Der Artikel untersucht den Größeneffekt auf den duktilen Bruch der Aluminiumlegierung 2024-T351 und beleuchtet die Herausforderungen und Vorteile von Miniaturprüfmethoden. Es werden die Grenzen dieser Methoden diskutiert, wie Schwierigkeiten bei der Präparation der Proben und der Einfluss des Größeneffekts auf die Materialeigenschaften. Die Studie konzentriert sich außerdem auf die Kalibrierung duktiler Bruchkriterien mittels maschineller Lerntechniken und zeigt das Potenzial fortgeschrittener Rechenmethoden in der Materialwissenschaft auf. Ergänzt wird die Forschung durch experimentelle Daten und Computersimulationen, die eine ganzheitliche Sicht auf das Thema bieten. Der Artikel schließt mit der Betonung der Bedeutung von Miniaturtests in verschiedenen technischen Bereichen und des Potenzials für weitere Fortschritte bei der Materialcharakterisierung.
KI-Generiert
Diese Zusammenfassung des Fachinhalts wurde mit Hilfe von KI generiert.
Abstract
Background
Reliably calibrated criteria are needed for an accurate prediction of fracture of various components. However, there is not always a sufficient amount of material available. Therefore, miniature testing provides an alternative that is researched together with the following calibration of the ductile fracture criteria and investigating the size effect.
Objective
The aim is to design miniature testing equipment and specimens for tensile testing, which covers various stress states. This is supplemented by the small punch test, which has the same specimen thickness, taken from the literature to broaden the portfolio for calibration. The second part deals with conducting the finite element analysis, which provided a basis for the calibration of the phenomenological ductile fracture criterion applicable to crack-free bodies to indicate the crack initiation.
Methods
The steel frame to test thin specimens is designed with optical measurement of deformations. The finite element method is used, within Abaqus and user subroutines, to simulate the tests to obtain the variables needed for the calibration. In addition, the calibration of the criterion using machine learning is explored.
Results
The feasibility of the proposed experimental program is tested on the aluminium alloy 2024-T351. Moreover, the numerical simulations, which showed a good match with experiments in terms of force responses, adds to the knowledge of modelling in the scope of continuum damage mechanics.
Conclusions
The presented results provide a material basis for the aluminium alloy studied on a lower scale, while they broaden the testing possibilities and analyses the calibration strategies for the best failure predictability possible.
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Introduction
Ductile fracture is a common failure mechanism that causes significant issues in many engineering fields. Most of the metals used in civil engineering, transportation and military industry exhibit ductile behaviour to some degree. Better understanding of that behaviour allows further reduction of the weight of structures while maintaining their reliability. The fracture strains for different stress states can be obtained experimentally only for the sheet metal (even though not in all cases), while numerically for cylindrical, tubular and other specimens manufactured from bulk material, for which the cracks can often initiate below the surface, which is not trackable by optics. The experiments can be performed on macroscopic or miniature samples. Miniaturized ones have roots in the nuclear industry, where material is scarce or too dangerous to handle in large quantities. Thus, it is important to find a way to determine the properties of a small amount of material without risking excessive irradiation or the properties of a new material that is being developed. The number of possible miniature testing methods has gradually been extended and has begun to be utilized in various fields of application, providing information on stress–strain behaviour, fatigue, creep or fracture. Although miniature testing methods require less material, it should be noted that they also have limitations and disadvantages related mainly to the size effect and problematic sample preparation. On the other hand, such testing is suitable for estimating the remaining life of structures with the minimum amount of material needed. The repair process after the removal of the material is then either unnecessary or simple, unlike the extraction of standard specimens, which significantly disrupts the structure.
When the size of a material is reduced to a microscale level, the sample often consists of only a few grains in the deformation zone. Then, an individual grain plays a more significant role in the deformation process, which causes the material to behave differently from macroscale parts. This phenomenon is known as the size effect, which affects the properties of a material and makes conventional models and methods no longer applicable in the analysis of microscale specimens [1‐4]. The surface grains are less restricted than the inner grains, leading to lower hardening and lower resistance against deformation and rotation, resulting in lower flow stresses [5, 6]. It is also easier for voids to coalesce and cracks to grow throughout the thickness of microscale materials [2]. The decrease in flow stress with increasing miniaturization can be explained by a surface layer model. Based on that, Lai et al. [5] proposed a new constitutive model to express the stress–strain relationship. In this model, the internal layer of the sample is treated as a polycrystalline material, whereas the surface grains are treated as a single crystal. Then, the fracture strain and energy are much lower in positive stress triaxiality than in negative stress triaxiality regardless of the shape and grain size of the sample [6].
Anzeige
There are many issues in the testing of miniaturized specimens associated with specimen geometry, preparation methods or measuring devices [7, 8]. Therefore, Beerli et al. [9] suggested a semi-automatic miniature testing method to prevent damage to the specimens during handling. The test results can also be affected by the aforementioned size effect, which becomes more significant if the sample thickness was only six to ten times greater than the average grain size [10]. Moreover, the small dimensions of these specimens do not allow the use of standard extensometers for measurements. Thus, optics is used to record the strains [11, 12]. Finally, the specimens have a limited grip section and often require the development of special grippers [7, 13]. The most commonly used testing method for miniaturized specimens is a tension. The specimens generally have rectangular or cylindrical cross-sections. Regarding the device, it must be able to ensure a very slow feed rate because of a very small gauge length. The advantage of the tensile testing method is that it allows for the achievement of different stress states in the crack initiation location by varying the geometry and features of the specimens (Fig. 1), which can also be helpful in investigating the properties of functionally graded materials [14], for example. These tests can be combined with other miniature experiments, such as the Small Punch Test (SPT), to capture a wider range of stress states. SPT is a punching experiment in which a ball or pin is pressed through the centre of a small disk-shaped specimen [15]. Wang et al. [16] used the SPT to evaluate the fracture strain and fracture toughness of the reactor vessel steel. SPT can also be used to obtain experimental creep data, as conducted in a study of AlSi9Cu3 and AZ31 alloys by Andrés et al. [17]. Furthermore, SPT is also a valuable tool for fatigue analysis, as shown by Lewis et al. [18], who carried out experiments on additively manufactured materials. Moreover, the testing devices have been developed to allow elevated temperatures [13], in situ electron microscopy [19, 20], in situ optical microscopy [21, 22] or high strain rates [23]. Last but not least, Watanabe et al. [24] used the in situ miniature testing for subsequent crystal plasticity finite element analysis. A basic overview of the possible miniature specimen geometries is shown in Fig. 1.
The ductile fracture criteria can also be calibrated using butterfly specimens that usually have a small-scale process area. The butterfly specimen allows multiple stress states to be reached by varying the loading angle, as done by Mae et al. [25]. A similar case could be with a cruciform specimen [26].
Rice and Tracey [27] analysed the growth of the spherical void and proposed a ductile fracture criterion dependent on the stress triaxiality
where \(\sigma_{1}\) is the maximum principal stress, \(\sigma_{2}\) is the middle principal stress, \(\sigma_{3}\) is the minimum principal stress and \(\overline{\sigma }\) is the equivalent stress. Decades later, the criteria have been extended by the dependence on the Lode angle \(\theta\) [28‐30]
where \(J_{3}\) is the third invariant of deviatoric stress tensor. Today, machine learning has been incorporated [31, 32], such as Particle Swarm Optimization (PSO) for example [33], which can even replace the criterion itself [34]. Finally, the ductile fracture criteria can be coupled or uncoupled, influencing hardening or not [35]. Coupling the criteria with a hardening results in a weakening (softening) effect on the constitutive law in the scope of Continuum Damage Mechanics (CDM), as in this work. The material parameters of the aforementioned models are generally calibrated using the hybrid experimental–numerical approach, which has also been used in this work. Yao et al. [36] reported on machine learning to provide similar results compared to the hybrid experimental–numerical approach. However, models with a large number of material parameters presented a challenge for the machine learning technique [36]. Regardless of the calibration procedure, the models work with stress triaxiality and Lode angle. This space must be reasonably covered by fracture tests for a reliable calibration of the ductile fracture criteria. Basic fracture tests are highlighted in Fig. 2 on the plane of stress triaxiality and normalized Lode angle
$$\overline{\theta }=1-\frac{6}{\pi }\theta ,$$
(3)
which can consequently be rewritten for the plane stress as
Fig. 2
Representation of stress states under a plane stress condition coinciding with other stress states [37]
In conclusion, miniature testing attracts growing interest that focuses on many aspects. This work aims on the calibration of the ductile fracture criterion using monotonic tensile tests, apart from the small punch test, carried out in a miniature testing device developed in order to show the size effect in ductile fracture for aluminium alloy 2024-T351. It is supplemented by a possible computational modelling in the scope of CDM to show the predictability.
Anzeige
Material Model
The material model can be established after explaining the stress states and quantities. Also, the model is described within this section so that the computational results could be plotted together with experiments in the following parts.
First, the flow curve is estimated. The hardening law is coupled with damage to describe the equivalent stress–equivalent plastic strain relationship in the scope of CDM, obeying the von Mises yield criterion with associated flow rule. A simple power law is chosen as it has a minimum number of material parameters to be calibrated, yet enough for accurate description of the behaviour as shown later. Then, the equivalent stress is
where \(K\) is the strength coefficient, \(\overline{\varepsilon }_{p}\) is the equivalent plastic strain, \(n\) is the strain hardening exponent and \(D\) is the damage parameter, for which a nonlinear accumulation is anticipated according to Xue [38], which was also used later by Papasidero et al. [39] for the same material
where \(\overline{\varepsilon }_{f} \left( {\eta ,\overline{\theta }} \right)\) is the fracture strain as a function of the stress state variables and \(m\) is the damage exponent. The damage parameter reaches the critical value when unity. The smooth specimen tensile test was numerically simulated to obtain the aforementioned material parameters related to plasticity and damage, together with elastic constants, Young’s modulus \(E\) and Poisson’s ratio \(\nu\). All, summarized in Table 1, were found using the trial and error method so that the simulated responses corresponded to those obtained experimentally. This was at the expense of a higher damage exponent than reported by Xue [38] and Papasidero et al. [39], who based it on the low-cycle fatigue and tension–torsion tests ended up with \(m = 1.73\) and \(m = 0.71\), respectively. The weakening exponent is sometimes introduced to the damage parameter in equation (5), which can then artificially decrease the value of the damage exponent. It was not the case in the present study, where a higher damage exponent was a matter of fit in the scope of the coupled damage used. Moreover, it cannot be directly compared with Papasidero et al. [39], who used the uncoupled approach, so the damage exponent did not influence the stress–strain curve contrary to the present analysis. All experiments were simulated within Abaqus/Standard, employing the USDFLD and UHARD user subroutines. Stress triaxiality and normalized Lode angle were defined as state variables within the USDFLD code, while the ductile fracture criterion, damage law and hardening rule were defined in UHARD.
Table 1
Elasticity-, plasticity- and damage-related material parameters
\(E\) [GPa]
\(\nu\) [–]
\(K\) [MPa]
\(n\) [–]
\(m\) [–]
70
0.33
540
0.15
7
Finally, the ductile fracture criterion is defined according to Ganjiani [40] as
where \(\overline{\sigma }^{ref}\) is the reference equivalent stress, \(\eta^{ref}\) is the reference stress triaxiality, \(\overline{\theta }^{ref}\) is the reference normalized Lode angle and \(c_{\eta }\) and \(c_{{\overline{\theta }}}\) are the material parameters related to the stress triaxiality and normalized Lode angle, respectively. Equation (7) has to be transformed from stress space to strain space using a power law considered in equation (5), only without brackets with the damage as
The remaining material parameters were calibrated using an evolutionary algorithm, namely PSO within MATLAB with 10000 particles in the swarm, which found a global minimum (as listed in Table 2) for the four miniaturized fracture tests. The goodness of fit is clear from the following section, where the corresponding experiments and results of the described calculations are explained.
Table 2
Fracture-related material parameters for Ganjiani ductile fracture criterion
\(\eta^{ref}\) [–]
\(\overline{\theta }^{ref}\) [–]
\(\overline{\varepsilon }_{f}^{ref}\) [–]
\(c_{\eta }\) [–]
\(c_{{\overline{\theta }}}\) [–]
0.336
0.992
0.240
0.4007
0.1178
Experiments and Simulations
Material
The material is EN AW 2024-T351 aluminium alloy with copper and magnesium as the main additives with solution heat treatment, stress relief and natural ageing. The grain size allows one to have at least 10 grains across the respective dimension of the designed specimens, which is completely different from studies like [41]. The bulk material is tested, not the sheet, so isotropy is assumed as in the studies used for comparison later [39, 42] and others for the same material [38, 43].
Miniature Test Apparatus
An in-house miniature testing device was designed and built for tensile tests of thin flat specimens. The apparatus consists of the MDI3 NEMA 23 Programmable Motion Control IP20 stepper motor, which drives the upper crossbar guided by two rods. The stationary bottom crossbar is attached to the HBM S2M Force Transducer with a nominal (rated) force of up to 1 kN. The transducer is connected to the HBM CLIP AE101 analogue amplifier that provides the voltage output, which is digitized on a control board using a 16-bit analogue-to-digital converter. The control board based on the ATmega328P microcontroller is responsible for controlling the stepper motor with a screw mechanism that is used to convert the rotary motion into the translational motion of the upper crossbar, from 0.038 to 600 mm/min, while the guide rods have been lubricated for smoother operation. Guide rods are mounted on a stiff structure made from a tool steel as depicted in Fig. 3.
The control board also communicates via USB with an in-house Python code, which provides interface for a manual control, live view, data logging and automated loading of specimens. The code is linked to Alpha 2.1.27, planar Digital Image Correlation (DIC) software from X-Sight. This software records strains via CMOS digital camera Teledyne FLIR BFS-U3-51S5M-C: 5.0 Mpx (2448 × 2048 px), 75 FPS, Sony IMX250, pixel size of 3.45 μm, Mono (Blackfly S), through the telecentric lens AZURE - 6505THM, magnification of 0.5 × , working distance of 65 mm, distortion less than 0.03%, attached to additively manufactured plastic adjustable holder, with lighting provided by OPTIKA CLD-01 LED cold light generator. The DIC is synchronized in real time with the force measurement (with the same sampling frequency). The square-shaped subset size of 101 px (0.697 mm) was used for the line measurement, with an affine subset shape function. Displacement was calculated from the virtual strain gauge (line probe) based on the Lagrange formulation. The interpolant was B-spline and the matching criterion was the normalized sum of squared differences. All optical measurements were performed with respect to the ISO 9513 standard [44] according to the 0.2 accuracy class.
The specimens (Fig. 4) were prepared by means of Electrical Discharge Machining (EDM) using CHMER G32S. Their individual thicknesses were measured with a Schut electronic outside micrometer having a measuring range of 0–25 mm and a resolution of 0.001 mm. The scatter in thickness also contributed to the scatter in the following responses. However, the load was measured during the clamping so that no significant pre-stress was introduced. The pre-loading of 2 N was applied after the clamping before each test similarly as on the commercial testing machines. The alignment of clamped specimens is ensured by simple additively manufactured plastic blocks with grooves. However, the misalignment cannot be fully omitted, which adds to the scatter in the results. This can be minimized by automating the process [9], but it is not easy to quantify the role of misalignment compared to other aspects such as manufacturing, handling and microstructure. It should be noted that Papasidero et al. [39] reported up to 16% difference between individual repeated tests for the same material, but on a macroscale level, and similarly calibrated the deterministic phenomenological ductile fracture criterion as in this case. The mean response (with mean displacement to fracture) are used as a representative sample for the most realistic modelling in this work. However, the three-sigma rule of thumb could be used to obtain conservative results, if needed in industry, for example.
Fig. 4
Smooth, notched and shear specimens with scale (all dimensions in mm)
SPT is another miniature test. It is taken from the literature [45] to supplement the experiments on the previously mentioned specimens for the calibration of the ductile fracture criterion. The ball of 2.5 mm in diameter is made of cemented carbide, while the sensor arm in Fig. 5 belongs to the Zwick multiXtens extensometer, used with the Zwick Z250 Allround-Line, tCII [45]. The ball with punch highlighted in red in Fig. 5 are the only moving parts to deform the specimen, as indicated by the white down-going arrow, while the other parts are stationary.
Fig. 5
Small punch test device with model (all dimensions in mm) [45]
The dog bone specimen had a gauge length of 4 mm and an average thickness of 0.474 mm. The loading rate was 0.05 mm/min. The experimental force–displacement curves served for the calibration of model of elasticity, plasticity and damage as discussed earlier, while the result is presented in Fig. 7 with highlighted onset of fracture. The post-mortem specimen is depicted in Fig. 8 along with the damage contours on the deformed mesh from computation, which predicts the slant fracture. However, the inclined shear band is barely observable in the experiment, in contrast to the notched specimen described later. It should be noted that the models within CDM are generally prone to shear band formation, which does not necessarily occur in reality, as in this case.
Fig. 7
Force–displacement relationship of the smooth specimen – experiments and simulation
The geometry was meshed with 8-node C3D8R brick elements having a size of 0.05 mm in the critical zone, while 4 elements were used through thickness. It should be noted that the convergence analysis was carried out with five different element sizes to ensure that the mesh is fine enough not to influence the results. Then, no remeshing has to be applied. One end of the geometry was fixed, while the displacement was applied to the other end.
Tensile Test of the Notched Specimen
The notched specimen had a gauge length of 2 mm and an average thickness of 0.463 mm. The loading rate was again 0.05 mm/min. The experimental force–displacement curves served to calibrate the Ganjiani ductile fracture criterion described earlier. Experiments with numerical calculation are plotted in Fig. 9, where a moment of assumed crack initiation is also highlighted. The post-mortem specimen is depicted in Fig. 10 with the damage contours on the deformed mesh showing a slight reproduction of the observed slant fracture. This result points to the formation of an inclined shear band, which is typical for metal sheets.
Fig. 9
Force–displacement relationship of the notched specimen – experiments and simulation
Post-mortem notched specimen compared to the contours of the damage parameter on the deformed mesh (approximately in scale)
Bild vergrößern
The geometry was again meshed with 8-node C3D8R brick elements. The element size was 0.05 mm in the critical zone with 4 elements through thickness. The geometry was fixed at one end and displaced at the other end to realize the tension.
Tensile Test of the Shear Specimen
The shear specimen had a gauge length of 6 mm and an average thickness of 0.465 mm. Finally, the loading rate was 0.05 mm/min. The largest scatter in the experimental force–displacement curves in the present study is observable in Fig. 11. These curves served to calibrate the Ganjiani ductile fracture criterion discussed earlier. An instant of assumed crack initiation is highlighted in Fig. 11 with a circle.
Fig. 11
Force–displacement relationship of the shear specimen – experiments and simulation
Once more, the geometry was meshed with 8-node C3D8R brick elements of 0.05 mm size in the critical zone. Again, 4 elements were used through thickness. One end was fixed and the displacement was applied to the other end of the geometry.
The global deformed mesh with detail on the damage contours is given in Fig. 12 with a post-mortem specimen after experiment. It should be noted that Tancogne-Dejean et al. [46] observed crack initiation in the centre of the shear specimen by X-ray synchrotron laminography for the same material, 2024 aluminium alloy, only with different temper – T3.
Fig. 12
Post-mortem shear specimen compared to the global deformed mesh and detail with the contours of the damage parameter (approximately in scale except for the detail)
The deflection of the specimen was measured on a punch using an extensometer, as described earlier, with the punch and ball much stiffer than the SPT specimen, which had an average thickness of 0.502 mm. The loading rate was 1 mm/min for SPT. The force–displacement curves that served to calibrate the Ganjiani ductile fracture criterion are plotted with the computation in Fig. 13, where there is low scatter in the experiments and a satisfactory match with the computation. Again, the moment of crack initiation is highlighted in Fig. 13 by a circle.
Fig. 13
Force–displacement relationship of the SPT specimen – experiments and simulation
The geometry was meshed with 4-node CAX4R bilinear axisymmetric quadrilateral elements with the size of 0.05 mm. The ball and dies were meshed with 2-node RAX2 linear axisymmetric rigid links as they are much stiffer than the specimen. The simulation is compared to the experiment in Fig. 14, where the experiment is represented by the disk cut with a metallographic saw. First, the cut was conventionally grinded and polished. Then, it was polished using the Stuers oxide polishing suspension to be photographed using an inverted metallurgical microscope Olympus GX51. The indicated crack propagation from the calculation is in good agreement with the observation.
Fig. 14
Post-mortem disk specimen cut compared to the contours of the damage parameter on the deformed mesh (approximately in scale)
The Ganjiani ductile fracture criterion was calibrated earlier toward four fracture tests (crack initiation points are highlighted in Figs. 7, 9, 11 and 13) with stress triaxiality and normalized Lode angle averaged with respect to equation (6) as follows
where \(\hat{\varepsilon }_{f}\) is the fracture strain for the respective fracture test. All respective calibration points are summarized in Table 3. These were obtained using previously presented numerical simulations.
Table 3
Averaged state variables and fracture strains for respective fracture tests
Experiment
Averaged stress triaxiality [–]
Averaged normalized Lode angle [–]
Fracture strain [–]
Tensile test of smooth specimen
0.336
0.990
0.236
Tensile test of notched specimen
0.354
0.952
0.229
Tensile test of shear specimen
0.020
0.027
1.020
Small punch test
0.592
–0.119
0.339
The results for miniature specimens (Table 3) are compared in Fig. 15 with the results for standard-sized specimens taken from [42] for the same heat (these specimens were manufactured from the same block of material to eliminate the role of different microstructure) and for standard-sized specimens taken from [39] for the same alloy. Visual comparison reveals no significant size effect for the ductile fracture of 2024-T351, which is consistent with a similar study on a pipeline steel [47]. Only the miniature shear specimen (no such geometry was included in the macroscale batch – the geometry producing the closest stress state was the torsional notched tube) exhibited significantly higher fracture strain, making the fracture envelope higher in generalized shear compared to uniaxial tension condition (Fig. 16a). However, testing shear properties is globally problematic and needs more research [48], to determine whether the geometry of the specimen (sharp corners compared to notched tube), different boundary conditions (tension versus torsion) or other aspects (misalignment) is to blame. Finally, this observation is in correspondence with the results of Li et al. [49], who studied the nickel-based superalloy K418 on a similar microscale. It should be noted that there were no data on the uniaxial compression condition on the microscopic scale. Therefore, the fracture envelope naturally increased with decreasing normalized Lode angle in Fig. 16(a).
Fig. 15
Fracture points for miniature and macroscopic specimens [39, 42]
The Ganjiani ductile fracture criterion based on the previous calibration on miniature specimens is given in Fig. 16(a). This fracture envelope has a similar shape as its macroscale counterpart (Fig. 16(b)) in the region where calibration points are present. The miniature-based fracture envelope along with the cut-off significantly deviates only in the region of normalized Lode angle of –1, where no calibration point is present as discussed above. This can be seen for the plane stress condition in Fig. 17, where the difference is within percent points up to the stress triaxiality of 1/3. It should also be noted that the results for macroscale specimens from [42] are based on an uncoupled approach compared to the coupled approach in this study. However, it should produce similar results regardless of the calibration procedure, confirming that there is no distinctive size effect on ductile fracture within this study in terms of the phenomenological ductile fracture criterion. Also, presented approach is intended for crack-free bodies and primarily for failure initiation, which does not compete with classical fracture mechanics.
Fig. 17
Ganjiani ductile fracture criterion under plane stress based on miniature and standard-sized (macroscale) specimens [42]
Miniature tensile testing becomes important in many engineering disciplines. Therefore, the detailed design of the apparatus for such experiments was developed. Subsequent modelling of elasticity and plasticity reveals a good conformity with the experimental results for the 2024-T351 aluminium alloy, including the addition of a small punch test and comparison with post-mortem specimens. Furthermore, the Ganjiani ductile fracture criterion is calibrated with the help of machine learning. However, artificial intelligence could hardly replace the fracture criteria themselves because of the lack of experimental data. On the other hand, neural networks could be directly trained using more calibrated ductile fracture criteria at once to incorporate the uncertainty of calibration, as some criteria may result in much better fit in terms of mean absolute percentage errors. Finally, the following analysis based on miniature and standard-sized specimens reveals no significant size effect on ductile fracture on a presented level. Therefore, such miniature specimens can be used to characterize the bulk ductile fracture behaviour of 2024-T351 without interrupting the structure with the extraction of a considerable amount of material needed for large-scale testing.
Acknowledgements
This work is an output of the project Computational modelling of ductile fracture of identical wrought and printed metallic materials under ultra-low-cycle fatigue created with financial support from the Czech Science Foundation under the registration no. 23-04724S.
Declarations
Conflict of Interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Ran J, Fu MW, Chan WL (2013) The influence of size effect on the ductile fracture in micro-scaled plastic deformation. Int J Plast 41:65–81
2.
Xu Z, Peng L, Lai XM, Fu MW (2014) Geometry and grain size effects on the forming limit of sheet metals in micro-scaled plastic deformation. Mater Sci Eng A 611:345–353
3.
Fu MW, Chan WL (2011) Geometry and grain size effects on the fracture behavior of sheet metal in micro-scale plastic deformation. Mater Des 32(10):4738–4746
4.
Raulea LV, Goijaerts AM, Govaert LE, Baaijens FPT (2001) Size effects in the processing of thin metal sheets. J Mater Process Technol 115(1):44–48
5.
Lai X, Peng L, Hu P, Lan S, Ni J (2008) Material behavior modelling in micro/meso-scale forming process with considering size/scale effects. Comput Mater Sci 43(4):1003–1009
6.
Wang J, Fu MW, Shi SQ (2019) Influences of size effect and stress condition on ductile fracture behavior in micro-scaled plastic deformation. Mater Des 131(2):69–80
7.
Kumar K, Pooleery A, Madhusoodanan K, Singh RN, Chakravartty RK, Dutta BK, Sinha R (2014) Use of miniature tensile specimen for measurement of mechanical properties. Procedia Eng 86:899–909
8.
Zheng P, Chen R, Liu H, Chen J, Zhang Z, Liu X, Shen Y (2020) On the standards and practices for miniaturized tensile test – A review. Fusion Eng Des 161:112006
9.
Beerli T, Roth ChC, Mohr D (2024) Semi-automatic miniature specimen testing method to characterize the plasticity and fracture properties of metals. Acta Mater 263:119539
10.
Zhang L, Will H, Yar MA, Brown SGR, Lavery NP (2021) The development of miniature tensile specimens with non-standard aspect and slimness ratios for rapid alloy prototyping processes. J Mater Res Technol 15:1830–1843
11.
Rahman F, Ngaile G, Hassan T (2021) Non-contact temperature control and stereo digital image correlation for high-temperature testing of miniature tubular specimens. Rev Sci Instrum 92(11):115104
12.
Rahman F, Ngaile G, Hassan T (2021) Optimized stereo digital image correlation setup for miniature round specimen: framework development and implementation. Opt Lasers Eng 144:106555
13.
Li L, Chan Y-C, Ngaile G, Hassan T (2020) A novel gripper for multiaxial mechanical testing of microtubes at elevated temperatures. Rev Sci Instrum 91(5):055007
14.
Melzer D, Koukolíková M, Rzepa S, Li Y, Džugan J, Brázda M (2023) Influence of interface orientation and surface quality on structural and mechanical properties of SS316L/IN718 block produced using directed energy deposition. Surf Interfaces 41:103139
15.
Hafeez F, Husain A (2019) Standardization of diminutive specimen testing techniques for predicting the mechanical properties of metals. AIP Conf Proc 2148(1):030012
16.
Wang ZX, Shi HJ, Lu J, Shi P, Ma XF (2008) Small punch testing for assessing fracture properties of the reactor vessel steer with different thicknesses. Nucl Eng Des 238(12):3186–3193
17.
Andrés D, Lacalle R, Álvarez RA (2016) Creep property evaluation of light alloys by means of the Small Punch test: Creep master curves. Mater Des 96:122–130
18.
Lewis DTS, Lancaster RJ, Jeffs SP, Illesley HW, Davies SJ, Baxter GJ (2019) Characterising the fatigue performance of additive materials using the small punch test. Mater Sci Eng 754:719–727
19.
Rahman F, Ngaile G, Hassan T (2019) Development of scanning electron microscope-compatible multiaxial miniature testing system. Meas Sci Technol 30(10):105902
20.
Basu S, Jaya BN, Ganguly S, Dutta M, Samajdar I (2023) Novel miniature in situ hole expansion test coupled with microscopic digital image correlation. Rev Sci Instrum 94(10):105104
21.
Gorji MB, Mohr D (2017) Micro-tension and micro-shear experiments to characterize stress-state dependent ductile fracture. Acta Mater 131:65–76
22.
Li J, Xu J, Guo B, Shan D, Langdon TG (2017) Shear fracture mechanism in micro-tension of an ultrafine-grained pure copper using synchrotron radiation X-ray tomography. Scr Mater 132:25–29
23.
Kim JS, Huh H (2011) Evaluation of the material properties of an OFHC copper film at high strain rates using a micro-testing machine. Exp Mech 51:845–855
24.
Watanabe Y, Matsuno T, Hama T, Matsuda T, Okitsu Y, Hayashi S, Takada K, Naito T (2023) Microstructural ductile fracture analysis of 1180-MPa class martensite-matrix dualphase steel via in-situ tensile test. Tetsu to Hagane 109(6):536–546
25.
Mae H, Teng X, Bai Y, Wierzbicki T (2007) Calibration of ductile fracture properties of a cast aluminum alloy. Mater Sci Eng A 459(1–2):156–166
26.
Gerke S, Valencia FR, Norz R, Volk W, Brünig M (2023) Damage and fracture in thin metal sheets: New biaxial experiments. Adv Indust Manuf Eng 7:100129
27.
Rice JR, Tracey DM (1969) On the ductile enlargement of voids in triaxial stress fields. J Mech Phys Solids 17(3):201–217
28.
Wierzbicki T, Bao Y, Lee Y-W, Bai Y (2005) Calibration and evaluation of seven fracture models. Int J Mech Sci 47(4–5):719–743
29.
Ganjiani M, Homayoufard M (2021) Development of a ductile failure model sensitive to stress triaxiality and Lode angle. Int J Solids Struct 225:111066
30.
Ding H, Zhu T, Wang X, Yang B, Xiao S, Yang G (2023) An uncoupled ductile fracture model considering void shape change and necking coalescence. Eng Fract Mech 292:109612
31.
Shang H, Wang S, Zhou L, Lou Y (2023) Neural network-based ductile fracture model for 5182-O aluminum alloy considering electroplastic effect in electrically-assisted processing. Eng Fract Mech 290:109476
32.
Li X, Roth ChC, Mohr D (2023) Neural network based rate- and temperature-dependent Hosford-Coulomb fracture initiation model. Int J Mech Sci 260:108643
33.
Zhong J, Xu T, Guan K, Zhou B (2016) Determination of ductile damage parameters using hybrid particle swarm optimization. Exp Mech 56:945–955
34.
Schowtjak A, Gerlach J, Muhammad W, Brahme AP, Clausmeyer T, Inal K, Tekkaya AE (2022) Prediction of ductile damage evolution based on experimental data using artificial neural networks. Int J Solids Struct 257:111950
35.
Lee S-Y, Kim J-H, Barlat F, Kim HS (2023) Numerical fracture prediction of martensitic steel by uncoupled and coupled type fracture models. Eng Fract Mech 289:109396
36.
Yao D, Pu S, Li M, Guan Y, Duan Y (2022) Parameter identification method of the semi-coupled fracture model for 6061 aluminium alloy sheet based on machine learning assistance. Int J Solids Struct 254–255:111823
37.
Quach H, Kim J-J, Nguyen DT, Kim YS (2020) Uncoupled ductile fracture criterion considering secondary void band behaviors for failure prediction in sheet metal forming. Int J Mech Sci 169:105297
38.
Xue L (2007) Damage accumulation and fracture initiation in uncracked ductile solids subject to triaxial loading. Int J Solids Struct 44(16):5163–5181
39.
Papasidero J, Doquet V, Mohr D (2015) Ductile fracture of aluminum 2024–T351 under proportional and non-proportional multi-axial loading: Bao-Wierzbicki results revisited. Int J Solids Struct 69–70:459–474
40.
Ganjiani M (2020) A damage model for predicting ductile fracture with considering the dependency on stress triaxiality and Lode angle. Eur J Mech Solids A 83:104048MathSciNet
41.
Mine Y, Hirashita K, Takashima H, Matsuda M, Takashima K (2013) Micro-tension behaviour of lath martensite structures of carbon steel. Mater Sci Eng A 560:535–544
42.
Vobejda R, Šebek F, Kubík P, Petruška J (2022) Solution to problems caused by associated non-quadratic yield functions with respect to the ductile fracture. Int J Plast 154:103301
43.
Khan AS, Liu H (2012) Strain rate and temperature dependent fracture criteria for isotropic and anisotropic metals. Int J Plast 37:1–15
44.
ISO 9513:2012 (2022) Metallic materials: Calibration of extensometer systems used in uniaxial testing. Geneva: International Organization for Standardization
45.
Šebek F, Park N, Kubík P, Petruška J, Zapletal J (2019) Ductile fracture predictions in small punch testing of cold-rolled aluminium alloy. Eng Fract Mech 206:509–525
46.
Tancogne-Dejean T, Roth ChC, Morgeneyer TF, Helfen L, Mohr D (2021) Ductile damage of AA2024-T3 under shear loading: Mechanism analysis through in-situ laminography. Acta Mater 205:116556
47.
Gorji MB, Furmanski J, Mohr D (2021) From macro- to micro-experiments: Specimen-size independent identification of plasticity and fracture properties. Int J Mech Sci 199:106389
48.
Han G, He J, Li S, Lin Z (2024) Simple shear methodology for local structure–property relationships of sheet metals: State-of-the-art and open issues. Prog Mater Sci 143:101266
49.
Li B, Cui Y, Wang S, Gao L, Xu Z, Hou X, Li Y (2024) Experimental and numerical investigation of ductile fracture of K418 on in-situ tensile tests. Fatigue Fract Eng Mater Struct 47(4):1361–1378
Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.