Next Article in Journal
Near-Surface and Bulk Dissolution Behavior of γ′ Precipitates in Nickel-Based VDM® Alloy 780 Studied with In-Situ Lab-Source and Synchrotron X-ray Diffraction
Next Article in Special Issue
Effect of Wall Thickness and Surface Conditions on Creep Behavior of a Single-Crystal Ni-Based Superalloy
Previous Article in Journal
Evaluation of Microstructure and Mechanical Properties of a Ti10Mo8Nb Alloy for Biomedical Applications
Previous Article in Special Issue
VHCF Behavior of Inconel 718 in Different Heat Treatment Conditions in a Hot Air Environment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Approach to Estimate the Fatigue Limit of Steels Based on Conventional and Cyclic Indentation Testing

Institute of Materials Science and Engineering, TU Kaiserslautern, 67663 Kaiserslautern, Germany
*
Author to whom correspondence should be addressed.
Metals 2022, 12(7), 1066; https://doi.org/10.3390/met12071066
Submission received: 17 May 2022 / Revised: 15 June 2022 / Accepted: 18 June 2022 / Published: 22 June 2022

Abstract

:
For a reliable design of structural components, valid information about the fatigue strength of the material used is a prerequisite. As the determination of the fatigue properties, and especially the fatigue limit σw, requires a high experimental effort, efficient approaches to estimate the fatigue strength are of great interest. Available estimation approaches using monotonic properties, e.g., Vickers hardness (HV), and in some cases the cyclic yield strength, only allow a rough estimation of σw. The approaches solely based on monotonic properties lead to substantial deviations of the estimated σw in relation to the experimentally determined fatigue limit as they do not consider the cyclic deformation behavior. In this work, an estimation approach was developed, which is based on a correlation analysis of the fatigue limit σw, HV, and the cyclic hardening potential obtained in instrumented cyclic indentation tests (CIT). For this, eleven conditions from five different low-alloy steels were investigated. The CIT enable an efficient and quantitative determination of the cyclic hardening potential, i.e., the cyclic hardening exponentCHT eII, and thus, the consideration of the cyclic deformation behavior in an estimation approach. In this work, a strong correlation of σw with the product of HV and |eII| was observed. In relation to an existing estimation approach based solely on HV, considering the combination of HV and |eII| enables the estimation of σw with an enormously increased precision.

1. Introduction

Materials used in engineering applications are oftentimes subjected to cyclic loadings. Therefore, the knowledge of cyclic properties, and especially the fatigue limit, is indispensable for a safe and reliable design. To determine the fatigue properties, a high experimental effort is required, i.e., a relatively large number of fatigue tests and specimens. Consequently, methods to reduce this effort are of great industrial and scientific interest. For this purpose, short-time procedures [1,2,3,4,5] and correlations between the fatigue limit and mechanical parameters that can be determined with a low experimental effort (e.g., hardness and ultimate tensile stress (UTS)) [6,7,8] were elaborated.
Hardness tests are one of the simplest methods to characterize the monotonic strength of a material. By investigating six different steels, Garwood et al. [6] showed that Rockwell hardness (HRC) roughly correlates with the fatigue limit σw. Note that the fatigue limit σw discussed in the presented work is defined as the stress amplitude σa that leads to an ultimate number of cycles of 106–107, depending on the material investigated and the test frequency used. According to Murakami [9,10], the linear relationship found by Garwood et al. [6] can be expressed based on the Vickers hardness (HV) using Equation (1) (compare Figure 1), with σw in MPa. This relationship is valid up to a material-specific hardness value, ranging between 400 and 500 HV [6,9,10]. Note that the fatigue limits given in Figure 1 were determined in rotation bending tests and the deviation observed for materials with a higher hardness is partially caused by residual stresses at the surface [11]. As this loading condition results in a stress gradient from the surface to the inner material, an increased influence of residual stresses in relation to uniaxial fatigue testing occurs. In addition to the residual stresses and the loading condition, it must be considered that a higher surface roughness significantly decreases the fatigue limit [9,12].
σw = 1.6 HV ± 0.1 HV
In addition to Equation (1), the ASTM Handbook [13] provides a linear correlation of the fatigue limit and Brinell hardness (HB) for some steels, which is valid for a hardness up to 500 HB. The resulting correlation factor is 1.72, which is relatively close to Equation (1). However, this is only a rough estimation, as it is based on the assumption that the fatigue limit is half of the UTS [13]. Furthermore, the hardness can also be used to estimate the fatigue limit in case of defect-based failure, as shown by the investigations of Murakami [9,10] and Casagrande et al. [14].
In addition to hardness, the tensile properties, i.e., the yield strength σy or the UTS, were also used in different scientific works to assess the fatigue limit [7,8,15,16,17].
Although the abovementioned models provide sufficient estimations for some steels, a reliable prediction of the fatigue limit exclusively based on hardness or tensile properties is not achieved [15,18,19]. This is illustrated by the big scatter band shown in Figure 1, while the deviation between the estimated and experimentally determined fatigue limit increases with increasing hardness [10]. Furthermore, Fleck et al. [15] and Rennert et al. [17] show that the relation between the fatigue limit and yield strength, i.e., the ratio σw/σy, depends on the type and the strength of the respective steel. In general, in these works, an increase in σy leads to a less pronounced growth of σw, and thus, with an increase in σy, the ratio σw/σy decreases. This is explained by Fleck et al. [15] with the cyclic deformation behavior: While steels with a lower strength in the initial state show cyclic hardening, most high-strength steels exhibit cyclic softening. This is also underlined by the work of Lopez and Fatemi [19], who observed for two steels with the same hardness a different fatigue limit caused by a different cyclic deformation behavior. Consequently, for a reliable estimation of the fatigue limit σw, the cyclic deformation behavior must be considered.
Beyond the possibilities of conventional hardness measurement, instrumented indentation testing enables the analysis of the deformation behavior. Thus, other material properties, e.g., the Young’s modulus E [20,21], can be determined, too. Moreover, the cyclic deformation behavior of a material can be characterized using instrumented cyclic indentation tests (CIT) [22,23]. The short-time procedure PhyBaLCHT, which is based on CIT and enables determining the cyclic hardening potential, was used in [23] to describe the cyclic deformation behavior of differently heat-treated conditions of the low-alloy steel 42CrMo4. The cyclic hardening potential refers to the capacity of a metallic material to increase its strength during cyclic loading, especially via dislocation activities, which counteracts local stress concentrations, e.g., at microstructural inhomogeneities. As demonstrated in [23], the cyclic deformation behavior obtained with this testing approach corresponds to the cyclic deformation curves determined in uniaxial compressive fatigue tests. Furthermore, Kramer et al. [24] showed for the steel 18CrNiMo7-6 that the cyclic hardening potential determined in CIT correlates with the amount of cyclic hardening obtained in uniaxial push-pull fatigue tests.
In our various investigations, it was demonstrated that the cyclic hardening observed in CIT highly depends on different microstructural phenomena, e.g., the size and distribution of precipitates [25,26,27], the dislocation density [28], and the grain size [24]. Moreover, in [26,27], the results obtained in CIT showed a higher sensitivity to microstructural changes than conventional hardness measurements. Accordingly, in another of our own studies [29] on two differently heat-treated Cu-alloyed steels, X0.5CuNi2-2 and X21CuNi2-2, the ranking of the fatigue strength could only be explained by a combined consideration of the hardness and cyclic hardening potential, as conditions with nearly identical hardness showed a significant difference in σw, depending on their cyclic hardening potential.
Based on the results outlined above, PhyBaLCHT is a promising method to improve the estimation of the fatigue limit of steels as it enables consideration of the cyclic deformation behavior, i.e., the cyclic hardening potential. The high potential of this method is especially underlined by the results reported in [29]. Based on that, in the present work, the relationship between the hardness, the cyclic hardening potential obtained in CIT, and the fatigue limit σw was analyzed. For this purpose, different types of low-alloy steels with a bcc lattice structure were investigated. In addition to existing data from [29] for Cu-alloyed steels and [30,31] for different batches of C50E, hardness measurements, CIT, and fatigue tests were performed on differently heat-treated 42CrMo4 and a batch of bainitic 100CrMnSi6-4.

2. Materials and Methods

2.1. Materials

For the correlation analysis presented in this work, which is the basis for an improved estimation of the fatigue limit σw, three differently heat-treated batches of 42CrMo4, one batch of 100CrMnSi6-4, five batches of differently heat-treated Cu-alloyed steels with two different Cu contents, and two batches of C50E (railway steel R7) were used. In summary, the presented analyses are based on eleven conditions of five different low-alloy steels.
As a condition with a relatively high hardness, a bainitic 100CrMnSi6-4 with the chemical composition given in Table 1 was investigated. To generate a bainitic microstructure, the material was austenitized at 880 °C for 30 min and subsequently bainitized at 220 °C for 8 h, resulting in a retained austenite fraction of 20 vol.-%.
The 42CrMo4 conditions investigated comprise one normalized as well as two quenched and differently tempered conditions. As these variants are from the identical batch investigated in [23], a more detailed description of the microstructure and heat treatment parameters can be found in this preliminary publication. However, the chemical composition is given in Table 1 while the annealing temperatures Ta and times ta are listed in Table 3.
For the Cu-alloyed steels, a detailed description of the chemical compositions, heat treatment parameters, and the resulting microstructure is given in [29,32]. The variants considered in this work differ in their C contents, i.e., 0.005 wt.-% (X0.5CuNi2-2) and 0.21 wt.-% (X21CuNi2-2), and in their annealing times ta, respectively. Moreover, a quenched condition of X0.5CuNi2-2 was analyzed. Please note that these laboratory melts were named with our own labeling based on DIN EN 10027-1 [33] and constitute two low-alloy steels.
Moreover, two batches of C50E, which were extracted from a railway wheel and differ in extraction positions and thus, cooling conditions, were used. As these batches were extracted from a component, no defined heat treatment parameters can be provided. However, the resulting microstructures are characterized in detail in [27,28].

2.2. Mechanical Characterization

To analyze the correlation of the fatigue limit σw and the cyclic deformation behavior obtained in CIT, the respective data is required. Thus, for all material conditions, the cyclic hardening potential was determined using the short-time procedure PhyBaLCHT, which is described in detail in [23,24]. For this, instrumented cyclic indentation tests (CIT) were performed using a Fischerscope H100C device and a Fischerscope HM2000 device (both from Helmut Fischer GmbH, Sindelfingen, Germany), which are both equipped with a Vickers indenter and enable a continuous measurement of the indentation force F and the indentation depth h. For the specimens of the high-strength steel 100CrMnSi6-4, a maximum indentation force of Fmax = 2000 mN was used while the other steels were tested with Fmax = 1000 mN. In CIT, a sinusoidal load function with a frequency of 1/12 Hz and, in total, 10 cycles were used. For each variant, 20 indentations were performed at 2 different polished sections, respectively. Thus, the results of each variant were determined based on 40 CIT, resulting in a high statistical reliability. To exclude any interference between the indentation points, they were placed at a distance of at least five times the indent diagonal, which is in accordance with [34].
From the second cycle on, the continuously measured signals result in an F-h hysteresis (compare Figure 2a). In analogy to the plastic strain amplitude obtained from a stress–strain hysteresis, the half width of the F-h hysteresis at mean loading is defined as the plastic indentation depth amplitude ha,p. Similar to the cyclic deformation curve, the development of ha,p versus the number of cycles N is used to describe the cyclic deformation behavior (compare Figure 2b). The resulting ha,pN curve shows a stabilized slope from the fifth cycle on, which indicates saturation of the macro plastic deformation and a domination of microplasticity. This regime of the ha,pN curve can further be expressed by the power function ha,p II given in Equation (2) [24]:
5 ≤ N ≤ 10: ha,p II = aII·NeII
The exponent eII describes the slope of ha,p II and, thus, of the ha,pN curve in the stabilized regime. As a steeper slope of the ha,pN curve indicates a more pronounced cyclic hardening, eII is used to quantify the cyclic hardening potential of the investigated material and, thus, is called the cyclic hardening exponentCHT. A steeper slope of the curve, which is described by a higher absolute value of the exponent |eII|, indicates a higher cyclic hardening potential, i.e., a higher capacity to counteract local stress concentrations [23,24,26,29]. As already discussed in the introduction, the fatigue strength relates to the cyclic hardening potential and, thus, |eII| is considered for the correlation analyses. Note that the cyclic hardening exponentCHT is not equivalent to the cyclic strain hardening exponent n’ in the Ramberg–Osgood curve as eII is, in contrast to n’, determined in a multi-axial compressive stress state.
In addition to eII, the Vickers hardness was measured for all material variants, as Equation (1) is the starting point of the correlation analyses presented in this work. The Vickers hardness (HV) was measured with a ZwickRoell ZHU 250top universal hardness testing device (ZwickRoell GmbH & Co. KG, Ulm, Germany). In the case of the conditions of the steels X0.5CuNi2-2, X21CuNi2-2, and C50E, a test force of 98.07 N was used (HV 10) while the variants of steel 42CrMo4 and 100CrMnSi6-4 were tested with 294.12 N (HV 30). However, as the Vickers hardness is considered to be independent of the test force, the hardness values can be used adequately. Each HV value was determined by calculating the arithmetic mean of eight measurements. Please note that the values of HV of the steels X0.5CuNi2-2 and X21CuNi2-2 previously published in [32] differ slightly from the ones used in this work because additional hardness measurements were taken into account.
To correlate the values of hardness and cyclic hardening exponentCHT with the fatigue limits, valid data of σw is required for each condition. The fatigue limits of the different variants of X0.5CuNi2-2 and X21CuNi2-2 were obtained by extrapolating the power functions of the Woehler curves given in [29] to the ultimate number of cycles NL = 2 × 106, which is commonly used to determine the fatigue limit of these types of steels. Because of the small lifetime scatter observed for this material, this approach provides a sufficient reliability. For the steel C50E, σw was determined by VHCF experiments with NL = 2 × 108, as presented in [30]. However, because no failure occurred beyond ~106 cycles, the fatigue limit can be defined at NL = 2 × 106.
As for the 100CrMnSi6-4 variant investigated in this work, no fatigue data were available, a Woehler curve was determined for this condition. To this end, fatigue specimens with a polished gauge length and the geometry given in Figure 3a were used. Before polishing, the specimens were hard-turned, which could lead to pronounced residual stresses in the surface. Thus, the process-induced residual stresses were measured at a 40 kV tube voltage, 40 mA tube current and a scanning speed of 0.004°/s using an X-ray diffractometer of the type PANalytical Empyrean (Malvern Panalytical B.V., Almelo, the Netherlands) equipped with a Cu Kα1 tube. The fatigue tests were stress-controlled with a stress ratio R = −1 and a frequency of 10 Hz. For this, a servo-hydraulic testing system of the type Schenck PSA 40 was utilized. The tests were performed at ambient temperature up to NL = 2 × 106.
Similarly, for the variants of 42CrMo4, fatigue experiments were required, as in [23] only single fatigue tests were performed, which did not enable the estimation of σw. The fatigue specimens of 42CrMo4 had a geometry as illustrated in Figure 3b with a polished gauge length. For the quenched and tempered conditions, the fatigue tests were conducted in stress control with the resonant testing device RUMUL TESTRONIC 200 kN (RUSSENBERGER PRÜFMASCHINEN AG, Neuhausen am Rheinfall, Switzerland) and a frequency of approximately 86 Hz while for the normalized variant, the resonant testing device RUMUL TESTRONIC 100 kN (RUSSENBERGER PRÜFMASCHINEN AG, Neuhausen am Rheinfall, Switzerland) and a frequency of approximately 72 Hz were used. All tests were conducted at ambient temperature up to NL = 107 cycles.
To determine the fatigue limits σw of the 42CrMo4 variants, the staircase method, which is described in detail in [35], was used. For each condition, the first experiment of this approach was performed at a stress amplitude σa,start, being in the range of σw. To roughly estimate σa,start, preliminary fatigue tests were performed for each condition. Each following step depended on the results of the prior experiment. If the specimen had reached the maximum number of cycles NL (run-out), the following experiment was conducted at a higher stress amplitude, which was obtained by multiplying the former stress amplitude with the staircase factor dlog. If an experiment led to failure, the stress amplitude was reduced by dividing the previous stress amplitude by dlog. Therefore, all stress amplitudes σn were calculated as described in Equation (3) [35].
The staircase factor dlog depends on the scatter of the tested material and is calculated using the expected or known standard deviation of the whole data set (compare Equation (4)). For the tested conditions, the exponent slog was estimated based on the preliminary fatigue experiments. For the normalized condition and the variant annealed at Ta = 650 °C, slog was set to 0.01 while for the variant annealed at Ta = 550 °C, an slog of 0.03 was chosen. After finishing the test series, a fictitious data point can be set, depending on the result of the last experiment. The fictitious point was set to the next smaller stress amplitude in case of failure or at the next higher stress amplitude if the experiment resulted in a run-out [35]. An exemplary fictitious test series of the staircase method is illustrated in Figure 4.
σn = σa,start × (dlog)n
dlog = 10slog
The fatigue limit σw was calculated based on the number of stress levels i and the number of events fi on each stress level. The lowest stress level was assigned to i = 0 and each higher stress level had the next larger number i + 1. Using Equations (5) and (6), the factors F and A were calculated, which were used to determine σw with Equation (7). Based on this, the fatigue limit σw was determined with a failure probability of 50% [35]:
F = Σfi
A = Σ(i × fi)
σw = σ0 × dlog(A/F)

3. Results

3.1. Determination of the Fatigue Limit of Differently Heat-Treated 42CrMo4 and 100CrMnSi6-4

To obtain statically reliable fatigue limits σw of the different conditions of 42CrMo4, the staircase method was used, leading to the results summarized in Table 2 and Figure 5a. As expected, the normalized condition showed the lowest σw while the condition annealed at Ta = 550 °C yielded the highest fatigue limit.
Considering the failure mechanisms that led to these σw, differences can be observed between the conditions. While for the normalized condition, crack initiation was only observed at surface slip bands, the quenched and tempered conditions also exhibited crack initiation at nonmetallic inclusions. In addition to crack initiation at defects, the condition annealed at Ta = 650 °C also showed crack initiation at slip bands at the surface. Note that for this condition, only one defect with √area = 40 µm was observed, which did not lead to a decrease in the number of cycles to failure Nf. Thus, no relevant influence of microstructural defects on the σw determined is expected for this condition. In contrast to the other conditions, the variant annealed at Ta = 550 °C only showed fatigue crack initiation at nonmetallic inclusions. However, apart from one specimen fractured at a stress amplitude of σa = 450 MPa, which was caused by a defect with √area = 190 µm, only small defect sizes (√area < 90 µm) were observed. As for the small defects, no correlation between the defect size and Nf was obtained, and as the one specimen with a bigger defect has only an insignificant influence on the value σw for Ta = 550 °C, the influence of microstructural defects on σw could also be neglected.
Since for the 100CrMnSi6-4 a smaller number of specimens was available, for this condition, a conventional Woehler curve was determined, which is shown in Figure 5b. The results obtained show a, in comparison with typical high-strength steels, relatively low scatter, which is caused by the small crack-initiating defect sizes, i.e., √area < 25 µm. Consequently, it is expected that also for this variant, the influence of crack-initiating defects on the fatigue limit is weak. Because of the small lifetime scatter, the Woehler curve shown in Figure 5b enables a valid determination of the fatigue limit of this material. For this purpose, the regression line was extended to Nf = 2 × 106 and based on this, the fatigue limit σw is assessed to be approximately 930 MPa (compare Figure 5b).
The fatigue limits of the material conditions of Cu-alloyed steels and the C50E were determined in analogy to 100CrMnSi6-4. For these conditions, only fatigue crack initiation at the surface and no crack-initiating defects were observed and thus, for these variants, the influence of microstructural defects can be neglected, too.

3.2. Relationship between the Fatigue Limit, Hardness, and Cyclic Hardening Potential

To elaborate an improved approach for an estimation of the fatigue limit based on cyclic indentation testing, for all material conditions considered, the fatigue limit σw, the macro hardness (HV), and the cyclic hardening exponentCHT eII were required. Thus, all conditions were characterized by hardness measurements and CIT. The results obtained in indentation testing and the experimentally determined fatigue limits are listed in Table 3, which is the data basis for the correlation analyses shown in the following. Table 3 also contains the standard deviation of HV and the 90% confidence interval of eII, illustrating that the scatter of HV and |eII| is low for all materials.
The σw values shown in Table 3 were determined using fatigue specimens with polished surfaces, leading to an elimination of the influence of the surface roughness. As the specimens were manufactured from bars with diameters significantly bigger than the gauge diameters using low feed rates and manual polishing, thermally and finishing-induced residual stresses are assumed to be relatively low. Additionally, only push-pull fatigue testing (R = −1) was applied and, thus, the whole material volume in the gauge length was loaded homogenously, leading to a less pronounced impact of the residual stresses and surface roughness in relation to other loading conditions, e.g., rotating bending. Consequently, the influence of the residual stresses was neglected.
Note that the specimens of 100CrMnSi6-4 were hard-turned and subsequently polished. Due to the hard-turning, for this condition, pronounced compressive residual stresses in the loading direction (−733 MPa) were measured in the surface-near area, which corresponds to [36]. However, the crack initiation started at defects located at a distance to the surface bigger than 800 µm, which, hence, significantly exceeds the depth of the residual stresses induced by hard-turning (compare [36]). Consequently, the influence of the residual stresses induced by hard-turning on σw can be neglected in this case, too.
In summary, the values of σw given in Table 3 are mainly related to the mechanical properties of the material volume. In this context, the properties of the material volume refer to the integral, global material properties in contrast to local ones, e.g., at the surface. More details about the microstructure and the properties of the material volume are given in the reference mentioned in Table 3. As the results determined in indentation testing are also dominated by the properties of the material volume, the material parameters listed in Table 3 constitute an excellent basis for the approach proposed.
Initially, for all material conditions considered in this work, the relation between the fatigue limit and the Vickers hardness was analyzed. For this purpose, the σw are plotted against the respective HV in Figure 6a. Moreover, in Figure 6a, the commonly used Equation (1) is represented by a straight line, showing a rough correlation with the experimental data. However, some conditions exhibit a rather large deviation from the relation given in Equation (1), especially in the case of 100CrMnSi6-4 and the Cu-alloyed steels. In accordance with [6,9], the relatively hard 100CrMnSi6-4 (706 HV) yields lower σw than expected from Equation (1). However, for the softer Cu-alloyed steels, a significantly higher σw can be observed in relation to Equation (1). It should, thus, be noted that the deviation from this estimation approach is not limited to high-strength conditions.
Comparing eII of these materials, the overestimated 100CrMnSi6-4 exhibits low |eII| while the underestimated Cu-alloyed steels all show strong cyclic hardening and, thus, a high |eII| (see Table 3). Consequently, the relative deviation between the experimentally determined σw and the estimation based on Equation (1) was quantified by σw/(1.6 × HV). These relative deviations were related to |eII|, which is illustrated in Figure 6b. Applying a linear regression of σw/(1.6 × HV) as a function of |eII| results in an impressively good correlation, which can be expressed by Equation (8). Note that Equation (8) contains all three material parameters considered in this work. By rearranging Equation (8), the fatigue limit σw can be expressed as the dependency of HV and |eII| (see Equation (9)):
σw/(1.6 × HV) = 2.946 × |eII| + 0.006
σw = 4.7136 × |eII| × HV + 0.0096 × HV
The summand 0.0096×HV of Equation (9) results from the mathematical regression analysis and represents the section point with the σw-axis. This summand is considered as a mathematical artifact. Note that even for a very high hardness of 1000 HV, this summand would shift the estimated fatigue limit by only 9.6 MPa and consequently, it can be neglected.
Considering Equation (9), the fatigue limit correlates with the product of HV and |eII|. However, the factor (4.7136) given in Equation (9) was calculated based on the estimation from Equation (1), which yields pronounced deviations. Thus, a regression analysis of σw and HV × |eII| was examined only based on the data given in Table 3, which is illustrated in Figure 7.
The diagram shown in Figure 7 demonstrates an excellent correlation between the fatigue limit σw and the product HV × |eII|, which can be expressed mathematically by Equation (10). Thus, this equation can be used to estimate σw based on the Vickers hardness (HV) and the cyclic hardening exponentCHT |eII|. Note that this relation covers data with a high range of HV × eII| (56 to 194) and σw (261 MPa to 930 MPa):
σw = 4.82 × HV × |eII| − 7.18
However, the factor (4.82) and the subtrahend (7.18) of this equation result from regression analysis and, hence, highly depend on the data basis used. Although the data used in the presented work is sufficient to verify the fundamental applicability of the relation, the explicit quantification requires a substantially bigger data basis, especially in the range between σw = 500–900 MPa. Consequently, from the available data, only the general form of this relation can be derived, which is given in Equation (11):
σw = C1 × HV × |eII| + C2

4. Discussion

The regression analysis conducted in Section 3.2 results in a new estimation approach of the fatigue limit σw based on the hardness and the cyclic hardening potential determined in CIT (see Equations (10) and (11)). To analyze the improvement of the estimation achievable with this approach, an additional regression analysis was performed on a reduced data set, without the conditions X0.5CuNi2-2 (ta = 2400 s), X21CuNi2-2 (ta = 2400 s), C50E (position 3), and 42CrMo4 (Ta = 650 °C). As illustrated in Figure 8a, the regression based on Equation (11) and the reduced data set leads to C1 = 4.82 and C2 = −2.34. These coefficients were used to estimate the fatigue limit of the omitted material conditions. Figure 8b demonstrates that the experimentally determined fatigue limits σw, exp. are very close to the estimated fatigue limit σw, est.. To quantitatively compare the estimation based on HV × |eII| with the estimation based solely on HV, the relative differences (Δσw, est.) between σw, exp. and σw, est. estimated with Equations (1) and (11), respectively, were determined for each condition by Equation (12). The resulting deviations are illustrated in Figure 8c:
Δσw, est. = (σw, exp.σw, est.)/σw, est.
As demonstrated by the results shown in Figure 8c, a strong improvement of the fatigue limit estimation is realized by considering the cyclic hardening potential in addition to the hardness. In accordance with the results shown by Fleck et al. [15] and Fatemi and Lopez [19], this improvement is caused by the integration of the cyclic deformation behavior, i.e., the cyclic hardening exponentCHT eII.
Although eII is determined based on a relatively small number of cycles (N = 10), it was shown in preliminary work that this material parameter strongly corresponds to the cyclic deformation curves [23,24] as well as the fatigue strength and lifetime [29,32,37] determined in uniaxial fatigue tests. Moreover, the limited number of cycles enables a fast (about 2 h) analysis of the cyclic deformation characteristics. Consequently, the cyclic hardening exponentCHT obtained with PhyBaLCHT allows the consideration of the cyclic deformation behavior in an efficient approach to estimate the fatigue limit σw.
Note that the estimation approach presented here was only validated for low-alloy steels with several carbon contents and a bcc lattice structure. Moreover, only push-pull fatigue experiments with R = −1 were considered and, thus, this approach needs to be extended to other loading conditions, e.g., rotational bending testing. In addition to this, the data basis is limited to polished surfaces and materials whose fatigue limit is not affected by microstructural defects and residual stresses.
Despite the abovementioned constraints, this approach yields promising results for a wide range of construction materials and, thus, has high potential for industrial and scientific applications. However, a bigger data basis is required, which is an objective of future research.

5. Conclusions and Prospects

In this work, a new approach to estimate the fatigue limit σw of low-alloy steels is presented. For this purpose, the correlation of the fatigue limit σw with the Vickers hardness (HV) and the cyclic hardening potential was analyzed based on eleven material conditions, which comprise five different low-alloy steels. To determine the cyclic hardening potential, instrumented cyclic indentation tests (CIT) were performed under each material condition. By analyzing the cyclic deformation behavior in CIT, the cyclic hardening exponentCHT |eII|, which represents the cyclic hardening potential of the material, was determined.
The results obtained demonstrate that an estimation of σw simply based on hardness measurement is insufficient. However, by considering the combination of HV and the cyclic deformation behavior, i.e., the cyclic hardening potential, the estimation of the fatigue limit σw can be improved enormously. Consequently, the cyclic hardening potential, which represents the cyclic deformation behavior, and can be easily determined by cyclic indentation testing, was used for the first time to obtain a valid estimation of σw, requiring only a small volume of material and a short testing time. Note that for this testing approach, the microstructure and residual stress state of the samples should be equivalent to the specimens or components considered. Thus, this approach has high potential for industrial and scientific applications.
Although the data considered in the presented work is sufficient for verification of the fundamental viability, an extended data base is required for a valid quantification. Moreover, the transferability of this approach to other materials, e.g., aluminum alloys, austenitic steels, or titanium alloys, needs to be investigated [38,39]. Furthermore, the cyclic hardening potential might also be used to improve the estimation of the fatigue limit in the case of defect-based failure. In addition to this, the scatter of the hardness, cyclic hardening potential, and resulting fatigue limit must be integrated [40]. Finally, an analysis of the influence of the loading condition, surface roughness, residual stress state, and microstructural defects, which were excluded in this work by the experimental design, is required, too. These aspects will be objectives of future work.

Author Contributions

Conceptualization, D.G., P.O., P.L., B.B., D.E. and T.B.; methodology, D.G., P.O., P.L., B.B., D.E. and T.B.; formal analysis, D.G.; investigation, D.G. and P.O.; writing—original draft preparation, D.G., P.O. and B.B.; writing—review and editing, P.L., D.E. and T.B.; visualization, D.G., P.O. and P.L.; supervision, B.B. and T.B.; project administration, B.B. and T.B.; funding acquisition, T.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the German Research Foundation (DFG), grant numbers “BE 2350/9-2” (project number 335746905) and “BE 2350/14-1” (project number 420401443), as well as by the priority research activity of Rhineland Palatinate “Advanced Materials Engineering (AME)”.

Data Availability Statement

Not applicable.

Acknowledgments

The authors want to the thank all the co-authors from preliminary work, especially Hendrik Kramer, Marcus Klein, and Peter Starke, which was the basis for the presented approach.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blinn, B.; Beck, T.; Jost, B.; Klein, M.; Eifler, D. PhyBaLSL—Short-time procedure for the determination of the fatigue lifetime of metallic materials under service loading. Int. J. Fatigue 2021, 144, 106060. [Google Scholar] [CrossRef]
  2. Jost, B.; Klein, M.; Beck, T.; Eifler, D. Out-of-Phase TMF lifetime calculation of EN-GJS-600 (ASTM 80-55-06) ductile cast iron based on strain increase tests and evaluation of cyclic deformation behavior in isothermal measuring intervals. Int. J. Fatigue 2018, 117, 274–282. [Google Scholar] [CrossRef]
  3. Starke, P.; Walther, F.; Eifler, D. “PHYBAL” a Short-Time Procedure for a Reliable Fatigue Life Calculation. Adv. Eng. Mater. 2010, 12, 276–282. [Google Scholar] [CrossRef]
  4. Starke, P.; Bäumchen, A.; Wu, H. SteBLife—A new short-time procedure for the calculation of S-N curves and failure probabilities. Mater. Test. 2018, 60, 121–127. [Google Scholar] [CrossRef]
  5. Wu, H.; Bäumchen, A.; Engel, A.; Acosta, R.; Boller, C.; Starke, P. SteBLife—A new short-time procedure for the evaluation of fatigue data. Int. J. Fatigue 2019, 124, 82–88. [Google Scholar] [CrossRef]
  6. Garwood, M.F.; Zurburg, H.H.; Erickson, M.A. Correlation of laboratory tests and service performance. In Interpretation of Tests and Correlation with Service; Garwood, M.F., Zurburg, H.H., Erickson, M.A., Gensamer, M., Burwell, J.T., La Que, F.L., Eds.; American Society for Metals: Cleveland, OH, USA, 1951; pp. 1–77. [Google Scholar]
  7. Pang, J.C.; Li, S.X.; Wang, Z.G.; Zhang, Z.F. General relation between tensile strength and fatigue strength of metallic materials. Mater. Sci. Eng. A 2013, 564, 331–341. [Google Scholar] [CrossRef]
  8. Pang, J.C.; Li, S.X.; Wang, Z.G.; Zhang, Z.F. Relations between fatigue strength and other mechanical properties of metallic materials. Fatigue Fract. Eng. Mater. Struct. 2014, 37, 958–976. [Google Scholar] [CrossRef]
  9. Murakami, Y. Metal Fatigue: Effects of Small Defects and Nonmetallic Inclusions, 1st ed.; Elsevier: Amsterdam, The Netherland, 2002; ISBN 0080440649. [Google Scholar]
  10. Murakami, Y. Material defects as the basis of fatigue design. Int. J. Fatigue 2012, 41, 2–10. [Google Scholar] [CrossRef]
  11. Thompson, S.W.; Parthasarathi, V.; Findley, K.O. A comparison of bending-fatigue properties of surface-induction-hardened SAE 1045 bar steels with and without vanadium and the influence of comparable low-temperature induction-tempering and furnace-tempering treatments. Mater. Sci. Eng. A 2021, 807, 140812. [Google Scholar] [CrossRef]
  12. Novovic, D.; Dewes, R.C.; Aspinwall, D.K.; Voice, W.; Bowen, P. The effect of machined topography and integrity on fatigue life. Int. J. Mach. Tools Manuf. 2004, 44, 125–134. [Google Scholar] [CrossRef]
  13. Mitchell, M.R. Fundamentals of Modern Fatigue Analysis for Design. In ASM Handbook: Volume 19 Fatigue and Fracture; Lampman, S.R., Ed.; ASM International: Materials Park, OH, USA, 1996; pp. 227–249. ISBN 978-1-62708-193-1. [Google Scholar]
  14. Casagrande, A.; Cammarota, G.P.; Micele, L. Relationship between fatigue limit and Vickers hardness in steels. Mater. Sci. Eng. A 2011, 528, 3468–3473. [Google Scholar] [CrossRef]
  15. Fleck, N.A.; Kang, K.J.; Ashby, M.F. Overview no. 112. Acta Metall. Mater. 1994, 42, 365–381. [Google Scholar] [CrossRef]
  16. Liu, J.; Zenner, H. Dauerschwingfestigkeit und zyklisches Werkstoffverhalten. Mater. Werkst. 1989, 20, 327–332. [Google Scholar] [CrossRef]
  17. Rennert, R.; Kullig, E.; Vormwald, M.; Esderts, A.; Luke, M. Analytical Strength Assessment of Components: Made of Steel, Cast Iron and Aluminium Materials: FKM Guideline; 7th Revised Edition 2020; VDMA Verlag GmbH: Frankfurt am Main, Germany, 2021; ISBN 978-3-8163-0745-7. [Google Scholar]
  18. Lee, K.; Song, J. Estimation methods for strain-life fatigue properties from hardness. Int. J. Fatigue 2006, 28, 386–400. [Google Scholar] [CrossRef]
  19. Lopez, Z.; Fatemi, A. A method of predicting cyclic stress–Strain curve from tensile properties for steels. Mater. Sci. Eng. A 2012, 556, 540–550. [Google Scholar] [CrossRef]
  20. Hay, J.L. Introduction to instrumented indentation testing. Exp. Tech. 2009, 33, 66–72. [Google Scholar] [CrossRef]
  21. 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]
  22. Saraswati, T.; Sritharan, T.; Mhaisalkar, S.; Breach, C.D.; Wulff, F. Cyclic loading as an extended nanoindentation technique. Mater. Sci. Eng. A 2006, 423, 14–18. [Google Scholar] [CrossRef]
  23. Blinn, B.; Görzen, D.; Klein, M.; Eifler, D.; Beck, T. PhyBaLCHT—Influence of indentation force on the results of cyclic hardness tests and investigations of comparability to uniaxial fatigue loading. Int. J. Fatigue 2019, 119, 78–88. [Google Scholar] [CrossRef]
  24. Kramer, H.S.; Starke, P.; Klein, M.; Eifler, D. Cyclic hardness test PHYBALCHT—Short-time procedure to evaluate fatigue properties of metallic materials. Int. J. Fatigue 2014, 63, 78–84. [Google Scholar] [CrossRef]
  25. Blinn, B.; Görzen, D.; Fischer, T.; Kuhn, B.; Beck, T. Analysis of the Thermomechanical Fatigue Behavior of Fully Ferritic High Chromium Steel Crofer®22 H with Cyclic Indentation Testing. Appl. Sci. 2020, 10, 6461. [Google Scholar] [CrossRef]
  26. Schwich, H.; Görzen, D.; Blinn, B.; Beck, T.; Bleck, W. Characterization of the precipitation behavior and resulting mechanical properties of copper-alloyed ferritic steel. Mater. Sci. Eng. A 2020, 772, 138807. [Google Scholar] [CrossRef]
  27. Blinn, B.; Winter, S.; Weber, M.; Demmler, M.; Kräusel, V.; Beck, T. Analyzing the influence of a deep cryogenic treatment on the mechanical properties of blanking tools by using the short-time method PhyBaLCHT. Mater. Sci. Eng. A 2021, 824, 141846. [Google Scholar] [CrossRef]
  28. Klein, M.W.; Blinn, B.; Smaga, M.; Beck, T. High cycle fatigue behavior of high-Mn TWIP steel with different surface morphologies. Int. J. Fatigue 2020, 134, 105499. [Google Scholar] [CrossRef]
  29. Görzen, D.; Schwich, H.; Blinn, B.; Bleck, W.; Beck, T. Influence of different precipitation states of Cu on the quasi-static and cyclic deformation behavior of Cu alloyed steels with different carbon contents. Int. J. Fatigue 2020, 136, 105587. [Google Scholar] [CrossRef]
  30. Wagner, V. Wechselverformungsverhalten Hochbeanspruchter Eisenbahnradstähle im Very High Cycle Fatigue (VHCF-) Bereich; Dissertation; TU Kaiserslautern, Lehrstuhl für Werkstoffkunde (WKK): Kaiserslautern, Germany, 2011; ISBN 3-932066-26-X. [Google Scholar]
  31. Wagner, V.; Starke, P.; Kerscher, E.; Eifler, D. Cyclic deformation behaviour of railway wheel steels in the very high cycle fatigue (VHCF) regime. Int. J. Fatigue 2011, 33, 69–74. [Google Scholar] [CrossRef]
  32. Görzen, D.; Schwich, H.; Blinn, B.; Song, W.; Krupp, U.; Bleck, W.; Beck, T. Influence of Cu precipitates and C content on the defect tolerance of steels. Int. J. Fatigue 2021, 144, 106042. [Google Scholar] [CrossRef]
  33. DIN EN 10027-1; DIN Deutsches Institut für Normung e. V. Designation Systems for Steels—Part 1: Steel Names. Beuth Verlag GmbH: Berlin, Germany, 2017.
  34. DIN EN ISO 14577-1; DIN Deutsches Institut für Normung e. V. Metallic Materials—Instrumented Indentation Test for Hardness and Materials Parameters—Part 1: Test Method. Beuth Verlag GmbH: Berlin, Germany, 2015.
  35. DIN 50100; DIN Deutsches Institut für Normung e. V. Load Controlled Fatigue Testing: Execution and Evaluation of Cyclic Tests at Constant Load Amplitudes on Metallic Specimens and Components. Beuth Verlag GmbH: Berlin, Germany, 2021.
  36. Ostermayer, P.; Ankener, W.; Blinn, B.; Smaga, M.; Eifler, D.; Beck, T. Analysis of the Subsurface Volume of Differently Finished AISI 52100 by Cyclic Indentation and X-Ray Diffraction. Steel Res. Int. 2021, 92, 2100253. [Google Scholar] [CrossRef]
  37. Blinn, B.; Krebs, F.; Ley, M.; Teutsch, R.; Beck, T. Determination of the influence of a stress-relief heat treatment and additively manufactured surface on the fatigue behavior of selectively laser melted AISI 316L by using efficient short-time procedures. Int. J. Fatigue 2020, 131, 105301. [Google Scholar] [CrossRef]
  38. Niu, X.; Zhu, S.-P.; He, J.-C.; Liao, D.; Correia, J.A.; Berto, F.; Wang, Q. Defect tolerant fatigue assessment of AM materials: Size effect and probabilistic prospects. Int. J. Fatigue 2022, 160, 106884. [Google Scholar] [CrossRef]
  39. Niu, X.-P.; Wang, R.-Z.; Liao, D.; Zhu, S.-P.; Zhang, X.-C.; Keshtegar, B. Probabilistic modeling of uncertainties in fatigue reliability analysis of turbine bladed disks. Int. J. Fatigue 2021, 142, 105912. [Google Scholar] [CrossRef]
  40. Chausov, M.; Pylypenko, A.; Maruschak, P.; Menou, A. Phenomenological Models and Peculiarities of Evaluating Fatigue Life of Aluminum Alloys Subjected to Dynamic Non-Equilibrium Processes. Metals 2021, 11, 1625. [Google Scholar] [CrossRef]
Figure 1. Fatigue limit as a function of hardness [6], based on [9,10].
Figure 1. Fatigue limit as a function of hardness [6], based on [9,10].
Metals 12 01066 g001
Figure 2. Schematic of the PhyBaLCHT procedure: (a) Fh hysteresis and (b) ha,pN curve [26].
Figure 2. Schematic of the PhyBaLCHT procedure: (a) Fh hysteresis and (b) ha,pN curve [26].
Metals 12 01066 g002
Figure 3. Fatigue specimen geometry of (a) 100CrMnSi6-4 and (b) 42CrMo4.
Figure 3. Fatigue specimen geometry of (a) 100CrMnSi6-4 and (b) 42CrMo4.
Metals 12 01066 g003
Figure 4. Schematic of the staircase method according to [35].
Figure 4. Schematic of the staircase method according to [35].
Metals 12 01066 g004
Figure 5. (a) Results obtained with the staircase method for different variants of 42CrMo4 and (b) the Woehler curve of 100CrMnSi6-4.
Figure 5. (a) Results obtained with the staircase method for different variants of 42CrMo4 and (b) the Woehler curve of 100CrMnSi6-4.
Metals 12 01066 g005
Figure 6. (a) Relationship of σw as a function of HV and (b) the deviation of the experimental determined to the estimated fatigue limit σw/(1.6 × HV) plotted versus |eII|.
Figure 6. (a) Relationship of σw as a function of HV and (b) the deviation of the experimental determined to the estimated fatigue limit σw/(1.6 × HV) plotted versus |eII|.
Metals 12 01066 g006
Figure 7. Relationship of σw as a function of HV×|eII|.
Figure 7. Relationship of σw as a function of HV×|eII|.
Metals 12 01066 g007
Figure 8. (a) Regression analysis of σw as a function of HV × |eII| for selected material conditions; (b) verification of the regression analysis by selected material conditions; (c) relative deviation of σw experimentally determined to the estimated fatigue limit based on HV and HV × |eII|, respectively.
Figure 8. (a) Regression analysis of σw as a function of HV × |eII| for selected material conditions; (b) verification of the regression analysis by selected material conditions; (c) relative deviation of σw experimentally determined to the estimated fatigue limit based on HV and HV × |eII|, respectively.
Metals 12 01066 g008
Table 1. Chemical compositions of 42CrMo4 [23] and 100CrMnSi6-4.
Table 1. Chemical compositions of 42CrMo4 [23] and 100CrMnSi6-4.
Element in wt.-%CSiMnPSCrMoNiCuAl
100CrMnSi6-40.930.591.090.010.011.46-0.030.01<0.001
42CrMo40.370.200.760.0020.0271.020.160.200.190.02
Table 2. Parameters of the staircase method F, A, and fatigue limit σw of the 42CrMo4 conditions.
Table 2. Parameters of the staircase method F, A, and fatigue limit σw of the 42CrMo4 conditions.
Material ConditionFAσw in MPa
normalized1011261
Ta = 650 °C, ta = 2 h810402
Ta = 550 °C, ta = 2 h1019479
Table 3. Fatigue limit σw, Vickers hardness (HV), and cyclic hardening exponentCHT |eII| of the investigated steels.
Table 3. Fatigue limit σw, Vickers hardness (HV), and cyclic hardening exponentCHT |eII| of the investigated steels.
MaterialConditionσw in MPaHV|eII|References
X0.5CuNi2-2quenched400182 ± 30.453 ± 0.027[29,32]
Ta = 550 °C, ta = 120 s460205 ± 20.463 ± 0.022
Ta = 550 °C, ta = 2400 s510207 ± 30.537 ± 0.023
X21CuNi2-2Ta = 550 °C, ta = 120 s480257 ± 30.383 ± 0.024[29,32]
Ta = 550 °C, ta = 2400 s530263 ± 30.429 ± 0.023
C50Eposition 1370275 ± 30.290 ± 0.007[30]
position 3340260 ± 40.289 ± 0.008
42CrMo4normalized261192 ± 60.292 ± 0.010[23]
Ta = 650 °C, ta = 2 h402286 ± 20.292 ± 0.009
Ta = 550 °C, ta = 2 h479364 ± 50.270 ± 0.007
100CrMnSi6-4-930709 ± 50.274 ± 0.007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Görzen, D.; Ostermayer, P.; Lehner, P.; Blinn, B.; Eifler, D.; Beck, T. A New Approach to Estimate the Fatigue Limit of Steels Based on Conventional and Cyclic Indentation Testing. Metals 2022, 12, 1066. https://doi.org/10.3390/met12071066

AMA Style

Görzen D, Ostermayer P, Lehner P, Blinn B, Eifler D, Beck T. A New Approach to Estimate the Fatigue Limit of Steels Based on Conventional and Cyclic Indentation Testing. Metals. 2022; 12(7):1066. https://doi.org/10.3390/met12071066

Chicago/Turabian Style

Görzen, David, Pascal Ostermayer, Patrick Lehner, Bastian Blinn, Dietmar Eifler, and Tilmann Beck. 2022. "A New Approach to Estimate the Fatigue Limit of Steels Based on Conventional and Cyclic Indentation Testing" Metals 12, no. 7: 1066. https://doi.org/10.3390/met12071066

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