Next Article in Journal
Effect of Electric Pulse Treatment on Hot-Dip Aluminizing Coating of Ductile Iron
Previous Article in Journal
Prediction of Geometric Characteristics of Melt Track Based on Direct Laser Deposition Using M-SVR Algorithm
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hardness–Deformation Energy Relationship in Metals and Alloys: A Comparative Evaluation Based on Nanoindentation Testing and Thermodynamic Consideration

1
Yamamoto Scientific Tool Lab. Co., Ltd., Funabashi, Chiba 273-0018, Japan
2
Department of Materials, Kyushu University, Nishi-ku, Fukuoka 819-0395, Japan
3
Center for Elements Strategy Initiative for Structural Materials, Kyoto University, Sakyo-ku, Kyoto 606-8501, Japan
*
Author to whom correspondence should be addressed.
Materials 2021, 14(23), 7217; https://doi.org/10.3390/ma14237217
Submission received: 25 October 2021 / Revised: 21 November 2021 / Accepted: 23 November 2021 / Published: 26 November 2021
(This article belongs to the Topic Metallurgical and Materials Engineering)

Abstract

:
Nanoindentation testing using a Berkovich indenter was conducted to explore the relationships among indentation hardness (H), elastic work energy (We), plastic work energy (Wp), and total energy (Wt = We + Wp) for deformation among a wide range of pure metal and alloy samples with different hardness, including iron, steel, austenitic stainless steel (H ≈ 2600–9000 MPa), high purity copper, single-crystal tungsten, and 55Ni–45Ti (mass%) alloy. Similar to previous studies, We/Wt and Wp/Wt showed positive and negative linear relationships with elastic strain resistance (H/Er), respectively, where Er is the reduced Young’s modulus obtained by using the nanoindentation. It is typically considered that Wp has no relationship with We; however, we found that Wp/We correlated well with H/Er for all the studied materials. With increasing H/Er, the curve converged toward Wp/We = 1, because the Gibbs free energy should not become negative when indents remain after the indentation. Moreover, H/Er must be less than or equal to 0.08. Thermodynamic analyses emphasized the physical meaning of hardness obtained by nanoindentation; that is, when Er is identical, harder materials show smaller values of Wp/We than those of softer ones during nanoindentation under the same applied load. This fundamental knowledge will be useful for identifying and developing metallic materials with an adequate balance of elastic and plastic energies depending on the application (such as construction or medical equipment).

1. Introduction

The mechanical properties of a material, particularly its strength and ductility, are the most fundamental metrics for evaluating its deformation behavior under applied stress. Ductility involves both uniform and local deformation, which occur by different mechanisms. Understanding the balance of these deformation modes is important for determining suitable materials for a given application. Hu et al. [1] showed that an increase in the local elongation of dual phase (DP) 980 steel correlated with an improvement in the hole expansion ratio obtained during hole piercing tests, classified as the edge stretchability. Taylor et al. [2] reported that the hole expansion ratio of DP980 steel decreased as the martensite hardness and martensite/ferrite hardness ratio (as determined by nanoindentation tests) increased. Furthermore, using synchrotron X-ray laminography of nanoscale precipitated steel and bainitic steel, Mugita et al. [3] showed that an increase in the nanoindentation hardness at the grain boundaries correlated with a decrease in local elongation, because the growth of voids was accelerated.
It is possible to analyze mechanical properties such as strength and elongation thermodynamically by integrating the stress–strain curve to determine the deformation energy [4,5]. However, thermodynamic analyses have rarely been used as a means of controlling the microstructure. Our research group has previously shown that the local deformation energy can be increased by controlling the microstructure of an alloy. Specifically, we clarified the effects of chromium-based steels with precipitations [6] and the austenite stability of duplex stainless steels [7] on the local deformation energy, as determined from the maximum load to fracture obtained from tensile stress–strain curves.
Hardness testing is widely used to evaluate the plastic deformability of materials. In the last few decades, nanoindentation testing methods [8,9,10,11,12,13,14,15,16,17,18] have been developed and are rapidly becoming widespread for the evaluation of various materials, including steels [19], tungsten [20], copper [21], and NiTi alloys [22]. This has produced new metallurgical knowledge about nanoscale deformation that is useful for the development of superior materials. Nanoindentation tests produce load–displacement curves that provide detailed information about the plastic deformation behavior during compression and unloading with nanoscale granularity, even under low loads (on the order of millinewtons). During the early development of nanoindentation testing techniques, Oliver and Pharr [23,24] proposed a method for determining the elastic modulus from the load–displacement curves. Sudden drops in deformation resistance during compression, called “pop-in events,” are sometimes observed in the load–displacement curves of nanoindentation tests [25]. This phenomenon is closely related to the dislocation nucleation behavior and indicates the transition point from elastic to elasto-plastic deformation [26]. Thus, nanoindentation testing can be applied in studies of both elastic and plastic deformation. Further, the elastic work energy (We), plastic work energy (Wp), and total work energy (Wt = We + Wp) can be determined by integrating the load–displacement curves, as described in an ISO standard [27], and most testing devices can now automatically calculate these energy values. Analyzing the deformation energy from hardness test results is therefore a valuable way of determining the relationships between the mechanical properties and microstructure of a material based on the thermodynamic properties.
Okoro et al. [28] added multi-walled carbon nanotubes (MWCNTs) to sintered Ti6Al4V-based nanocomposites and investigated the effect on the nanoindentation hardness (H), plasticity index (Wp/Wt), recovery index (We/Wt), and elastic strain resistance (H/Er). Er is the Young’s modulus obtained by nanoindentation (Er = π1/2/2 × S/Ac1/2, where S is the slope of the force–displacement curve during unloading, and Ac is the projected area of the indent). Their findings revealed that the addition of MWCNTs improved H, We/Wt, and H/Er, but decreased Wp/Wt. The anti-wear resistance was also improved. Cheng and Cheng [29] used finite element method (FEM) numerical analysis and experimental verification to show that, for various materials, H/Er has a negative linear relationship with Wp/Wt. In addition, Yang et al. [30] demonstrated that H/Er has a positive linear relationship with We/Wt for 20 sets of materials. The deformation energy obtained from nanoindentation tests is therefore a useful parameter for evaluating the deformation behavior of various materials. Moreover, it can be used in the development of new materials.
We and Wp represent the work required for different atomistical phenomena. Herein, We is the work required to change the atomic distance elastically, whereas Wp is the work required to move or multiplicate lattice defects such as dislocations and vacancies. The linear relationships between We/Wt and H/Er and between Wp/Wt and H/Er [29,30] strongly suggest that Wp/We can also be regarded as a function of H/Er. The correspondence between Wp/We and H/Er indicates that the mechanisms of elastic and plastic deformation are related. To explore this hypothesis, we measured H, We, and Wp by nanoindentation testing for six types of iron and steel, including austenitic stainless steels, with H values measured by nanoindentation (defined as HIT in ISO 14577 [31]) of approximately 2600–9000 MPa. In addition, high purity copper, single-crystal tungsten, and NiTi alloy were used as test materials. This paper analyzes the respective relationships between H, We, and Wp and determines the Wp/We ratio as a metric for identifying materials with an adequate balance of elastic and plastic energies for a given application, such as construction beams and guidewires for medical equipment. Finally, thermodynamic analysis was performed to elucidate the relationships among these properties.

2. Experimental Methods

2.1. Test Materials

A total of nine test materials was used, including six types of iron and steel: martensitic steel (Vickers hardness reference block, HMV700; Yamamoto Scientific Tool Lab. Co., Ltd., Funabashi, Japan; Commercial); bainitic steel (JFE Steel Co., Ltd., HQ: Tokyo, Japan; Laboratory); interstitial-free steel (JFE Steel Co., Ltd., HQ: Tokyo, Japan; Commercial); electrodeposited iron (MAIRON Grade SHP; Toho Zinc Co., Ltd., HQ: Tokyo, Japan; Laboratory); and two austenitic stainless steels (JIS SUS304 and JIS SUS316; Sanyo Special Steel Co., Ltd., HQ: Himeji, Japan; Commercial). The remaining test materials were high purity copper (Vickers hardness reference block, HMV40 JIS C1020P; Yamamoto Scientific Tool Lab. Co., Ltd., Funabashi, Japan; Commercial), single-crystal tungsten (MaTecK GmbH, HQ: Jülich, Germany; Commercial), and 55Ni–45Ti (mass%) alloy (prepared in a vacuum melting using a high frequency induction heating furnace with minimal contamination; heat-treated at 773 K for 3600 s in air followed by quenching; Daido Steel Co., Ltd. HQ: Nagoya, Japan; Laboratory) (hereafter referred to as NiTi alloy).
Table 1 lists the chemical compositions of the test alloys, as obtained by spark discharge emission spectroscopic analyses and wet chemical analyses. The electrodeposited iron, high purity copper, and single-crystal tungsten had purities of 99.98%, 99.99%, and 99.99%, respectively.

2.2. Nanoindentation Tests

The dimensions and preparation methods of the nanoindentation test specimens are shown in Table 2. Each specimen was wet-polished using emery paper, in sequential order from 320 to 2000 grit, and then mirror polished using a diamond abrasive and colloidal silica.
Nanoindentation tests were conducted using a nanoindentation hardness tester (Elionix Inc., ENT-2100) with a Berkovich indenter under a loading rate of 0.98 mN/s and maximum load of 9.8 mN. For each sample, H, We, and Wp were determined from the load–displacement curve (see Figure 1): H was determined as the ratio of the maximum test load to the projected contact area of the indenter [31], and We and Wp were determined as the areas of regions A and B, respectively [27]. These values were calculated automatically by software included with the nanoindentation tester. These measurement methods are described in ISO 14577 [27]. The number of measurement points was set at 30 points per specimen, and the values shown in the tables and figures are averages of these 30 points. We excluded abnormal data such as those of indentions on grain boundaries.

3. Results

Table 3 shows the nanoindentation test results measured under a load of 9.8 mN along with yield strength (YS), and tensile strength (TS) measured by tensile tests. The relationship between H/Er and We/Wt was determined based on these results (solid line in Figure 2). The relationship between H/Er and We/Wt was linear and intercepted the origin. The linear coefficient was obtained by using the least-squares method, including the origin, as denoted by the following equation:
H/Er = α(We/Wt)
where α is the slope of the line, which had a value of 0.16 in this study.
Yang et al. [30] previously investigated the relationship between H/Er and We/Wt by using FEM and experimental verification. They also found a good linear relationship between H/Er and We/Wt, which is shown as a dashed line in Figure 2. The gradient of the relationship obtained by Yang et al. is slightly higher than that in this study. This is because the regression line in their work was drawn directly from experimental data. However, they noted that the line should intercept the origin.
Figure 3 shows the relationship between H/Er and Wp/Wt for the experimental data obtained in this study (solid line) and data obtained by Cheng and Cheng [29] (dashed line). The data obtained by Cheng and Cheng are based on FEM and experimental verification (nanoindentation tests using a conical-type indenter). Both sets of data identified a linear relationship between H/Er and Wp/Wt. The linear relationship was approximated using the least-squares method as follows:
H/Er = β + γ (Wp/Wt)
where β and γ are constants, which had values of 0.16 and −0.16 in this study, respectively. Cheng and Cheng [29] also found a negative value of γ in Equation (2); however, the absolute value of γ obtained by Cheng and Cheng was slightly greater than that in this study.
As described above, We/Wt and Wp/Wt are both linearly related to H/Er, albeit in different directions (positive and negative). This suggests that there is a relationship between We and Wp with a basis in the mechanical properties. To investigate this further, we first plotted We and Wp against H, as shown in Figure 4. The We values were almost independent of H and similar for all tested materials, except for the NiTi alloy, which had a much larger We value than those of the other materials. The We value is determined from the Young’s modulus and elastic strain. Therefore, the larger We value for the NiTi alloy was attributed to the elastic strain contributing to a larger proportion of the same total strain because the Young’s modulus was lower. The Young’s moduli of NiTi, iron/steel (including stainless steel), copper, and tungsten are 60 GPa [32], 190–220 GPa [33], 105–130 GPa [33], and 370 GPa [33], respectively. Meanwhile, as shown in Figure 4, Wp decreased with increasing H, and the values for all materials followed a single curve (except the NiTi alloy, which showed a value below this curve).
Because H is related to the yield strength, work-hardening coefficient, and plastic strain of a material, it is not always an effective parameter for evaluating We. Therefore, we focused on the ratio of Wp/We in order to account for the effect of plastic deformation. Figure 5 shows the relationship between Wp/We and H. The results for iron, steel, and copper fell along a single curve, while the data for the NiTi alloy and tungsten deviated significantly from this curve.
The relationship between H/Er and Wp/We was determined by combining Equations (1) and (2) as follows:
W p W e = ( 1 β ( H E r ) 1 ) α γ = 1 + 0.16 ( H E r ) 1
According to this equation, Wp/We can be represented as a function of H/Er. This relationship is plotted in Figure 6, along with the experimental data for all the materials used herein. Notably, all the materials used herein were accurately represented by Equation (3). Hence, Wp/We was accurately determined from H/Er, where Wp/We and H/Er are both nondimensional values. It is to be noted that the maximum and minimum values of H/Er and Wp/We experimentally obtained in this sturdy were 0.08 and 1.0, respectively. The necessity of these values will be discussed next.

4. Discussion

This section considers the elastic and plastic deformation in the nanoindentation test from a thermodynamic perspective. Taking the specimen as the system, the change in Gibbs free energy (ΔG) due to an external force (i.e., indentation) is considered. Here, ΔG is generally given by the change in Helmholtz free energy (ΔA) and work (ΔW) as follows:
ΔG = ΔA + ΔW
If the applied load in the nanoindentation test is low enough so that only elastic deformation occurs, no indent is left on the specimen surface after the test. In that case, the elastic work energy We is stored in the specimen as reversible work. Therefore, ΔA in Equation (4) corresponds to We. Consequently, the change in Gibbs free energy for loading during elastic deformation, ΔGelastic, can be given as follows:
ΔGelastic = ΔA = We.
If the applied load is large enough for plastic deformation to occur, an indent will be left on the surface after the test, i.e., the sample will undergo elasto-plastic deformation with a change in Gibbs free energy of ΔGelasto-plastic. ΔA is also given as We, and ΔW corresponds to Wp in Equation (4). The sign of Wp is negative when the work done to the system is from an external force. Therefore,
ΔGelasto-plastic = WeWp.
Dividing both terms by We (≠0), Equation (6) can be given as
ΔGelasto-plastic/We = 1 − Wp/We.
If Wp/We < 1, ΔG is positive, and plastic deformation does not occur, because of an increase in the Gibbs free energy. In other words, only elastic deformation occurs during nanoindentation loading, and no indent would remain after unloading. It is therefore concluded that the condition Wp/We ≥ 1 is required for an indent to be observed after the nanoindentation test. Consequently, Equation (8) gives the relationship between hardness and Young’s modulus:
Wp/We = −1.0 + 0.16/(H/Er) ≥ 1
  H / E r     0.08 .
Figure 6 shows those conditions were well satisfied when indents were left on the surface after nanoindentation tests. Here, it should be noted that Wp/We of the NiTi alloy used in this study was 0.99 with H/Er ≈ 0.08, which is nearly the maximum value obtainable. It is also worth noting that Equation (8) reveals the physical meaning of hardness obtained by nanoindentation. That is, when Er is identical, harder materials show smaller values of Wp/We than those of softer ones during nanoindentation under the same applied load.
It is expected that the relationship between H/Er and Wp/We should not be influenced by the microstructure, even though the microstructure has an important influence on each H/Er and Wp/We. This is because the relationship between H/Er and Wp/We is universal, as shown in Figure 6, regardless of the specimen type. That is, even though the mechanical properties of a given material change with the microstructure, the relationship between H/Er and Wp/We should be maintained. The detail of the effects of microstructure will be explored further in our future work. In addition, the relationship might not be valid in alloys with special properties such as superplasticity or shape memory effect, such as that in NiTi alloys. Therefore, it will be interesting to investigate the effects of such special microstructures on the relationship between H/Er and Wp/We.
The findings of this study may have widespread industrial applications. For example, it is known that an increase in H/Er improves the wear resistance of a coating [34]. Based on the relationship between H/Er and Wp/We, it is therefore reasonable that wear resistance will be improved when the value of Wp is decreased. In addition, it is to be stressed that Equation (8) indicates that there is an upper limit of H/Er (or an upper limit of wear resistance) from a thermodynamic perspective.

5. Conclusions

Nanoindentation tests were performed to evaluate the relationships among the indentation hardness and energy components of various metals and alloys. The main conclusions are as follows.
(1)
We/Wt and Wp/Wt are linearly correlated with H/Er with positive and negative relationships, respectively.
(2)
For all the materials used in this study except NiTi alloy, We (Wp) increased (decreased) with increasing H.
(3)
A good correlation between Wp/We and H/Er was found for all materials used in this study, where Wp/We and H/Er are nondimensional values. This emphasizes the physical meaning of hardness obtained by nanoindentation. That is, when Er is identical, harder materials show smaller values of Wp/We than those of softer ones during nanoindentation testing under the same applied load.
(4)
Thermodynamic analyses showed that the indentation mark will be visible after nanoindentation testing when Wp/We ≥ 1, which suggests that the condition H/Er ≤ 0.08 should be true for any material.

Author Contributions

Conceptualization, M.T. and O.F.; methodology, M.Y.; validation, M.Y., M.T. and O.F.; formal analysis, M.T. and O.F.; investigation, M.T. and O.F.; data curation, M.Y.; writing—original draft preparation, O.F.; writing—review and editing, M.T.; visualization, M.Y. and M.T.; funding acquisition, M.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly supported by the Element Strategy Initiative of MEXT, Grant Number JPMXP0112101000.

Acknowledgments

We would like to thank Chiaki Kato (President of Withup Kaiyosha Co., Ltd.) for the insightful discussions on the industrial applications of our results; Hitoshi Kabasawa (President of Nihon Techno Co., Ltd.), Masazumi Oonishi (formerly of Toyota Motor Corporation), Yoshimasa Funakawa (JFE Steel Co., Ltd.), and Noritoshi Sueoka (Daido Steel Co., Ltd.) for providing specimens; and Kensuke Miyahara (National Institute for Materials Science) for his valuable comments when compiling this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, X.; Sun, X.; Raghavan, K.; Comstock, R.J.; Ren, Y. Linking constituent phase properties to ductility and edge stretchability of two DP 980 steels. Mater. Sci. Eng. A 2020, 780, 139176. [Google Scholar] [CrossRef]
  2. Taylor, M.D.; Choi, K.S.; Sun, X.; Matlock, D.K.; Packard, C.E.; Xu, L.; Barlat, F. Correlations between nanoindentation hardness and macroscopic mechanical properties in DP980 steels. Mater. Sci. Eng. A 2014, 597, 431–439. [Google Scholar] [CrossRef]
  3. Mugita, Y.; Aramaki, M.; Yamamoto, M.; Takeuchi, A.; Takeuchi, M.; Yokota, T.; Funakawa, Y.; Furukimi, O. Void nucleation, and growth during tensile deformation of nanoscale precipitated steel and bainitic steel. ISIJ Int. 2019, 59, 1362–1368. [Google Scholar] [CrossRef] [Green Version]
  4. Kato, M. Introduction to the Theory of Dislocations [in Japanese], 5th ed.; Shokabo: Tokyo, Japan, 2005. [Google Scholar]
  5. Kato, M. Selected topics in strength of materials II. Mater. Jpn. 2008, 47, 317–322. [Google Scholar] [CrossRef]
  6. Furukimi, O.; Niigaki, S.; Yamada, N.; Aramaki, M.; Oue, S.; Funakawa, Y. Local deformation energy and void formation behavior in 16%Cr ferritic steel. Tetsu-to-Hagané 2013, 99, 60–69. [Google Scholar] [CrossRef] [Green Version]
  7. Yamamoto, M.; Ochi, R.; Yasuda, K.; Aramaki, M.; Munetoh, S.; Furukimi, O. Effect of γ-phase stability on local deformation energy of α-γ duplex stainless steel. Mater. Trans. 2017, 58, 1379–1385. [Google Scholar] [CrossRef]
  8. Paul, V.; Wakeda, M.; Ameyama, K.; Ota-Kawabata, M.; Ohmura, T. Local deformation behavior of the copper harmonic structure near grain boundaries investigated through nanoindentation. Materials 2021, 14, 5663. [Google Scholar] [CrossRef]
  9. Ohmura, T.; Wakeda, M. Pop-In phenomenon as a fundamental plasticity probed by nanoindentation technique. Materials 2021, 14, 1879. [Google Scholar] [CrossRef]
  10. Liu, Z.; Zhang, J.; He, B.; Zou, Y. High-speed nanoindentation mapping of a near-alpha titanium alloy made by additive manufacturing. J. Mater. Res. 2021, 36, 2223–2234. [Google Scholar] [CrossRef]
  11. Endoh, K.; Ii, S.; Kimura, Y.; Sasaki, T.; Goto, S.; Yokota, T.; Ohmura, T. Effects of grain boundary geometry and boron addition on the local mechanical behavior of interstitial-free (IF) steels. Mater. Trans. 2021, 62, 1479–1488. [Google Scholar] [CrossRef]
  12. Meng, F.H.; Miyamoto, G.; Furuhara, T. Effects of alloying blements on microstructure, hardness and growth rate of compound layer in gaseous-nitrided ferritic alloys. Mater. Trans. 2021, 62, 596–602. [Google Scholar] [CrossRef]
  13. Xiang, S.H.; Yang, X.F.; Liang, Y.X.; Wang, L. Texture evolution and nanohardness in Cu-Nb composite wires. Materials 2021, 14, 5294. [Google Scholar] [CrossRef]
  14. Man, T.H.; Ohmura, T.; Tomota, Y. The effect of boundary or interface on stress-induced martensitic transformation in a Fe-Ni alloy. Mater. Today Commun. 2020, 23, 100896. [Google Scholar] [CrossRef]
  15. Zhu, Z.B.; Huang, H.F.; Liu, J.Z.; Ye, L.F.; Zhu, Z.Y. Nanoindentation Study on the Creep Characteristics and Hardness of Ion-Irradiated Alloys. Materials 2020, 13, 3132. [Google Scholar] [CrossRef]
  16. Pöhl, F.; Hardes, C.; Scholz, F.; Frenzel, J. Orientation-dependent deformation behavior of 316L steel manufactured by laser metal deposition and casting under local scratch and indentation load. Materials 2020, 13, 1765. [Google Scholar] [CrossRef]
  17. Wang, S.B.; Zhao, H.W. Low temperature nanoindentation: Development and applications. Micromachines 2020, 11, 407. [Google Scholar] [CrossRef]
  18. Tanaka, M.; Morikawa, T.; Yoshioka, S.; Takashima, K.; Kaneko, S. Inhomogeneity of Plastic Deformation after Yielding in Low-carbon Martensitic Steels. ISIJ Int. 2022, 62. in press. [Google Scholar] [CrossRef]
  19. Zhang, L.; Liang, M.; Feng, Z.; Zhang, L.; Cao, W.; Wu, G. Nanoindentation characterization of strengthening mechanism in a high strength ferrite/martensite steel. Mater. Sci. Eng. A 2019, 745, 144–148. [Google Scholar] [CrossRef]
  20. Xio, X.; Terentyev, D.; Ruiz, A.; Zinovev, A.; Bakav, A.; Zhurkin, E.E. High temperature nano-indentation of tungsten: Modelling and experimental validation. Mater. Sci. Eng. A 2019, 743, 106–113. [Google Scholar] [CrossRef]
  21. Sun, W.; Jiang, Y.; Sun, G.; Hu, J.; Zhou, T.; Jiang, Z.; Lian, J. Nanoindentation creep behavior and its relation to activation volume and strain rate sensitivity of nanocrystalline Cu. Mater. Sci. Eng. A 2019, 751, 35–41. [Google Scholar] [CrossRef]
  22. Laursen, C.M.; Peter, N.J.; Gerstein, G.; Maier, H.J.; Dehm, G.; Frick, C.P. Influence of Ti3Ni4 precipitates on the indentation-induced two-way shape-memory effect in Nickel-Titanium. Mater. Sci. Eng. A 2020, 792, 139373. [Google Scholar] [CrossRef]
  23. Oliver, W.C.; Pharr, G.M. An improved technique for determining hardness and elastic modulus during load and displacement sensing indentation experiments. J. Mater. Res. 1992, 7, 1564–1583. [Google Scholar] [CrossRef]
  24. Oliver, W.C.; Pharr, G.M. Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding and refinements to methodology. J. Mater. Res. 2004, 19, 3–20. [Google Scholar] [CrossRef]
  25. Sekido, K.; Ohmura, T.; Zhang, L.; Hara, T.; Tsuzaki, K. The effect of interstitial carbon on the initiation of plastic deformation of steels. Mater. Sci. Eng. A 2011, 530, 396–401. [Google Scholar] [CrossRef]
  26. Gouldstone, A.; Koh, H.-J.; Zeng, K.-Y.; Giannakopoulos, A.; Suresh, S. Discrete and continuous deformation during nanoindentation of thin films. Acta Mater. 2020, 48, 2277–2295. [Google Scholar] [CrossRef]
  27. ISO14577-1. Plastic and Elastic Parts of the Indentation Work; A.8; ISO: London, UK, 2015; pp. 22–23. [Google Scholar]
  28. Okoro, A.M.; Machaka, R.; Lephuthing, S.S.; Oke, S.R.; Awotunde, M.A.; Olubambi, P.A. Nanoindentation studies of the mechanical behaviours of spark plasma sintered multiwall carbon nanotubes reinforced Ti6Al4V nanocomposites. Mater. Sci. Eng. A 2019, 765, 138320. [Google Scholar] [CrossRef]
  29. Cheng, Y.-T.; Cheng, C.-M. Relationships between hardness, elastic modulus, and the work of indentation. Appl. Phys. Lett. 1998, 73, 614–616. [Google Scholar] [CrossRef]
  30. Yang, R.; Zhang, T.; Jiang, P.; Bai, Y. Experimental verification and theoretical analysis of the relationships between hardness, elastic modulus, and the work of indentation. Appl. Phys. Lett. 2008, 92, 231906. [Google Scholar] [CrossRef]
  31. ISO14577-1. Indentation Hardness; A.4; ISO: London, UK, 2015; p. 15. [Google Scholar]
  32. Mita, T.; Kawashima, K.; Misumi, M.; Ohkubo, M. Poisson’s ratio and Young’s modulus of TiNi shape memory alloy measured by electromagnetic acoustic transducer. Trans. Jpn. Soc. Mech. Eng. A 2010, 76, 290–295. [Google Scholar] [CrossRef] [Green Version]
  33. The Materials Science Society of Japan. Strength Test Handbook for Metallic Materials; Yokendo: Tokyo, Japan, 1977. [Google Scholar]
  34. Leyland, A.; Matthews, A. On the significance of the H/E ratio in wear control: A nanocomposite coating approach to optimized tribological behaviour. Wear 2000, 246, 1–11. [Google Scholar] [CrossRef]
Figure 1. Definition of the elastic and plastic work energies based on load–displacement curves.
Figure 1. Definition of the elastic and plastic work energies based on load–displacement curves.
Materials 14 07217 g001
Figure 2. Ratio of nanoindentation hardness to reduced modulus (H/Er) as a function of the recovery index (We/Wt). The error bars are drawn with the standard deviation.
Figure 2. Ratio of nanoindentation hardness to reduced modulus (H/Er) as a function of the recovery index (We/Wt). The error bars are drawn with the standard deviation.
Materials 14 07217 g002
Figure 3. Ratio of nanoindentation hardness to reduced modulus (H/Er) as a function of the plasticity index (Wp/Wt). The error bars are drawn with the standard deviation.
Figure 3. Ratio of nanoindentation hardness to reduced modulus (H/Er) as a function of the plasticity index (Wp/Wt). The error bars are drawn with the standard deviation.
Materials 14 07217 g003
Figure 4. Elastic energy (We) and plastic energy (Wp) as a function of nanoindentation hardness (H). The error bars are drawn with the standard deviation.
Figure 4. Elastic energy (We) and plastic energy (Wp) as a function of nanoindentation hardness (H). The error bars are drawn with the standard deviation.
Materials 14 07217 g004
Figure 5. Ratio of plastic energy to elastic energy (Wp/We) as a function of nanoindentation hardness (H). The error bars are drawn with the standard deviation.
Figure 5. Ratio of plastic energy to elastic energy (Wp/We) as a function of nanoindentation hardness (H). The error bars are drawn with the standard deviation.
Materials 14 07217 g005
Figure 6. Ratio of plastic energy to elastic energy (Wp/We) as a function of the ratio of nanoindentation hardness to reduced modulus (H/Er). The error bars are drawn with the standard deviation.
Figure 6. Ratio of plastic energy to elastic energy (Wp/We) as a function of the ratio of nanoindentation hardness to reduced modulus (H/Er). The error bars are drawn with the standard deviation.
Materials 14 07217 g006
Table 1. Chemical composition of tested alloys (mass%).
Table 1. Chemical composition of tested alloys (mass%).
SpecimenCSiMnNiCrMoTiFe
Martensitic steel0.860.160.250.010.04--bal.
Bainitic steel0.090.71.5---0.12bal.
Interstitial-free steel0.0020.0020.14---0.046bal.
Stainless steel (SUS304)0.060.641.089.5218.50--bal.
Stainless steel (SUS316)0.060.561.3612.3417.572.4-bal.
NiTi alloy---55--45-
Table 2. Dimensions (mm) and preparation of nanoindentation test specimens.
Table 2. Dimensions (mm) and preparation of nanoindentation test specimens.
SpecimenShort NamePreparation
Bainitic steelBA Materials 14 07217 i001
Interstitial-free steelIF
Stainless steel (SUS304)S304 Materials 14 07217 i002
Stainless steel (SUS316)S316
Electrodeposited ironEI Materials 14 07217 i003
NiTi alloyNT Materials 14 07217 i004
Single crystal tungstenW Materials 14 07217 i005
Martensitic steelMA Materials 14 07217 i006
CopperCu
Table 3. Average values of H, We, Wp, and Wp/We obtained by nanoindentation tests (Fmax = 9.8 mN; n = 30), along with yield stress (YS) and tensile strength (TS) obtained by tensile testes for various specimens.
Table 3. Average values of H, We, Wp, and Wp/We obtained by nanoindentation tests (Fmax = 9.8 mN; n = 30), along with yield stress (YS) and tensile strength (TS) obtained by tensile testes for various specimens.
Short NameH
(MPa)
Er
(GPa)
We
× 10−10(J)
Wp
× 10−10(J)
Wp/WeYS
(MPa)
TS
(MPa)
MA a9030 ± 379223 ± 5.472.17 ± 0.086.08 ± 0.152.80 ± 0.14--
BA4255 ± 214213 ± 8.181.64 ± 0.0610.1 ± 0.336.16 ± 0.32696765
IF2266 ± 124238 ± 9.931.12 ± 0.0414.6 ± 0.5213.04 ± 0.73188275
EI2613 ± 227215 ± 13.41.31 ± 0.0913.3 ± 0.7210.15 ± 1.05--
S3045348 ± 411205 ± 7.551.87 ± 0.078.79 ± 0.454.70 ± 0.34368658
S3164948 ± 384206 ± 7.651.78 ± 0.079.17 ± 0.535.15 ± 0.42329635
W6006 ± 43.0372 ± 8.821.17 ± 0.028.88 ± 0.077.59 ± 0.14--
Cu b860 ± 38.3121 ± 8.091.21 ± 0.0426.8 ± 0.8122.15 ± 0.9265213
NT4829 ± 24760 ± 1.686.37 ± 0.166.30 ± 0.420.99 ± 0.08-1500
a Vickers hardness reference block HMV700. b Vickers hardness reference block HMV40 annealed at 613 K for 7200 s (Grain diameter ≒ 50 μm).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yamamoto, M.; Tanaka, M.; Furukimi, O. Hardness–Deformation Energy Relationship in Metals and Alloys: A Comparative Evaluation Based on Nanoindentation Testing and Thermodynamic Consideration. Materials 2021, 14, 7217. https://doi.org/10.3390/ma14237217

AMA Style

Yamamoto M, Tanaka M, Furukimi O. Hardness–Deformation Energy Relationship in Metals and Alloys: A Comparative Evaluation Based on Nanoindentation Testing and Thermodynamic Consideration. Materials. 2021; 14(23):7217. https://doi.org/10.3390/ma14237217

Chicago/Turabian Style

Yamamoto, Masayuki, Masaki Tanaka, and Osamu Furukimi. 2021. "Hardness–Deformation Energy Relationship in Metals and Alloys: A Comparative Evaluation Based on Nanoindentation Testing and Thermodynamic Consideration" Materials 14, no. 23: 7217. https://doi.org/10.3390/ma14237217

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