties and mechanical behaviors of LPBF-manufactured materials using conventional standardized specimens, e.g., [6,7,8]. The conventional standardized specimens usually require a bulk of material, which is difficult or unapplicable to extract from in-service components. Hence, effective testing methods for remaining-life assessment of LPBF-manufactured components using a small amount of material are desirable.
Small punch tests (SPTs) only require a small volume of material for sampling without affecting the integrity and operational performance of in-service components. SPTs are widely used to estimate various material properties, such as tensile properties [9,10,11], fracture toughness [12,13], creep properties [14,15], and fatigue properties [16,17]. In recent years, some studies have been performed to evaluate the material properties of additive manufactured materials.
Lucon et al. [13] performed SPTs on additively manufactured (ALM) Ti-6Al-4V specimens fabricated under various processing parameters and heat-treatment conditions to assess their tensile and fracture toughness. Hurst et al. [18] conducted SPTs to evaluate the tensile and creep characteristics of ALM alloys, including IN718 and Ti-6Al-4V, utilizing small disc specimens obtained from a thin-walled ALM aerofoil blade. Lulu-Bitton et al. [19] studied the effects of gaseous hydrogen charging and electrochemical hydrogen charging on the microstructure and mechanical properties of wrought and electron beam melted (EBM) Ti-6Al-4V alloys. Wang et al. [20] performed uniaxial creep and small punch creep tests on deposited and ultrasonic micro-forging treated (UMFT)-assisted AM Hastelloy X to assess creep damage induced by AM defects. Lewis et al. [21] examined the fatigue behavior of LPBF nickel-based superalloy C263 and EBM titanium alloy Ti-6Al-4V in comparison with conventionally manufactured versions of these alloys. Wang et al. [17] assessed the fatigue performance of AM GH4169 related to AM defects and utilized various machine learning algorithms to predict small punch fatigue life. Mao and Takahashi [22] conducted pioneering work to correlate the Fe and Fm values from SPTs with tensile test-obtained Rp0.2 and Rm values. Song et al. [23] developed empirical correlations between the results of SPTs and uniaxial tensile tests for an anisotropic ASTM A350 alloy extracted from a failed forging flange. García et al. [24] discussed the effects of using different methods to obtain Fe from SPTs on the accuracy of the yield stress estimation. Janča et al. [25] summarized and compared different methods for obtaining Fe.
Previous studies have predominantly examined the material properties and mechanical behavior of LPBF-manufactured materials using conventional standard specimens, which necessitate a significant material volume and are challenging to extract from operational components. Consequently, researchers have adopted the small punch test (SPT) method to analyze the properties of additive manufacturing materials and traditional non-additive manufacturing materials. However, prior investigations have been restricted to individual SPT specimens. This study employed both standard and notched specimens to assess the tensile and fracture characteristics of LPBF-manufactured stainless steels including 30Cr13, 316L, 15-5PH, and 17-4PH. Force-displacement profiles were obtained for two SPT specimens for each of the four LPBF-manufactured materials. By refining the precision and focus of the analysis on tensile properties and fracture behavior, this study enhances the accuracy and comprehensiveness of small-sample testing for evaluating the mechanical properties of LPBF stainless steels. This research presents a precise experimental approach for appraising the remaining operational lifespan of components fabricated from such materials.

2. Small Punch Testing

2.1. Materials and Test Specimen

This study utilizes SPT testing on specimens composed of various materials to offer insights for material selection and lightweight structural design in components of water hydraulic systems. The medium in water hydraulic systems, a water-based fluid, can lead to corrosion of metal parts. Hence, this investigation concentrates on examining four stainless steel grades: 30Cr13, 316L, 15-5PH, and 17-4PH. Notably, 30Cr13, with elevated carbon content, exhibits superior strength, hardness, and wear resistance but comparatively weaker corrosion resistance. Next, 316L, an austenitic stainless steel containing molybdenum, delivers exceptional corrosion resistance, good weldability, and ductility, facilitating ease of processing and forming. Moreover, 15-5PH, a precipitation-hardening stainless steel, combines high strength with commendable corrosion resistance, achieving high hardness and toughness post heat treatment, maintaining stability at elevated temperatures, and demonstrating good machinability. Additionally, 17-4PH, another precipitation-hardening stainless steel, showcases high strength, high toughness, and excellent corrosion resistance, along with remarkable fatigue resistance, consistent mechanical properties at room temperature, and good machinability. These stainless steels are produced using the LPBF process, with production equipment employing high-speed laser cladding equipment (output power approximately 6000 W, spindle speed 1500 r/min, three-axis translation speed approximately 10 mm/s, powder feeding accuracy approximately 1.5%). Self-fusing iron-based alloy powders are used (powder production process: gas atomization and air cooling), with a particle size range of −300 to +500 M and a weld overlay reference hardness of 44-47HRC. The chemical composition of stainless steel is shown in Table 1. Uniaxial tensile tests were conducted on four types of stainless steel to determine their yield stress and ultimate strength. The tests were conducted in accordance with GB/T228.1-2021, using a computer-controlled electronic universal testing machine (WDW-300) with a test temperature of 23 °C and a humidity of 51% RH. One sample was tested for each type of stainless steel (manufactured using traditional processes), and the strength values are shown in Table 2.
The standard SPT specimen used in this study [11] has a diameter of 10 mm, a thickness of 0.5 mm, and an arithmetic mean roughness (Ra) of 0.4 µm. EN 10371 [26] outlines a method for determining fracture toughness, J-integral (JIC), utilizing a notched SPT specimen with identical dimensional parameters to the standard SPT specimen. The notched SPT specimen features a lateral notch at its center, simulating a crack. To secure the specimen in the test rig, the effective crack length (aeff) is approximately 4 mm. The correlations between the SPT data and uniaxial test results were determined following the guidelines outlined in CEN CWA 15672 [27] and EN ECISS/TC101/WG1 [28]. The wear mechanism of the standard SPT specimens was examined through scanning electron microscopy (SEM) analysis. Refer to Figure 1 for the dimensions of the notched SPT specimen.

2.2. Test Apparatus and Procedure

The test rig depicted in Figure 2 was utilized to determine the tensile properties and fracture toughness of stainless steel through SPTs; Figure 2b shows the various components of the SPT testing device and the positions of the test samples. The dimensions of the test rig align with those detailed in a prior investigation [11]. In the SPTs, a compression bar drives a spherical indenter to compress the clamped specimens at a constant rate of 0.5 mm/min under an external force. Simultaneously, a data acquisition system records the force on the compression bar and the deflection change at the center of the specimen’s lower surface to generate force-deflection curves. The axisymmetrically tested specimens primarily experience radial and circumferential tensile stresses. With increasing displacement applied to the specimen, its deformation escalates until eventual fracture. Four types of LPBF-manufactured stainless steel are used as test objects, with three samples prepared for each type. Three repeated tests are conducted for each type of stainless steel.

3. Results and Discussion

3.1. Metallographic Analyses

Metallographic analysis was conducted on samples of 15-5Ph, 17-4Ph, 3Cr13, and 316L, prepared through wire cutting, sandpaper grinding, polishing, and etching. The metallographic structures of the LPBF-manufactured stainless steels are depicted in Figure 3. Figure 3a,d illustrate tempered martensite in LPBF-manufactured 15-5PH and austenite structures in 316L, respectively. Conversely, Figure 3b,c exhibit the metallographic structure of 17-4PH and 30Cr13, likely composed of tempered troostite with notable carbide presence at the grain boundaries.

3.2. SPT Aanalysis

3.2.1. Method for Determining the Elastic–Plastic Transition Force

Numerous studies have examined the empirical relationships between the strength properties (Rp0.2 and Rm) and the characteristic parameters derived from SPT force-deflection outcomes. Equations (1)–(3), established in prior research [23,24,25], are commonly acknowledged correlations:
R p 0.2 = α 1 ( F e h 0 2 ) + α 2
R m = β 1 ( F m u m h 0 ) + β 2
R m = β 1 ( F m u m h 0 ) + β 2
where α 1 , α 2 , β 1 , β 2 , β 1 , and β 2 are the empirical correlation factors which depend on the geometry of the test rig.
In our previous study [11], we utilized two different methods to determine the elastic–plastic transition force, Fe, as recommended by CEN CWA 15672 and EN ECISS/TC101/WG1. In the current study, we also employed these two methods to determine Fe and plotted the results according to standard documents. Figure 4 illustrates a typical result for determining Fe in LPBF-manufactured 15-5PH using both methods. Table 3 provides a summary of the F e ( C W A ) and F e ( t w o   s e c a n t s ) values determined for the four LPBF-manufactured stainless steels. The empirical correlation factors in Equations (1)–(3) can be derived by correlating F e h 0 2 , F m u m h 0 , and F m h 0 2 values calculated from the SPT test data with the uniaxial tensile test data for Rp0.2 and Rm in the four LPBF-manufactured stainless steels. Table 4 presents the empirical correlation factor values.

3.2.2. SPT Results and Estimation of Strength Properties

Figure 5 displays the force-deflection curves of four materials manufactured using LPBF. Among these materials, 30Cr13 exhibits the highest scatter in results compared to the other stainless steels. Notably, all materials, except for 316L, demonstrate a characteristic behavior in SPT curve: initially increasing with deflection, then experiencing a sudden decrease (pop-in), followed by a subsequent increase. Upon reaching maximum load, the load decreases with further deflection, indicating material cracking. The deflection at which cracking occurs is observed to be the smallest for 30Cr13, followed by 15-5PH and 17-4PH, with 316L exhibiting the highest. Consequently, the ductility ranking of the stainless steels is as follows: 316L > 17-4PH > 15-5PH > 30Cr13. Prior to cracking, 30Cr13 demonstrates the highest SPT force at the same deflection, followed by 17-4PH, 15-5PH, and 316L, suggesting that 30Cr13 possesses the highest tensile strength, followed by 17-4PH, 15-5PH, and 316L.
It can be seen in Figure 6 and Figure 7 that the offsets are distributed in the lower and upper bounds within a 2-factor and a 1.5-factor of the unity lines, respectively. This could give reasonably accurate predictions for Rp0.2 and Rm.

3.3. SEM Analyses on the Ruptured Surfaces of Standard SPT Specimens

SEM fractographs of the ruptured surfaces of SPT specimens for four LPBF-manufactured stainless steels are presented in Figure 8a–h. Figure 8a displays a low-magnification SEM micrograph of the 30Cr13 LPBF-manufactured SPT specimen, revealing cracks originating from the specimen’s bottom and extending radially towards it, with no circumferential cracking observed. In contrast, Figure 8c shows the SEM micrograph of the 316L LPBF-manufactured SPT specimen, where radial cracks are absent. For the LPBF-manufactured 15-5PH and 17-4PH SPT specimens, circumferential cracking is predominant, as depicted in Figure 8e,g, with some radial cracks also visible, particularly pronounced in the 15-5PH specimens compared to the 17-4PH specimens. This suggests slightly better ductility in 17-4PH, consistent with the force-deflection curve results in Figure 5. High-magnification images in Figure 8d,f,h reveal numerous ductile dimples on the fracture surfaces of the LPBF-manufactured 316L, 15-5PH, and 17-4PH SPT specimens. However, a cleavage fracture is evident at high magnification in the LPBF-manufactured 30Cr13 SPT specimen.

3.4. EBSD Analyses on the LPBF-Manufactured Four Stainless Steels

Samples of 30Cr13, 316L, 15-5PH, and 17-4PH prepared using the LPBF process were analyzed for grain orientation, phase distribution, and grain boundary distribution. The microstructural information of 30Cr13 is shown in Figure 9a–c. The inverse phase figure (IPF) of 30Cr13 (Figure 9a) indicates the absence of a distinct microstructure. The phase distribution diagram (Figure 9b) shows a combination of martensite and partial austenite phases. Similarly, 316L also lacks a distinct microstructure, primarily composed of austenite, as shown in Figure 10a,). The IPF of 15-5PH (Figure 11a) also indicates the absence of microstructural features, while the phase distribution diagram (Figure 11b) shows a primary martensite phase (BCC) with a small amount of austenite phase (FCC). For 17-4PH, Figure 12a,b show the absence of microstructural features, with the primary phases being martensite and austenite. EBDS analysis of the phase distribution of the four stainless steels further validates the results of the metallographic analysis. EBDS analysis of 17-4PH and 30Cr13 shows that both have martensite as the primary phase, and the presence of carbides may be related to precipitation strengthening of the martensite matrix, further supporting the observation of “grain boundary carbides” in the metallographic analysis. The austenite content decreases progressively in the order of 316L, 17-4PH, 30Cr13, and 15-5PH. The austenite phase exhibits excellent plasticity, and EBDS analysis shows that 316L has the highest austenite content, supporting the finding that 316L has the best ductility. Grain size is largest in 316L, followed by 17-4PH, 15-5PH, and 30Cr13. Grain size is a key factor influencing strength (fine-grain strengthening effect), so 30Cr13 exhibits the highest tensile strength due to fine-grain strengthening, while 316L has the lowest strength due to coarse grains. As shown in Figure 9c, Figure 10c, Figure 11c and Figure 12c, 15-5PH has the highest content of high-angle grain boundaries (the red portions in the grain boundary information diagram represent high-angle grain boundaries). Next, 30Cr13 has a lower content of high-angle grain boundaries, followed by 17-4PH, while 316L has the lowest content of high-angle grain boundaries. Moreover, 30Cr13 has a low austenite content (with brittle martensite dominating) and a low content of high-angle grain boundaries (weak resistance to crack propagation), resulting in typical cleavage brittle fracture during fracture, consistent with the SEM features of “no ductile pits and rapid radial crack propagation.” Furthermore, 316L, 17-4PH, and 15-5PH all contain austenite, whose plastic deformation capability can absorb fracture energy, resulting in numerous ductile dimples on the fracture surface (a characteristic of plastic fracture); among these, 17-4PH has a higher austenite content than 15-5PH and a lower content of high-angle grain boundaries than 15-5PH (though the influence of grain boundaries on toughness is dominated by the microstructural composition), resulting in fewer radial cracks and slightly better ductility in 17-4PH compared to 15-5PH, consistent with SEM observation results.

3.5. Estimations of Fracture Toughness, JIC

The force-deflection curves of notched SPT specimens for the four AM stainless steels are depicted in Figure 13. The deflection at the maximum force, um, can be derived from Figure 13 and is presented in Table 5. In accordance with EN 10371 [26], the deflection at the maximum force, um, of the tested notched SPT specimen can be correlated with the notch opening displacement, δIC, corresponding to crack initiation, as illustrated in Figure 14. The linear relationship between fracture toughness, JIC, and the notch opening displacement, δIC, is defined by Equation (4):
J I C S P T = R p 0.2 × δ I C
The fracture toughness, J I C S P T , estimated through notched opening displacement is detailed in Table 6. It is evident from Table 6 that the highest J I C S P T value is observed in LPBF-manufactured 316L, followed by the second highest J I C S P T in LPBF-manufactured 17-4PH. In contrast, 30Cr13, among the other three LPBF-manufactured stainless steels, exhibits the lowest J I C S P T .

3.6. Summary

The study confirms the efficacy of the SPT in assessing mechanical properties of additive manufacturing materials and establishing correlations with traditional uniaxial test results. Previous research has mainly examined single SPT specimens, focusing on materials like titanium and nickel alloys. This study investigates four types of LPBF stainless steel (30Cr13, 316L, 15-5PH, 17-4PH) using both standard and notched SPT specimens. Through microstructural analysis, SEM, and EBSD analysis, it elucidates differences in tensile properties, fracture modes, and the relationship between microstructural characteristics (composition, grain size, grain boundary distribution) and macroscopic properties. This research addresses gaps in SPT studies on additively manufactured stainless steels, offering insights into their unique performance patterns and the impact of texture on performance, surpassing existing studies focused on macroscopic parameters.

4. Conclusions

In this study, the tensile and fracture properties of four LPBF-manufactured stainless steels were explored through SPTs using two types of small disc specimens. The following are the conclusions.
Metallographic analysis of stainless steels produced by LPBF reveals that 15-5Ph exhibits tempered martensite, 316L shows austenite, and both 17-4PH and 30Cr13 display tempered martensite with abundant carbides at grain boundaries. Tensile testing indicates that 316L possesses the highest ductility, followed by 17-4PH, 15-5PH, and 30Cr13, while 30Cr13 exhibits the highest tensile strength, followed by 17-4PH, 15-5PH, and 316L. Differences in metallographic structure significantly influence mechanical properties. For instance, 316L, characterized by an austenitic structure, exhibits superior ductility, whereas 30Cr13, with carbide inclusions, demonstrates the highest tensile strength but the lowest ductility. The correlation between Rp0.2 and Rm values obtained from SPT testing and those acquired from uniaxial tensile testing was established for the four stainless steel variants produced via LPBF.
High magnification SEM fracture images reveal that SPT specimens of 30Cr13 steel grade produced by LPBF display cleavage fracture, demonstrating distinct brittle fracture features. In contrast, SPT specimens of 15-5PH, 17-4PH, and 316L steel grades manufactured by LPBF exhibit numerous ductile pits on the fracture surface, indicating clear plastic fracture characteristics. Notably, 17-4PH shows fewer radial cracks compared to 15-5PH, suggesting slightly superior ductility, aligning with the SPT curve outcomes.
EBSD analysis revealed that none of the four stainless steel grades displayed significant texture. The order of austenite content was found to be 316L > 17-4PH > 30Cr13 > 15-5PH. The austenite phase demonstrated superior plasticity, with 316L exhibiting the highest austenite content (single austenite phase), correlating with its superior ductility. In terms of grain size, the sequence was 316L > 17-4PH > 15-5PH > 30Cr13. Grain size significantly impacts strength (fine-grain strengthening effect), leading to 30Cr13 displaying the highest tensile strength due to fine-grain strengthening, while 316L showed the lowest strength due to coarse grains. Regarding high-angle grain boundaries, the order was 15-5PH > 30Cr13 > 17-4PH > 316L.
Fracture toughness of notched SPT specimens was evaluated in accordance with the EN 10371 standard [26]. The J I C S P T values were ranked as follows, consistent with the observed ductility and grain size trends: 316L > 17-4PH > 15-5PH > 30Cr13.

Author Contributions

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

Funding

This research was supported by CCTEG Project (2023-TD-MS015, 2023-TD-QN004) and TMIC Project (2022TM-167M) funding, China.

Data Availability Statement

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

Conflicts of Interest

Authors Ran Li, Wenshu Wei, Mengyu Wu, Yuehua Lai, Rongming Chen, Hao Liu, Jian Ye, and Jianfeng Li were employed by the company Beijing Tianma Intelligent Control Technology Co. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest. The authors declare that this study received funding from CCTEG Project (2023-TD-MS015, 2023-TD-QN004) and the TMIC Project (2022TM-167M).

References

  1. Kong, X.D.; Zhu, Q.X.; Yao, J.; Shang, Y.X.; Zhu, Y. Basic Theory and Key Technology of “New Method for Lightweight Design and Manufacturing of Hydraulic Components and Systems”. J. Mech. Eng. 2021, 57, 4–12. [Google Scholar] [CrossRef]
  2. Han, M.; Kaşıkcıoğlu, S.; Ergene, B.; Atlıhanet, G.; Bolat, C. An Experimental Investigation on the Mechanical and Wear Responses of Lightweight AlSi10Mg Components Produced by Selective Laser Melting: Effects of Building Direction and Test Force. Sci. Sinter. 2025, 27, 87–101. [Google Scholar] [CrossRef]
  3. Álvarez-Trejo, A.; Cuan-Urquizo, E.; Bhate, D.; Roman-Flores, A. Mechanical metamaterials with topologies based on curved elements: An overview of design, additive manufacturing and mechanical properties. Mater. Des. 2023, 233, 112190. [Google Scholar] [CrossRef]
  4. Lashgari, H.R.; Ferry, M.; Li, S. Additive manufacturing of bulk metallic glasses: Fundamental principle, current/future developments and applications. J. Mater. Sci. Technol. 2022, 119, 131–149. [Google Scholar] [CrossRef]
  5. Harun, W.S.W.; Kamariah, M.; Muhamad, N.; Ghani, S.A.C.; Ahmad, F.; Mohamed, Z. A review of powder additive manufacturing processes for metallic biomaterials. Powder Technol. 2018, 327, 128–151. [Google Scholar] [CrossRef]
  6. Heo, S.; Lim, Y.; Kwak, N.; Jeon, C.; Choi, M.; Jo, I. Impact of Heat Treatment and Building Direction on Tensile Properties and Fracture Mechanism of Inconel 718 Produced by SLM Process. Metals 2024, 14, 440. [Google Scholar] [CrossRef]
  7. Zhang, S.M.; He, X.Q.; Shan, C.Y.; Xu, C.T.; Lu, Y.F.; Song, Z.K. Effect of Annealing on Microstructure and Tensile Properties of Selective Laser Melting MAR-M509 Superalloy. J. Mater. Eng. Perform. 2024, 34, 11779–11790. [Google Scholar] [CrossRef]
  8. Ponnusamy, P.; Sharma, B.; Masood, S.H.; Rashid, R.A.R.; Rashid, R.; Palanisamy, S.; Ruan, D. A study of tensile behavior of SLM processed 17-4 PH stainless steel. Materials 2021, 45, 4531–4534. [Google Scholar] [CrossRef]
  9. Bruchhausen, M.; Holmström, S.; Simonovski, I.; Austin, T.; Lapetite, J.M.; Ripplinger, S.; de Haan, F. Recent developments in small punch testing: Tensile properties and DBTT. Theor. Appl. Fract. Mech. 2016, 86, 2–10. [Google Scholar] [CrossRef]
  10. Haroush, S.; Priel, E.; Moreno, D.; Busiba, A.; Silverman, I.; Turgeman, A.; Shneck, R.; Gelbstein, Y. Evaluation of the mechanical properties of SS-316L thin foils by small punch testing and finite element analysis. Mater. Des. 2015, 83, 75–84. [Google Scholar] [CrossRef]
  11. Li, R.; Wei, W.S.; Chen, R.M.; Wu, M.Y.; Lai, Y.H. Evaluating the tensile properties of high-strength stainless steels using small punch testing. Sci. Prog. 2024, 107, 1–19. [Google Scholar] [CrossRef]
  12. Foulds, J.; Viswanathan, R. Determination of the toughness of in-service steam turbine disks using small punch testing. J. Mater. Eng. Perform. 2001, 10, 614–619. [Google Scholar] [CrossRef]
  13. Lucon, E.; Benzing, J.T.; Derimow, N.; Harbe, N. Small punch testing to estimate the tensile and fracture properties of additively manufactured Ti–6Al–4V. J. Mater. Eng. Perform. 2021, 30, 5039–5049. [Google Scholar] [CrossRef]
  14. Hyde, T.H.; Hyde, C.J.; Sun, W. Theoretical basis and practical aspects of small specimen creep testing. J. Strain Anal. Eng. Des. 2012, 48, 112–125. [Google Scholar] [CrossRef]
  15. Zhuang, F.K.; Zhou, G.Y.; Tu, S.T. Numerical investigation of frictional effect on measuring accuracy of different small specimen creep tests. Int. J. Press. Vessel. Pip. 2013, 110, 42–49. [Google Scholar] [CrossRef]
  16. Prakash, R.V.; Arunkumar, S. Evaluation of damage in materials due to fatigue cycling through static and cyclic small punch testing. Small Specim. Test Tech. 2014, 6, 168–186. [Google Scholar] [CrossRef]
  17. Wang, X.; Xu, L.; Zhao, L.; Han, Y. Evaluation of defect-related fatigue performance of additive manufacturing GH4169 via small punch test. Theor. Appl. Fract. Mech. 2023, 128, 104162. [Google Scholar] [CrossRef]
  18. Hurst, R.C.; Lancaster, R.J.; Jeffs, S.P.; Bache, M.R. The contribution of small punch testing towards the development of materials for aero-engine applications. Theor. Appl. Fract. Mech. 2016, 86, 69–77. [Google Scholar] [CrossRef]
  19. Lulu-Bitton, N.; Navi, N.U.; Rosen, B.A.; Haroush, S.; Sabatani, E.; Eretz-Kdosha, Y.; Agronov, G.; Eliaz, N. The influence of gaseous hydrogen charging on the microstructural and mechanical behavior of electron beam melted and wrought Ti-6Al-4V alloys using the small punch test. Int. J. Hydrogen Energy 2023, 48, 34077–34093. [Google Scholar] [CrossRef]
  20. Wang, X.; Xu, L.; Zhao, L.; Han, Y.; Liu, Z. Defect-based additive manufactured creep performance evaluation via small punch test. Int. J. Mech. Sci. 2024, 279, 109565. [Google Scholar] [CrossRef]
  21. Lewis, D.T.S.; Lancaster, R.J.; Jeffs, S.P.; Illsley, H.W.; Davies, S.J.; Baxter, G.J. Characterising the fatigue performance of additive materials using the small punch test. Mater. Sci. Eng. 2019, 754, 719–727. [Google Scholar] [CrossRef]
  22. Mao, X.; Takahashi, H. Development of a further-miniaturized specimen of 3 mm diameter for TEM disk (ø 3 mm) small punch tests. J. Nucl. Mater. 1987, 150, 42–52. [Google Scholar] [CrossRef]
  23. Song, M.; Guan, K.; Qin, W.; Jerzy, A.; Szpunar. Comparison of mechanical properties in conventional and small punch tests of fractured anisotropic A350 alloy forging flange. Nucl. Eng. Des. 2012, 247, 58–65. [Google Scholar] [CrossRef]
  24. García, T.E.; Rodríguez, C.; Belzunce, F.J.; Suárez, C. Estimation of the mechanical properties of metallic materials by means of the small punch test. J. Alloys Compd. 2014, 582, 708–717. [Google Scholar] [CrossRef]
  25. Janca, A.; Siegl, J.; Haušild, P. Small punch test evaluation methods for material characterisation. J. Nucl. Mater. 2016, 481, 201–213. [Google Scholar] [CrossRef]
  26. EN 10371; Metallic Materials. Small Punch Test Method. The European Committee for Standardization: Brussels, Belgium, 2021.
  27. CWA 15627:2006 E; Small Punch Test Method for Metallic Materials. European Committee for Standardization: Brussels, Belgium, 2006.
  28. Bruchhausen, M.; Altstadt, E.; Austin, T.; Dymacek, P.; Holmström, S.; Jeffs, S.; Lacalle, R.; Lancaster, R.; Matocha, K.; Petzova, J. European standard on small punch testing of metallic materials. Ubiquity Proc. 2018, 1, 11. [Google Scholar] [CrossRef]
Figure 1. The dimensions of the notched SPT specimen.
Figure 1. The dimensions of the notched SPT specimen.
Processes 13 02584 g001
Figure 2. The test apparatus of SPTs: (a) Physical picture of SPTs test apparatus; (b) schematic representation of SPTs test apparatus.
Figure 2. The test apparatus of SPTs: (a) Physical picture of SPTs test apparatus; (b) schematic representation of SPTs test apparatus.
Processes 13 02584 g002
Figure 3. Metallographic structure analyses for the four LPBF-manufactured stainless steels: (a) 15-5PH; (b) 17-4PH; (c) 30Cr13; (d) 316L.
Figure 3. Metallographic structure analyses for the four LPBF-manufactured stainless steels: (a) 15-5PH; (b) 17-4PH; (c) 30Cr13; (d) 316L.
Processes 13 02584 g003
Figure 4. A typical result for determining Fe on a SPT force-deflection curve of LPBF-manufactured 15-5PH.
Figure 4. A typical result for determining Fe on a SPT force-deflection curve of LPBF-manufactured 15-5PH.
Processes 13 02584 g004
Figure 5. The force-deflection curves of SPTs for LPBF-manufactured stainless steels.
Figure 5. The force-deflection curves of SPTs for LPBF-manufactured stainless steels.
Processes 13 02584 g005
Figure 6. Estimated R0.2 from SPTs against uniaxial results for the four LPBF-manufactured stainless steels by use of the CWA and two secants methods for determining Fe.
Figure 6. Estimated R0.2 from SPTs against uniaxial results for the four LPBF-manufactured stainless steels by use of the CWA and two secants methods for determining Fe.
Processes 13 02584 g006
Figure 7. Estimated Rm from SPTs against uniaxial results for the four LPBF-manufactured stainless steels for the normalization Fm by umh0 and h02.
Figure 7. Estimated Rm from SPTs against uniaxial results for the four LPBF-manufactured stainless steels for the normalization Fm by umh0 and h02.
Processes 13 02584 g007
Figure 8. SEM fractographs depicting the ruptured surfaces of the SPT specimens of the four LPBF-manufactured stainless steels: (a) 30Cr13 at low magnification; (b) 30Cr13 at high magnification; (c) 316L at low magnification; (d) 316L at high magnification; (e) 15-5PH at low magnification; (f) 15-5PH at high magnification; (g) 17-4PH at low magnification; (h) 17-4PH at high magnification.
Figure 8. SEM fractographs depicting the ruptured surfaces of the SPT specimens of the four LPBF-manufactured stainless steels: (a) 30Cr13 at low magnification; (b) 30Cr13 at high magnification; (c) 316L at low magnification; (d) 316L at high magnification; (e) 15-5PH at low magnification; (f) 15-5PH at high magnification; (g) 17-4PH at low magnification; (h) 17-4PH at high magnification.
Processes 13 02584 g008aProcesses 13 02584 g008b
Figure 9. EBSD analyses on the LPBF-manufactured 30Cr13: (a) inverse pole figure; (b) phase composition; (c) grain boundary distribution.
Figure 9. EBSD analyses on the LPBF-manufactured 30Cr13: (a) inverse pole figure; (b) phase composition; (c) grain boundary distribution.
Processes 13 02584 g009aProcesses 13 02584 g009b
Figure 10. EBSD analyses on the LPBF-manufactured 316L: (a) inverse pole figure; (b) phase composition; (c) grain boundary distribution.
Figure 10. EBSD analyses on the LPBF-manufactured 316L: (a) inverse pole figure; (b) phase composition; (c) grain boundary distribution.
Processes 13 02584 g010aProcesses 13 02584 g010b
Figure 11. EBSD analyses on the LPBF-manufactured 15-5PH: (a) inverse pole figure; (b) phase composition; (c) grain boundary distribution.
Figure 11. EBSD analyses on the LPBF-manufactured 15-5PH: (a) inverse pole figure; (b) phase composition; (c) grain boundary distribution.
Processes 13 02584 g011aProcesses 13 02584 g011b
Figure 12. EBSD analyses on the LPBF-manufactured 17-4PH: (a) inverse pole figure; (b) phase composition; (c) grain boundary distribution.
Figure 12. EBSD analyses on the LPBF-manufactured 17-4PH: (a) inverse pole figure; (b) phase composition; (c) grain boundary distribution.
Processes 13 02584 g012aProcesses 13 02584 g012b
Figure 13. The force-deflection curves of notched SPTs for LPBF-manufactured stainless steels.
Figure 13. The force-deflection curves of notched SPTs for LPBF-manufactured stainless steels.
Processes 13 02584 g013
Figure 14. Correlation between the deflection at the maximum force, um, of the tested notched SPT specimen and the notch opening displacement, δIC, corresponding to crack initiation.
Figure 14. Correlation between the deflection at the maximum force, um, of the tested notched SPT specimen and the notch opening displacement, δIC, corresponding to crack initiation.
Processes 13 02584 g014
Table 1. Chemical compositions of stainless steels specimens (wt%).
Table 1. Chemical compositions of stainless steels specimens (wt%).
MaterialsCPSiMoCrMnFeAlNiNbSCuTa
30Cr130.320.020.83 13.580.89Bal 0.42 0.013
316L0.0210.0320.81 16.761.53Bal 12.56 0.019
15-5PH0.0560.0280.6 14.320.58Bal 4.210.330.00283.150.0006
17-4PH0.0610.0230.74 16.310.67Bal 4.360.320.0033.860.0005
Table 2. The yielding stresses and ultimate strengths of stainless-steel specimens.
Table 2. The yielding stresses and ultimate strengths of stainless-steel specimens.
MaterialsYielding Stress (MPa)Ultimate Strength (MPa)
30Cr137531829
316L373620
15-5PH8061431
17-4PH7451162
Table 3. Determined F e ( C W A ) and F e ( t w o   s e c a n t s ) values for the four LPBF-manufactured stainless steels.
Table 3. Determined F e ( C W A ) and F e ( t w o   s e c a n t s ) values for the four LPBF-manufactured stainless steels.
Materials F e ( C W A ) , N F e ( t w o   s e c a n t s ) , N
30Cr13666.531137
621.06976
7201122
316L162.37165.66
169.26216.07
621.06994.63
15-5PH479.06775.2
399694
602.48781
17-4PH339.93462.59
562.7800.85
601.06788.23
Table 4. The values of empirical correlations factors.
Table 4. The values of empirical correlations factors.
α 1 ( C W A ) α 2 ( C W A ) α 1 ( t w o   s e c a n t s ) α 2 ( t w o   s e c a n t s ) β 1 β 2 β 1 β 2
0.0477550.620.1196365.770.142727.71−0.0981921.4
Table 5. Deflection at the maximum force, um, of the tested notched SPT specimen.
Table 5. Deflection at the maximum force, um, of the tested notched SPT specimen.
Materialsum for Different Specimens, mmAverage Value of um, mm
30Cr130.2130.256
0.264
0.290
316L1.351.39
1.42
1.40
15-5PH0.7690.760
0.729
0.781
17-4PH1.041.04
1.04
1.04
Table 6. The fracture toughness, J I C S P T , estimated by using the notched opening displacement.
Table 6. The fracture toughness, J I C S P T , estimated by using the notched opening displacement.
Materials δ I C Determined by um, mm J I C S P T
30Cr130.06750
316L1.218455
15-5PH0.291234
17-4PH0.569424
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

Li, R.; Wei, W.; Wu, M.; Liu, F.; Li, W.; Lai, Y.; Chen, R.; Liu, H.; Ye, J.; Li, J.; et al. Tensile and Fracture Properties Evaluation of Additively Manufactured Different Stainless Steels via Small Punch Testing. Processes 2025, 13, 2584. https://doi.org/10.3390/pr13082584

AMA Style

Li R, Wei W, Wu M, Liu F, Li W, Lai Y, Chen R, Liu H, Ye J, Li J, et al. Tensile and Fracture Properties Evaluation of Additively Manufactured Different Stainless Steels via Small Punch Testing. Processes. 2025; 13(8):2584. https://doi.org/10.3390/pr13082584

Chicago/Turabian Style

Li, Ran, Wenshu Wei, Mengyu Wu, Fengcai Liu, Wenbo Li, Yuehua Lai, Rongming Chen, Hao Liu, Jian Ye, Jianfeng Li, and et al. 2025. "Tensile and Fracture Properties Evaluation of Additively Manufactured Different Stainless Steels via Small Punch Testing" Processes 13, no. 8: 2584. https://doi.org/10.3390/pr13082584

APA Style

Li, R., Wei, W., Wu, M., Liu, F., Li, W., Lai, Y., Chen, R., Liu, H., Ye, J., Li, J., & Cao, T. (2025). Tensile and Fracture Properties Evaluation of Additively Manufactured Different Stainless Steels via Small Punch Testing. Processes, 13(8), 2584. https://doi.org/10.3390/pr13082584

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