Skip to main content
Erschienen in: Journal of Materials Science 14/2024

Open Access 27.02.2024 | Processing Bulk Nanostructured Materials

Atomistic modeling of idealized equal channel angular pressing process

verfasst von: Aruna Prakash

Erschienen in: Journal of Materials Science | Ausgabe 14/2024

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Severe plastic deformation (SPD) processes are forming techniques that impose large plastic strains to achieve ultra-fine grained and nanocrystalline microstructures. The refined microstructure results in a significant increase in strength and in many cases without a considerable loss in ductility. Despite their long existence, a comprehensive understanding of the deformation and grain refinement mechanisms during SPD is still lacking. Atomistic simulations that are regularly used to obtain insights into material behavior have rarely been used to study SPD processes. In this work, we use large scale atomistic simulations of the molecular statics/dynamics kind to model the equal channel angular pressing (ECAP) process, a specific SPD process. Three different materials—Al, Ni and Cu—are used to model high, medium and low stacking fault energy materials. Large simulation cells are used to facilitate grain refinement at the nanoscale. The simulations are carefully analyzed in terms of stress–strain behavior, dislocation activity and grain refinement. All samples show the formation of deformation twins at the nanoscale, which generally detwin at increased strains via dislocation twin interactions. A significant number of stair-rod dislocations form via dislocation–dislocation interactions and are present at low angle grain boundaries in the early stages of deformation. Such stair-rod dislocations also result in the formation of stacking fault tetrahedra which are present in significant numbers in all samples. Finally, difficulties in studying grain refinement at very large strains in atomistic simulations are discussed.
Begleitmaterial
Supplementary file 1 (mp4 127541 KB)
Supplementary file 2 (mp4 80985 KB)
Supplementary file 3 (mp4 87543 KB)
Supplementary file 4 (mp4 167375 KB)
Supplementary file 5 (mp4 177221 KB)
Supplementary file 6 (mp4 15866 KB)
Supplementary file 7 (mp4 33206 KB)
Supplementary file 8 (mp4 22315 KB)
Supplementary file 9 (mp4 13630 KB)
Hinweise
Handling Editor: Megumi Kawasaki.

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1007/​s10853-024-09414-w.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Introduction

Severe plastic deformation (SPD) processes have established themselves as prominent forming techniques for producing materials with very high strengths [1, 2]. Typically in any SPD process, large plastic strains are applied on the feedstock in a single pass of the process. The feedstock is subjected to repeated multiple passes, thus allowing for further accumulation of the large plastic strains. As a result, pronounced grain refinement occurs, with grain sizes in the ultra-fine grained (\(<\,1\,\upmu\)m) and nanocrystalline (\(<\,100\) nm) regime. The decreased grain size results in a significant increase in strength—by up to a factor of 8—and is attributed to the Hall–Petch effect [3]. Generally, any increase in strength is offset by a loss in ductility, often presented as the strength-ductility paradox [4]. Nonetheless, a few studies have indeed reported enhanced ductility in SPD processed materials [58]. This combination of high strength and increased ductility in the processed material makes SPD processes very interesting and attractive for many applications.
A proper description of the enhanced mechanical properties of SPD processed materials requires a fundamental understanding of the grain refinement mechanisms leading to ultra-fine grained (UFG) or nanocrystalline (NC) microstructures, and the complex interactions at play in such microstructures. However, as will be evident by the discussion below, details of such microstructure level processes are still missing. Furthermore, atomistic simulation studies that can provide a fundamental understanding of such critical mechanisms have been rare.
Experimental studies have tried to illuminate the deformation and grain refinement mechanisms during SPD. Investigations on single crystals of Al [911], Cu [1215] and Ni [16] have focused on the influence of orientation on the deformation mechanisms and texture evolution. Further studies on polycrystals have tried to elucidate the influence of stacking fault energy (SFE) on the mechanisms [17]. In high and medium SFE materials like Al and Ni, dislocations interact, intersect and tangle with each other to form sub-structures such as cells and cell blocks [18, 19]. With increasing deformation, it is expected that the misorientation between neighboring cells increases. Two types of boundaries are discussed in this regard: geometrically necessary boundaries (GNBs) and incidental dislocation boundaries (IDBs). While GNBs manifest in transmission electron microscopy (TEM) images as parallel extended lamellar structures and are known to form between regions of different strain patterns, IDBs are randomly oriented structures that are thought to form by random trapping of dislocations. Such lamellar boundaries are observed not only in Al [20, 21] and Ni [22] but also in low SFE materials like Cu [23, 24] and Cu-Al alloys [25, 26].
In materials with low SFE, it is expected that deformation twinning plays an important role in grain refinement [17]. Twin boundaries (TBs) act as effective barriers to dislocation motion, forcing pile-up of dislocations and facilitate dislocation TB interactions. The latter ostensibly transform coherent TBs to incoherent TBs and subsequently to high angle grain boundaries. Further emission of partial dislocations result in the formation of secondary twins whose interaction with themselves and grain boundaries (GB), together with nanoscale grain rotation results in the formation of the refined microstructure [27]. Cao et al. [28] suggest a different mechanism in fcc austenite phase of a duplex steel, where grain refinement is still driven by dislocation twin interactions, but via detwinning of the primary twin.
However, finer details on the mechanisms are missing. For example, it is unclear what the characteristics of the cell wall boundaries are, the mechanism that results in cell boundaries transforming into large angle grain boundaries, and if there is a limiting mechanism for the formation of such cells. It is also an open question if grain refinement indeed extends to the nanoscale, i.e., would grain refinement be observable if the initial characteristic length of the sample is already in the nanometer regime.
It is hence evident that the deformation mechanisms during SPD are far from being well-understood and a clear picture is still lacking. Atomistic simulations are ideally placed in this regard. The atomic resolution provided by these techniques allows for tracking the trajectory of individual atoms and has led to unprecedented insights into mechanical behavior of materials [29, 30], leading to the term computational microscopy [31]. Previous studies have been successful in elucidating mechanisms of deformation in NC materials. For instance, it is now well accepted that GBs act as sources and sinks for dislocations in the absence of dislocation generating sources inside grains due to the small grain sizes [32, 33]. Non-equilibrium GBs that can be found in SPD processed materials enhance this effect due to lower activation energy barriers, and furthermore, exhibit reduced resistance to GB sliding and migration [34]. However, rarely have atomistic simulations been used to study SPD processes, or even, deformation behavior at large strains. A recent atomistic study of SPD has raised the possibility GNBs being associated with disclinations [35].
In this work, we perform a large scale atomistic simulations study of the equal channel angular pressing (ECAP) process, which is perhaps the most widely studied SPD process [36, 37]. In ECAP, a billet is passed through a die consisting of two channels which are equal in cross section and meet at an angle at the center of the die (see Fig. 1a). When the billet is pressed through the die, it is subjected to simple shear when it passes through the plane where the two channels meet [38, 39]. As a result, the ECAP process is often idealized via simple shear boundary conditions in modeling studies, and is also the choice of boundary conditions in the current study.
Of particular focus in the current study is the SFE of the material, which is described by the generalized SFE curve of the interatomic potential used. Three different embedded atom method (EAM) interatomic potentials, which have been parameterized for basic material properties of Al, Ni and Cu, are chosen to model high, medium and low SFE materials. The potentials provide a relatively accurate description of both the unstable \(\gamma _\textrm{usf}\) and stable \(\gamma _\textrm{ssf}\) stacking fault energies compared to density functional theory calculations. The study incorporates large computation cells allowing for dislocation loops of significant sizes to form during the simulations. A single pass of ECAP is modeled by subjecting the simulation cell to 200% shear.

Methods

Single crystalline samples used in the current work are generated with a native cube orientation of \(\left\langle {100} \right\rangle\) and \(\left\langle {010} \right\rangle\) parallel to the x and y directions of the simulation box [40]. The lattice constants used correspond to that predicted by the three different interatomic potentials used for the simulations. The samples are then relaxed using the FIRE algorithm [41] in standard molecular statics simulations to a force norm of \(10^{-8}\) eV/Å. Periodic boundary conditions are imposed along all three directions of the simulation box. Two different sample sizes—with 50 and 100 nm side lengths—are used in the study.
The relaxed structures are then equilibrated at 300 K for 20 ps with the microcanonical ensemble in molecular dynamics (MD), and subsequently thermalized at 300 K for 80 ps with the npt ensemble using a Nosé–Hoover thermo-barostat [42] to ensure zero stresses in all box directions. These stress-free structures are then subjected to simple shear at a constant strain rate \(\dot{\epsilon }_{xy}\) with the npt ensemble (see Fig. 1b). Simple shear boundary conditions are used to mimic ideal ECAP deformation. Samples are subjected to two different strain rates of \(\dot{\epsilon }_{xy}=2 \times 10^8\) s\(^{-1}\) and \(\dot{\epsilon }_{xy}=2 \times 10^9\) s\(^{-1}\) in the current study. The samples are deformed to a total shear of 200% to reflect the strain after a single ECAP pass. A stable time step of 2 fs is used in all simulations. The interatomic interactions are modeled via the embedded atom method (EAM) potentials of Mishin et al. [43] for Al and Ni, and Mishin et al. [44] for Cu. All simulations are performed using the atomistic simulation code LAMMPS [45].
The simulations are analysed by writing out snapshots 0.05% strain increment. The open source visualization tool Ovito [46] is used for visualization. Defect structures are identified using the common neighbor analysis (CNA) [47] and polyhedral template matching (PTM) [48] algorithms. Dislocations are characterized by the dislocation extraction algorithm [49] in Ovito, and the interactions are analyzed using the Thompson tetrahedron notation [50].

Identification of grains

To study the formation of new grains during the process, simulation snapshots are subjected to further analysis with a recently developed tool Orisodata [51], which is an unsupervised machine learning algorithm to cluster atoms into grains. Orisodata constitutes a top–down approach toward grain segmentation and differs from the bottom–up approach used in graph-based clustering methods implemented in Ovito [46, 51]. Briefly, Orisodata works as follows: A user-defined initial number of clusters are randomly sampled in the orientation space. Atoms are assigned to individual clusters based on their proximity to the cluster centers. The orientation of each atom obtained via the PTM algorithm [48] is used for this purpose. The proximity to any cluster is defined by the disorientation between the atom and the cluster centroid. Once all atoms have been assigned to individual clusters, the cluster centroids are recalculated. Clusters are then subjected to split and merge procedures: each cluster is split if the disorientation spread is greater than a split-threshold value, and two clusters are merged if the disorientation between the clusters is less than a merge-threshold value. For more details on the working of the algorithm, the reader is referred to Ref. [51].

Simulation results

Macroscopic response

The macroscopic stress–strain response of the different samples is shown in Fig. 2. The elastic response of the samples follows the elastic moduli predicted by the three potentials with Ni exhibiting the stiffest response, followed by Cu and Al. The peak stress of the samples also follows similar trends: for the 50 nm samples, peak stresses of roughly 20, 10 and 5 GPa were observed for Ni, Cu and Al, respectively. Table 1 lists the values of the peak stress in all the samples.
An increase in the sample size to 100 nm decreases the value of the peak stress indicating a size dependent mechanical response, with the strongest size effect seen for Cu, followed by Ni and Al. Almost no size effect is visible in the case of Al. Plasticity is initiated earliest in the case of Cu, followed by Ni and Al. Despite the differences in peak stresses between the 50 and 100 nm samples, the flow stresses observed are almost identical. Interestingly, both Al and Cu predict almost similar values for the flow stress (\(\sim 1\) GPa); for Ni a higher flow stress of approximately 2 GPa is observed.
An increase in strain rate to \(2\times 10^9\) results in an increase in both the peak stress as well as the flow stresses. An increase in around 10–15% is seen in the case of Al and Ni, while in the case of Cu, a significant increase by around 50% is observed.
Table 1
Peak stresses of all samples used in the current work
 
Al (GPa)
Cu (GPa)
Ni (GPa)
50 nm, \(\dot{\epsilon }=2\times 10^8\)
4.6
9.2
20.2
100 nm, \(\dot{\epsilon }=2\times 10^8\)
4.5
6.7
18.7
100 nm, \(\dot{\epsilon }=2\times 10^9\)
4.9
10.1
21.8

Dislocation activity

To understand the plastic deformation activity in the samples, we look into the nucleation and propagation of dislocations and their reactions with each other. For the sake of simplicity, we present the results of the 50 nm sample. We note, nonetheless, that similar trends are visible in the 100 nm samples.
The initial dislocation activity in all samples is driven by the nucleation and propagation of partial dislocation loops, and is initiated slightly before the peak stress is reached. These loops nucleate on the (d) and (c) planes, with Burgers vectors of \(\delta C\) and \(\gamma D\), respectively (see Fig. 3). Both these dislocations belong to systems that exhibit the highest Schmid factor (cf. Table 2). The large simulation box allows for significant expansion of the loop creating a stacking fault in its wake (see Supplementary Fig. 8).
Table 2
Schmid factors for all possible slip systems under simple shear. The Schmid factors are computed according to the procedure outlined in Ref. [52]
(a)-plane
 
(b)-plane
 
BC(a)
0.408
DC(b)
0.0
DB(a)
0.408
CA(b)
0.408
CD(a)
0.0
AD(b)
0.408
D\(\alpha\)
0.236
C\(\beta\)
0.236
B\(\alpha\)
0.471
A\(\beta\)
0.471
C\(\alpha\)
0.236
D\(\beta\)
0.236
(c)-plane
 
(d)-plane
 
AB(c)
0.0
BA(d)
0.0
BD(c)
0.408
AC(d)
0.408
DA(c)
0.408
CB(d)
0.408
B\(\gamma\)
0.236
A\(\delta\)
0.236
D\(\gamma\)
0.471
C\(\delta\)
0.471
A\(\gamma\)
0.236
B\(\delta\)
0.236
In the case of Al, correlated nucleation of additional partial dislocation loops on adjacent atomic planes occurs, firstly on the (d)-plane and subsequently on the (c)-plane, resulting in the formation of deformation twins. The twin on the (d)-plane grows considerably to a thickness of around 15 atomic layers, while the twin on the (c)-plane attains a maximum thickness of around 8 atomic layers.
In the case of Ni, a single twin forms on the (c)-plane, which is also the plane on which dislocation activity is initiated and attains a maximum thickness of around 10 atomic layers.
In the case of Cu too, only a single twin forms, but on the (d)-plane. The twin, however, grows to be only around 3 layers thick. Subsequently, the trailing partial dislocation \(A \delta\) nucleates to remove the twin (see Supplementary Fig. 8). Trailing partials also nucleate to remove other stacking faults nucleated away from the twin.
The twinning mechanism is identical in all cases where twinning is observed. The first few twinning dislocations have the same Burgers vector as the initial partial dislocation that nucleated on the corresponding plane, i.e. \(\delta C\) and \(\gamma D\) on the (d)-plane and (c)-plane, respectively. Subsequently, after a critical twin thickness of at least 3 atomic layers has been achieved, partial dislocations with the other two possible Burgers vectors in the plane are nucleated.
The stress continues to increase until peak stress during this initial phase of dislocation activity on the (d) and (c) planes. Further partial dislocation loops with Burgers vector \(\delta C\) and \(\gamma D\) continue to nucleate on the (d) and (c) planes away from the initial twin. These result in the formation of stacking faults, but do not lead to twinning.
Subsequently, homogeneous dislocation nucleation occurs on the (a) and (b) planes. Here both leading and trailing partial dislocations are nucleated. Concurrently, a significant drop in the bulk shear stress is observed.
The twins in Al remain stable until beyond 60% applied shear. Subsequently, dislocation twin interactions result in detwinning and the twin is completely removed after approximately 100% applied shear (see Supplementary Movie M1). Similar reactions are also observed in the case of Ni (see Supplementary Movie M2); here the single twin that forms on the (c)-plane is completely detwinned after approximately 70% applied shear. In the case of Cu, however, detwinning occurs by the nucleation of the trailing partial dislocation (see Supplementary Movie M2 and Supplementary Fig. 8).
Twins also form in the 100 nm samples; the thickness of the twins remains the same as in the 50 nm samples (see Supplementary Movie M4). With an increase in the deformation rate, the thickness of the twin seemingly decreases (see Supplementary Movie M5). The detwinning behavior, however, is similar to that found in the 50 nm samples. All twins are completely detwinned after approximately 100% applied shear.
The motion and interaction of partial dislocations on different planes result in the formation of a large number of stair-rod dislocations. Two distinct stair-rod products are generally observed in the simulations over all samples. In the initial phase of plasticity, the partial dislocations \(\delta C\) and \(\gamma D\) meet to form:
$$\begin{aligned} \delta C + \gamma D \rightarrow \frac{\delta \gamma }{CD}, \end{aligned}$$
(1)
which is a 1/3 \(\left\langle {100} \right\rangle\) type of stair-rod product (see Fig. 4a). The second type of interaction occurs once homogeneous dislocation nucleation has been initiated on all planes and trailing partial dislocations are able to nucleate. The stair-rod product is of the more common type \(\delta \beta\) and occurs when, for instance, the leading partial dislocations \(\delta A\) and \(\beta A\) gliding on the (d) and (b) planes, respectively:
$$\begin{aligned} \delta A + A \beta \rightarrow \delta \beta , \end{aligned}$$
(2)
resulting in a 1/6 \(\left\langle {110} \right\rangle\) stair-rod product. Figure 4b shows the formation of such a stair-rod dislocation.
The second type of stair-rod product is of particular interest, since such reactions on multiple planes can result in the formation of a stacking fault tetrahedron (SFT). A significant number of such SFTs are found in all the samples, see supplementary Movies M1–M3. The process of formation of such SFTs is as follows. Leading partial dislocations \(\alpha D\), \(D \beta\) and \(\gamma D\) on the (a), (b) and (c)-planes, respectively, form three stair-rod dislocations on the three edges of the SFT (see Fig. 4c):
$$\begin{aligned} \begin{array}{rcl} \alpha D + D \beta &{} \rightarrow &{} \alpha \beta \\ \gamma D + D \beta &{} \rightarrow &{} \gamma \beta \\ \alpha D + D \gamma &{}\rightarrow &{} \alpha \gamma , \end{array} \end{aligned}$$
(3)
resulting in the formation of a SFT that is open on one face. The closure of such a SFT requires a full dislocation with the Burgers vector AB(d) on the (d)-plane. In the concrete example mentioned above, partial dislocations \(\gamma B\) and \(A \gamma\) join to form the constricted dislocation AB(c) which subsequently cross-slips onto the (d)-plane and dissociates into partial dislocations \(\delta B\) and \(A \delta\). These partial dislocations then interact with the trailing partial dislocations on the other three planes as follows:
$$\begin{aligned} \begin{array}{rcl} \delta B + B\alpha &{} \rightarrow &{} \delta \alpha \\ \delta A + A \gamma &{} \rightarrow &{} \delta \gamma \\ \delta A + A \beta &{} \rightarrow &{} \delta \beta , \end{array} \end{aligned}$$
(4)
resulting in complete closure of the SFT. In certain cases, multiple SFTs can be linked with each other as shown in Fig. 4d, where two SFTs are linked via a common edge.

Formation of new grains

To identify the formation of new grains during the deformation process, the Orisodata algorithm is applied onto select snapshots of the simulation. Figure 5 shows the result of the grain identification with Orisodata applied to a snapshot of the Al 50 nm sample deformed to approximately 15% shear. At this strain, 72 clusters are identified in the deformed configuration. Significant number of the identified grain boundaries (GBs) are low angle GBs (see Supplementary Fig. 9) with disorientation of 15\(^\circ\) or less. The high angle GBs are associated with the twinned regions in the sample. Similar trends are observed in Ni and Cu too.
The clustering pattern identified by Orisodata is consistent with the dislocation activity in the snapshot (cf. Fig. 5a, b). The twinned regions are identified as well-defined straight boundaries.
Since clusters are formed by assigning individual atoms one after the other, a well-defined grain like structure may not be always visible. Hence, the clustering pattern can differ from a microstructure seen in experimental micrographs. Regions between clusters with a low angle GB generally display an array of dislocations, as seen in Fig. 5c, for example. A significant portion of dislocations at such boundaries is made up of stair-rod dislocations, which are generally sessile (cf. Fig. 5c).
With increasing deformation, however, the number of grains identified by Orisodata in the respective snapshots decreases drastically. Beyond roughly 25% applied shear, very few grains are identified in the deformed snapshots (see Supplementary Fig. 9).
Figure 6 shows the orientation of clusters 1 and 66 identified in the snapshot as shown in Fig. 5. The evolution of the orientation in selected clusters from the initial state through to the deformed state as shown in Fig. 5, i.e., 15% applied shear, is shown as supplementary movies M6–M9. The initial elastic deformation results in a small orientation change of the cluster, which can be described by an affine transformation of the simulation cell due to the applied shear deformation. Once plasticity is initiated, atoms in the vicinity of regions where partial dislocations have traversed evidence a change in local orientation, and is visible by the deviation from the initial cluster of points in the pole figures. While cluster 1 undergoes purely dislocation mediated slip, cluster 66 forms a part of the twinned region in the sample. As a result, the orientation evolution of the two clusters as seen in Fig. 6 is quite different.

Discussion

In the current work, large scale atomistic simulations have been performed to study the deformation behavior and grain refinement of single crystalline samples subjected to idealized ECAP conditions. Specifically we seek answers to the following questions: (a) Is grain refinement observed during ECAP if the initial size of samples is in the nanometer regime?, and (b) What are the deformation and grain refinement mechanisms in such samples when deformed to large plastic strains?.
Of particular focus, in this study, has been the influence of SFE of the material on the mechanisms. Three different materials—Al, Ni and Cu—have been used in the current study. All three materials display the formation of twins, albeit to different amounts, during the initial stages of deformation. At the atomic scale, the SFE alone is insufficient to assess the formation of twins. The relative ease of twinning can be better explained by the generalized planar fault energy curves as defined by the interatomic potential of the material [53]. Specifically, the competition between the two ratios \(\kappa _1 = \gamma _{\textsc {ssf}}/\gamma _{\textsc {usf}}\) and \(\kappa _2 = \gamma _{\textsc {ssf}}/\gamma _{\textsc {utf}}\), where \(\gamma _{\textsc {ssf}}\) and \(\gamma _{\textsc {usf}}\) are the stable and unstable stacking fault energies, and \(\gamma _{\textsc {utf}}\) is the unstable twin fault energy, determines the ability to twin. If the two ratios are similar, the probability of twinning is high. A higher \(\kappa _1\), on the other hand, suggests that the barrier for nucleation of the trailing partial dislocation is low; thus the motion of full dislocations is preferred over twinning [54].
For the interatomic potentials used in the current work [43, 44], the values for the two ratios are \(\kappa _1^{Al}=0.87\), \(\kappa _1^{Ni}=0.34\), \(\kappa _1^{Cu}=0.28\) and \(\kappa _2^{Al}=0.65\), \(\kappa _2^{Ni}=0.66\), \(\kappa _2^{Cu}=0.66\). It is hence expected that the likelihood of twinning is greater in Cu when compared to Ni and Al. Further factors, like e.g., \(\gamma _{\textsc {ssf}}/\mu b\), where \(\mu\) is the shear modulus and b is the magnitude of the Burgers vector, have previously been used to rationalize dislocation absorption and slip transmission at GBs [5557]. The numerical values for the aforementioned factor are \(3\times 10^{-2}, 1\times 10^{-2}\) and \(7\times 10^{-3}\) for Al, Ni and Cu, respectively, resulting in the same conclusion as with the ratios \(\kappa _1\) and \(\kappa _2\): Twinning is more likely in Cu as compared to Ni and Al.
In the current work, however, we see twins not only in Cu and Ni, but also in Al with the highest amount of twinning in the latter. Furthermore, these twins are nucleated after considerable elastic straining (approx. 13% applied shear) of the initial cube shaped specimen. This suggests that the fault energies computed at 0 K in a strain-free configuration may by insufficient to completely explain the occurrence of twins in a significantly strained sample. Computation of strain dependent fault energy curves for both twins and stacking faults may be required to provide a clearer picture of twin nucleation. Additional factors like local stresses and activation volume are also likely to influence twinning.
Despite the differences in the amount of twinning, the mechanism itself is the same in all cases. Twinning occurs when a series of partial dislocations nucleate on adjacent atomic planes. A mono-layer twin is created when a partial dislocation propagates on the plane adjacent to an existing stacking fault. With the passage of every additional partial dislocation on an adjacent plane, the twin widens. On the basis of the Burgers vector of the individual partial dislocations, two mechanisms have been identified for the formation of twins: monotonic activation of partials (MAP) where the partial dislocations have the same Burgers vector, and random activation of partials (RAP) where the partial dislocations have different Burgers vectors [58]. In polycrystalline materials, MAP is the accepted mechanism in coarse-grained materials, while RAP has been implied as the more plausible mechanism in nanocrystalline materials since a smaller shear needs to be accommodated [59].
Twins in all samples used in the current work form initially via the MAP mechanism. Once a thickness of at least 3 atomic layers has been achieved, the RAP mechanism seemingly takes over. The large shear created by the MAP mechanism can only be accommodated to a certain extent in the sample. Further compatibility of the twin with the applied boundary conditions necessitates the activation of the RAP mechanism, which requires a lesser amount of shear to be accommodated in comparison to the MAP mechanism.
A particular aim of this study is to identify if grain refinement is indeed possible at the nanoscale. SPD processes are an attractive class of forming processes since they result in a refined microstructure in the UFG/NC regime. Previous studies on microstructure evolution during SPD processes have used samples where the characteristic length scale—generally, the grain size—is well above a micrometer in the initial specimen. Repeated processing of such specimens results in a refined grain microstructure in the UFG/NC regime. By contrast, in the current work, the initial characteristic length scale—the simulation cell size—is already in the nanometer regime. Using the Orisodata algorithm, multiple clusters were identified in the deformed state of the specimens of all three materials up to roughly 20% strain. However, the number of clusters identified by the algorithm progressively decreases after about 15% strain. Furthermore, beyond roughly 25% strain, very new clusters are identified in the samples.
Four parameters are available within the Orisodata algorithm to fine tune the cluster identification process. The initial number of clusters chosen has no significant effect on the final number of clusters identified. The split and merge thresholds do influence the final number of clusters, but do not explain the lack of any clusters beyond a certain strain. The minimum grain size, on the other hand, requires careful calibration in light of the very small grain sizes that are likely to form in the samples. In the current study, a minimum grain size of 5 nm was used for all materials and across all snapshots. In view of the increased dislocation activity and the small regions enveloped by dislocation structures, careful calibration of the minimum grain size, together with the split and merge thresholds might be necessary for obtaining a clear picture on the clustering pattern after very large strains.
An additional plausible reason for the lack of identification of any clusters with the Orisodata algorithm lies in the methodology used to calculate the local orientation, which forms the basis for delimiting regions as clusters or grains. In Orisodata, the orientation is obtained via the PTM algorithm, which computes the orientation of an atom by identifying the local neighborhood of each atom and comparing it with templates of local structural types [48]. Only the local structural environment is considered in this calculation. Previous history, particularly plastic events, do not play a role in such a calculation. As a result, the orientation of a region after the traversal of a dislocation is restored to that before the motion of the dislocation. Consequently, the calculated misorientation between regions decreases, resulting in fewer clusters being identified by Orisodata. This plausibility was also verified with the grain segmentation algorithms in Ovito. Both the graph clustering and minimum spanning tree algorithms, which also use the PTM algorithm for obtaining the orientation of atoms, also failed to recognize any clusters in snapshots at increased strain.
One method to overcome this limitation is to use continuum metrics like deformation gradient to obtain the orientation [6062]. Such metrics require the local neighborhood in the current configuration to be mapped to a reference configuration. As a result, the deformation history is automatically accounted for. The change in orientation with respect to the reference configuration can then be obtained by the polar decomposition of the deformation gradient \(\varvec{ F }= \varvec{ R }\varvec{ U }\), where \(\varvec{ R }\) and \(\varvec{ U }\) are the rotation tensors and the stretch tensors, respectively. Figure 7 shows the difference in orientations computed via PTM and the rotation tensor \(\varvec{ R }\). With the PTM algorithm, two distinct orientations are identified corresponding to the original cube orientation and the orientation of the twinned region. The usage of the rotation tensor clearly show a more heterogeneous distribution of orientations in the untwinned regions. Nonetheless, identifying such regions will require careful calibration of the available parameters, particularly the minimum grain size, within Orisodata or other grain segmentation algorithms, and is reserved for another study.
A primary motivation for the current work has been to investigate the atomic level details of deformation and grain refinement mechanisms during ECAP. Perhaps the most interesting of mechanisms found in the current study, in addition to the mechanism of twinning, is the formation of SFT. These are three dimensional defects in the shape of a tetrahedron consisting of six stair-rod dislocations along the edges of the tetrahedron bounded by triangular stacking faults on four equivalent 111 planes. Such defects were first observed in quenched gold by Silcox & Hirsch [63] who suggested that the tetrahedra were formed by the clustering of quenched-in vacancies to produce a Frank dislocation loop with a Burgers vector \(\varvec{ b }= 1/3 \left\langle {111} \right\rangle\), which then dissociates along the three \(\left\langle {110} \right\rangle\) directions in the 111 plane of the loop. This mechanism, also known as the Silcox-Hirsch mechanism, is one of the three mechanisms that have been proposed for the formation of SFT [64]. The second mechanism, also involving vacancies, has been proposed to explain the occurrence of SFT in high stacking fault energy materials like Al [65]; in situ experiments at very high strain rates showed a significant number of SFT forming, but no dislocations were found to propagate. As a result, it was concluded that isolated vacancies and complex vacancy arrays are involved in SFT production. Atomistic simulations have also provided evidence that SFT are among many major defects that may form under such conditions [66, 67].
The third mechanism proposes the formation of SFT without the involvement of vacancies, but with dislocation glide and cross slip, and has been used to explain the formation of SFT in a Ni-based superalloy after low cycle fatigue [68]. In this mechanism, dislocations interact to form a Frank loop which then follows the Silcox-Hirsch mechanism to produce an SFT. The mechanism of the formation of a triangular Frank loop is only deemed feasible if dislocation motion is able to reverse its glide direction, as is possible under fatigue loading conditions. Wang et al. [69] propose an alternative method that does not involve the Slicox–Hirsch mechanism. They suggest a sequence of dislocation interactions and dissociation that may finally lead to the formation of the SFT. These dislocation reactions were critically reviewed and questioned by Loretto et al. [64] who show that such dislocation reactions are energetically unfavorable compared to the Frank loop.
The novel mechanism observed in the current work is different from the previously observed mechanisms in literature. Firstly, the mechanism does not involve the collapse of vacancies. Secondly, and more importantly, the mechanism does involve the formation of a Frank loop. The observed mechanism is driven by partial dislocations forming stair-rods on three edges of the SFT, similar to the mechanism found by Wang et al. [69]. Note that if the trailing partial dislocations on the three faces were to move to the apex of the tetrahedron, the stacking fault and the stair-rod dislocations would be removed. The stacking faults, however, remain stable resulting in the trailing partial dislocations forming the base of the tetrahedron. The fourth face is now closed by a full dislocation whose partials form stair-rods with the trailing partial dislocations of three faces of the SFT. It must be noted that such reactions seem only possible due to the homogeneous dislocation motion on all four 111 planes. Further details on the dislocation reactions, energetics of dislocation dissociation and SFT formation will be reported elsewhere.
The SFT are highly stable defect structures that can considerably influence mechanical properties [70, 71]. In general, the stair-rod dislocations and Lomer–Cottrell locks that form such SFT are sessile structures, and act as obstacles to the motion of other dislocations. In the current work, we observe that beyond 150% applied shear a significant number of SFT have accumulated resulting in a slight hardening in the stress response (cf. Fig. 2). Furthermore, at low angle boundaries between clusters identified by the Orisodata algorithm, a considerable number of stair-rod dislocations and Lomer–Cotrell locks can be found (cf. Fig. 5c). At higher strains, although clusters could not be identified with Orisodata, dislocation structures identified by Ovito clearly show a network of dislocations of which SFT and stair-rod dislocations form a considerable portion (see Supplementary Fig. 10).
A few comments must be made at this juncture on the idealizations used in the current work, and in general, on the limitations of atomistic simulations. ECAP is modeled here by imposing simple shear boundary conditions. Friction between the billet and die walls, and the complex process parameters that form a part of the real process are simply neglected here. The idealized boundary conditions, however, allow for the usage of simulation cells of 100 nm side lengths, which are considerably large for typical atomistic simulations. The large strains of 200 % required to simulate the process result in computing times of approximately 2.5 Mio cpu hours per sample and simulation. The sample sizes used in the current work essentially denote a compromise between the maximum size usable under the limitations of computing time and the minimum size required for grain refinement to be observed. The single crystalline samples provide the best conditions—in terms of initial size—for grain refinement, and allow for delineating the influence of grain boundaries on the mechanisms if polycrystalline samples were to be used. Both the sample sizes used in the current work are well above the size where a transition of deformation mechanisms with decreasing characteristic size at the nanoscale is observed [72]. Furthermore, the observed mechanisms can be explained solely based on arguments of crystallography, stress state and dislocation theory. It may hence be concluded that the mechanisms hold importance for the experimental situation too, although the observations in the current work have been in nanoscale specimens.
Atomistic simulations of the molecular dynamics kind typically require the usage of very high strain rates that are necessitated by the notorious time limit of MD simulations. In the current work, strain rates of \(10^8\) s\(^{-1}\) and \(10^9\) s\(^{-1}\) have been employed, which are orders of magnitude higher than those used in the real ECAP process. Such high strain rates are a significant constraint in modeling thermally activated processes, and can lead to considerably different quantitative measures like peak and flow stresses. But they do not seem to have much influence on the mechanisms; the reported mechanisms are fairly robust at least in the regime of the two strain rates used in the current study. All mechanisms reported here are associated with dislocation mediated plasticity. Even though the strain rates are high, the overall simulated time is much longer than characteristic time scales of dislocation motion. For instance, it takes only a few picoseconds for a dislocation to move from one obstacle to another or to adjust its velocity in response to a change in stress. By contrast, the total time of the simulated trajectory is in the range of 10 ns, which provides ample time for the underlying dynamics of dislocation behavior. It has been previously suggested that even if flow stress, dislocation density and other rate-dependent measures are likely to be different, the mechanisms themselves can be deemed representative even in the low strain rate regime [31, 73, 74]. In this regard, computational experiments as reported here can be regarded just as quasi-static with respect to the underlying dynamics of dislocation behavior as the comparatively low-rate of straining in laboratory-scale ECAP processes.
Notwithstanding the limitations of the simulation technique, the results of the current work offer additional mechanistic insights into deformation and grain refinement during ECAP, albeit in an idealized version of the process. The formation of large number of stair-rod dislocations and stacking fault tetrahedra is seemingly inherent to severe plastic deformation of materials. Low angle grain boundaries formed during the process contain a significant number of such stair-rod dislocations and SFT.
It is hence evident that atomistic simulations are extremely useful in providing valuable insights into deformation and grain refinement mechanisms during SPD. These computations are particularly attractive since they offer spatial and temporal resolution that is generally not available to experiments [30, 75], and are rightly referred to as computational microscopy [31]. Nevertheless, there are clear reasons as to why such computations have rarely been used to study SPD processes. A single simulation with the 100 nm sample size results in approximately 100 gigabytes of overall processed data, and as mentioned before, requires approximately 2.5 million cpu hours of computing time. Incorporation of larger samples is not merely a question of computing time since modern MD codes scale very well on high performance computing machines. But even with larger simulation cells, the sample dimensions are still likely to be much smaller than the millimeter size specimens used in typical laboratory-scale ECAP processes. The large amount of data generated from such simulations, however, precludes the usage of manual techniques that are currently used to analyze dislocation reactions or to obtain critical stresses. Moving further, significant progress is required on the data curation, analysis and handling front to make such computations amenable for simulating large strains and multiple passes of an SPD process [76]. Automated analysis methods to supplant manual techniques and better algorithms to identify orientations at the atomic scale are necessary to obtain a better picture of microstructure evolution during SPD.
However, we must admit that the limitations of MD simulations pose strong restrictions on expanding our understanding of SPD processes. Simulations with simplified setups as used here can inhibit important mechanisms that might occur in the real world processes [29, 77]. High purity samples without defects or initial dislocation densities are used, a condition rarely satisfied in experimental situations. True rate limiting processes or thermally activated processes are extremely difficult to determine with these computations. The extrapolation of the knowledge to experiments requires a fundamental understanding of the temporal and spatial time scales accessible to such simulations. Since a direct comparison with experiments is seldom possible, the results from simulations must generally be viewed with caution. A complete picture of the deformation mechanism map cannot be obtained by simulations alone. In this regard, such computations must be seen as complementary to experiments; the tractability of the former can be improved with increased synergy with the latter.
The current study has implications not only for understanding the deformation and grain refinement mechanisms during SPD, but for modeling and simulation in general, both for atomic and continuum scale simulations. The usefulness of atomic scale modeling has been discussed above. Continuum scale simulations have been previously used to study various aspects of SPD [78], e.g., process modeling [79, 80], texture and microstructure evolution [8183], grain refinement [84], heterogeneity of stress and strain [85, 86], etc. Such modeling approaches, however, rely on microstructure related information from lower length scales [62]. For instance, crystal plasticity formulations at the mesoscale require details on the deformation mechanisms governing plasticity, as well as quantitative information like resolved shear stresses driving individual mechanisms. Such information—both qualitative and quantitative—can be provided by atomistic simulations to improve the tractability of simulations at the continuum scale. On the other hand, continuum scale simulations can be used to provide feedback in terms of realistic boundary and initial conditions that may not be directly realizable in atomistic simulations [77, 87], increasing the synergy between the modeling approaches.

Conclusions

In this work, large scale atomistic simulations of idealized ECAP process have been performed to study the deformation and grain refinement mechanisms during the process. To this end, simulation cells of two different side lengths—50 and 100 nm—with a native cube orientation and periodic boundary conditions are deformed to 200% under simple shear boundary conditions to mimic ideal ECAP conditions.
The findings of the study can be summarized as follows:
  • All samples show initial elastic deformation to at least 10 % applied shear. Plasticity is initiated very close to the point of peak stress in all samples. Both peak stress and elastic stiffness are the highest for Ni, followed by Cu and finally Al.
  • The initial phase of plasticity (approx. 0.25% applied shear) is dominated by the nucleation of partial dislocation loops creating a stacking fault in their wake as they expand. Partial dislocations are activated on slip systems with the highest Schmid factor.
  • Deformation twins of various thicknesses form in samples of all materials via a combination of the mechanisms monotonic activation and random activation of partials. The twins detwin either via nucleation of trailing partial dislocations or via twin dislocation interactions. The thickness of twins does not change with increased sample size.
  • With increasing plasticity, homogeneous dislocation activity is observed on all four possible {111} planes. Dislocation interactions result in annihilation of dislocation content, as well as the formation of stair-rod dislocations. While the former ensures an almost constant dislocation density and flow stress, the latter leads to the formation of stacking fault tetrahedra.
  • The formation of stacking fault tetrahedra reveals a new mechanism, which neither involves collapse of vacancies nor the formation of a Frank loop. The formation of SFT is driven by the motion and interaction of partial dislocations on four {111} planes.
  • Dislocation interactions critically determine the characteristics of boundaries between newly formed clusters. Such boundaries consist mainly of partial dislocation nodes, stair-rod dislocations and stacking fault tetrahedra.
These results mark the first steps in illuminating atomic scale details of deformation and grain refinement mechanisms during SPD processes. They also provide valuable pointers for future work, which can include realistic initial and boundary conditions to induce strain gradients as well as hardening, providing a possibility of direct comparison with experiments.

Acknowledgements

The author gratefully acknowledges the Center for Information Services and High Performance Computing [Zentrum für Informationsdienste und Hochleistungsrechnen (ZIH)], TU Dresden for providing the computing time for the atomistic simulations in the current work. The author thanks Prof. Laszlo Toth and Prof. Bernhard Eidel for fruitful discussions, and Mr. Raja Havish Seggoju for supporting work.

Declarations

Conflict of interest

That author declares that there is no conflict of interest or any competing interests that could have influenced the work reported in this article.

Ethical approval

Not applicable
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.
Anhänge

Supplementary Information

Below is the link to the electronic supplementary material.
Supplementary file 1 (mp4 127541 KB)
Supplementary file 2 (mp4 80985 KB)
Supplementary file 3 (mp4 87543 KB)
Supplementary file 4 (mp4 167375 KB)
Supplementary file 5 (mp4 177221 KB)
Supplementary file 6 (mp4 15866 KB)
Supplementary file 7 (mp4 33206 KB)
Supplementary file 8 (mp4 22315 KB)
Supplementary file 9 (mp4 13630 KB)
Literatur
1.
Zurück zum Zitat Zehetbauer MJ, Valiev RZ (eds) (2004) Nanomaterials by severe plastic deformation. Wiley, Hoboken Zehetbauer MJ, Valiev RZ (eds) (2004) Nanomaterials by severe plastic deformation. Wiley, Hoboken
2.
Zurück zum Zitat Valiev RZ, Islamgaliev RK, Alexandrov(2000) IV Bulk nanostructured materials from severe plastic deformation, vol 45 Valiev RZ, Islamgaliev RK, Alexandrov(2000) IV Bulk nanostructured materials from severe plastic deformation, vol 45
3.
Zurück zum Zitat Valiev RZ, Estrin Y, Horita Z, Langdon TG, Zechetbauer MJ, Zhu YT (2006) Producing bulk ultrafine-grained materials by severe plastic deformation. JOM 58(4):33–39CrossRef Valiev RZ, Estrin Y, Horita Z, Langdon TG, Zechetbauer MJ, Zhu YT (2006) Producing bulk ultrafine-grained materials by severe plastic deformation. JOM 58(4):33–39CrossRef
5.
Zurück zum Zitat Wang Y, Chen M, Zhou F, Ma E (2002) High tensile ductility in a nanostructured metal. Nature 419(6910):912–915PubMedCrossRef Wang Y, Chen M, Zhou F, Ma E (2002) High tensile ductility in a nanostructured metal. Nature 419(6910):912–915PubMedCrossRef
9.
Zurück zum Zitat Fukuda Y, Oh-ishi K, Furukawa M, Horita Z, Langdon TG (2004) The application of equal-channel angular pressing to an aluminum single crystal. Acta Materialia 52(6):1387–1395CrossRef Fukuda Y, Oh-ishi K, Furukawa M, Horita Z, Langdon TG (2004) The application of equal-channel angular pressing to an aluminum single crystal. Acta Materialia 52(6):1387–1395CrossRef
11.
Zurück zum Zitat Han W, Zhang ZF, Wu S, Li S (2007) Influences of crystallographic orientations on deformation mechanism and grain refinement of al single crystals subjected to one-pass equal-channel angular pressing. Acta Materialia 55(17):5889–5900CrossRef Han W, Zhang ZF, Wu S, Li S (2007) Influences of crystallographic orientations on deformation mechanism and grain refinement of al single crystals subjected to one-pass equal-channel angular pressing. Acta Materialia 55(17):5889–5900CrossRef
12.
Zurück zum Zitat Miyamoto H, Erb U, Koyama T, Mimaki T, Vinogradov A, Hashimoto S (2004) Microstructure and texture development of copper single crystals deformed by equal-channel angular pressing. Philos Mag Lett 84(4):235–243CrossRef Miyamoto H, Erb U, Koyama T, Mimaki T, Vinogradov A, Hashimoto S (2004) Microstructure and texture development of copper single crystals deformed by equal-channel angular pressing. Philos Mag Lett 84(4):235–243CrossRef
13.
Zurück zum Zitat Miyamoto H, Fushimi J, Mimaki T, Vinogradov A, Hashimoto S (2005) Dislocation structures and crystal orientations of copper single crystals deformed by equal-channel angular pressing. Mater Sci Eng A 405(1–2):221–232CrossRef Miyamoto H, Fushimi J, Mimaki T, Vinogradov A, Hashimoto S (2005) Dislocation structures and crystal orientations of copper single crystals deformed by equal-channel angular pressing. Mater Sci Eng A 405(1–2):221–232CrossRef
15.
Zurück zum Zitat Guo T, Wei S, Wang C, Li Q, Jia Z (2019) Texture evolution and strengthening mechanism of single crystal copper during ECAP. Mater Sci Eng A 759:97–104CrossRef Guo T, Wei S, Wang C, Li Q, Jia Z (2019) Texture evolution and strengthening mechanism of single crystal copper during ECAP. Mater Sci Eng A 759:97–104CrossRef
16.
Zurück zum Zitat Skrotzki W, Toth LS, Klöden B, Brokmeier HG, Arruffat-Massion R (2008) Texture after ECAP of a cube-oriented NI single crystal. Acta Materialia 56(14):3439–3449CrossRef Skrotzki W, Toth LS, Klöden B, Brokmeier HG, Arruffat-Massion R (2008) Texture after ECAP of a cube-oriented NI single crystal. Acta Materialia 56(14):3439–3449CrossRef
18.
Zurück zum Zitat Hansen N (2001) New discoveries in deformed metals. Metall Mater Trans A 32:2917–2935CrossRef Hansen N (2001) New discoveries in deformed metals. Metall Mater Trans A 32:2917–2935CrossRef
19.
Zurück zum Zitat Huang X, Tsuji N, Hansen N, Minamino Y (2003) Microstructural evolution during accumulative roll-bonding of commercial purity aluminum. Mater Sci Eng A 340(1–2):265–271CrossRef Huang X, Tsuji N, Hansen N, Minamino Y (2003) Microstructural evolution during accumulative roll-bonding of commercial purity aluminum. Mater Sci Eng A 340(1–2):265–271CrossRef
20.
Zurück zum Zitat Iwahashi Y, Horita Z, Nemoto M, Langdon T (1997) An investigation of microstructural evolution during equal-channel angular pressing. Acta Materialia 45(11):4733–4741CrossRef Iwahashi Y, Horita Z, Nemoto M, Langdon T (1997) An investigation of microstructural evolution during equal-channel angular pressing. Acta Materialia 45(11):4733–4741CrossRef
21.
Zurück zum Zitat Iwahashi Y, Horita Z, Nemoto M, Langdon TG (1998) The process of grain refinement in equal-channel angular pressing. Acta Materialia 46(9):3317–3331CrossRef Iwahashi Y, Horita Z, Nemoto M, Langdon TG (1998) The process of grain refinement in equal-channel angular pressing. Acta Materialia 46(9):3317–3331CrossRef
22.
Zurück zum Zitat Zhang H, Huang X, Hansen N (2008) Evolution of microstructural parameters and flow stresses toward limits in nickel deformed to ultra-high strains. Acta Materialia 56(19):5451–5465CrossRef Zhang H, Huang X, Hansen N (2008) Evolution of microstructural parameters and flow stresses toward limits in nickel deformed to ultra-high strains. Acta Materialia 56(19):5451–5465CrossRef
24.
Zurück zum Zitat Mishra A, Kad B, Gregori F, Meyers M (2007) Microstructural evolution in copper subjected to severe plastic deformation: experiments and analysis. Acta materialia 55(1):13–28CrossRef Mishra A, Kad B, Gregori F, Meyers M (2007) Microstructural evolution in copper subjected to severe plastic deformation: experiments and analysis. Acta materialia 55(1):13–28CrossRef
26.
Zurück zum Zitat Qu S, An XH, Yang HJ, Huang CX, Yang G, Zang QS, Wang ZG, Wu SD, Zhang ZF (2009) Microstructural evolution and mechanical properties of Cu–Al alloys subjected to equal channel angular pressing. Acta Materialia 57(5):1586–1601CrossRef Qu S, An XH, Yang HJ, Huang CX, Yang G, Zang QS, Wang ZG, Wu SD, Zhang ZF (2009) Microstructural evolution and mechanical properties of Cu–Al alloys subjected to equal channel angular pressing. Acta Materialia 57(5):1586–1601CrossRef
27.
Zurück zum Zitat Wang YB, Liao XZ, Zhao YH, Lavernia EJ, Ringer SP, Horita Z, Langdon TG, Zhu YT (2010) The role of stacking faults and twin boundaries in grain refinement of a Cu–Zn alloy processed by high-pressure torsion. Mater Sci Eng A 527(18–19):4959–4966CrossRef Wang YB, Liao XZ, Zhao YH, Lavernia EJ, Ringer SP, Horita Z, Langdon TG, Zhu YT (2010) The role of stacking faults and twin boundaries in grain refinement of a Cu–Zn alloy processed by high-pressure torsion. Mater Sci Eng A 527(18–19):4959–4966CrossRef
29.
Zurück zum Zitat Prakash A, Guénolé J, Wang J, Müller J, Spiecker E, Mills MJ, Povstugar I, Choi P, Raabe D, Bitzek E (2015) Atom probe informed simulations of dislocation-precipitate interactions reveal the importance of local interface curvature. Acta Materialia 92:33–45CrossRef Prakash A, Guénolé J, Wang J, Müller J, Spiecker E, Mills MJ, Povstugar I, Choi P, Raabe D, Bitzek E (2015) Atom probe informed simulations of dislocation-precipitate interactions reveal the importance of local interface curvature. Acta Materialia 92:33–45CrossRef
30.
Zurück zum Zitat Zepeda-Ruiz LA, Stukowski A, Oppelstrup T, Bulatov VV (2017) Probing the limits of metal plasticity with molecular dynamics simulations. Nature 550(7677):492–495PubMedCrossRef Zepeda-Ruiz LA, Stukowski A, Oppelstrup T, Bulatov VV (2017) Probing the limits of metal plasticity with molecular dynamics simulations. Nature 550(7677):492–495PubMedCrossRef
31.
Zurück zum Zitat Zepeda-Ruiz LA, Stukowski A, Oppelstrup T, Bertin N, Barton NR, Freitas R, Bulatov VV (2021) Atomistic insights into metal hardening. Nat Mater 20(3):315–320PubMedCrossRef Zepeda-Ruiz LA, Stukowski A, Oppelstrup T, Bertin N, Barton NR, Freitas R, Bulatov VV (2021) Atomistic insights into metal hardening. Nat Mater 20(3):315–320PubMedCrossRef
32.
Zurück zum Zitat Schiøtz J (2004) Atomic-scale modeling of plastic deformation of nanocrystalline copper. Scripta Materialia 51(8):837–841CrossRef Schiøtz J (2004) Atomic-scale modeling of plastic deformation of nanocrystalline copper. Scripta Materialia 51(8):837–841CrossRef
33.
Zurück zum Zitat Van Swygenhoven H, Derlet PM, Frøseth AG (2006) Nucleation and propagation of dislocations in nanocrystalline FCC metals. Acta Materialia 54(7):1975–1983CrossRef Van Swygenhoven H, Derlet PM, Frøseth AG (2006) Nucleation and propagation of dislocations in nanocrystalline FCC metals. Acta Materialia 54(7):1975–1983CrossRef
34.
Zurück zum Zitat Tucker GJ, McDowell DL (2011) Non-equilibrium grain boundary structure and inelastic deformation using atomistic simulations. Int J Plast 27(6):841–857CrossRef Tucker GJ, McDowell DL (2011) Non-equilibrium grain boundary structure and inelastic deformation using atomistic simulations. Int J Plast 27(6):841–857CrossRef
36.
Zurück zum Zitat Valiev RZ, Langdon TG (2006) Principles of equal-channel angular pressing as a processing tool for grain refinement. Prog Mater Sci 51(7):881–981CrossRef Valiev RZ, Langdon TG (2006) Principles of equal-channel angular pressing as a processing tool for grain refinement. Prog Mater Sci 51(7):881–981CrossRef
38.
Zurück zum Zitat Segal VM (1995) Materials processing by simple shear. Mater Sci Eng A 197(2):157–164CrossRef Segal VM (1995) Materials processing by simple shear. Mater Sci Eng A 197(2):157–164CrossRef
39.
Zurück zum Zitat Segal VM (2002) Severe plastic deformation: simple shear versus pure shear. Mater Sci Eng A 338(1–2):331–344CrossRef Segal VM (2002) Severe plastic deformation: simple shear versus pure shear. Mater Sci Eng A 338(1–2):331–344CrossRef
40.
Zurück zum Zitat Prakash A, Hummel M, Schmauder S, Bitzek E (2016) NanoSCULPT: a methodology for generating complex realistic configurations for atomistic simulations. MethodsX 3:1–9CrossRef Prakash A, Hummel M, Schmauder S, Bitzek E (2016) NanoSCULPT: a methodology for generating complex realistic configurations for atomistic simulations. MethodsX 3:1–9CrossRef
41.
Zurück zum Zitat Guénolé J, Nöhring WG, Vaid A, Houllé F, Xie Z, Prakash A, Bitzek E (2020) Assessment and optimization of the fast inertial relaxation engine (FIRE) for energy minimization in atomistic simulations and its implementation in LAMMPS. Comput Mater Sci 175:109–584CrossRef Guénolé J, Nöhring WG, Vaid A, Houllé F, Xie Z, Prakash A, Bitzek E (2020) Assessment and optimization of the fast inertial relaxation engine (FIRE) for energy minimization in atomistic simulations and its implementation in LAMMPS. Comput Mater Sci 175:109–584CrossRef
42.
Zurück zum Zitat Hoover WG (1985) Canonical dynamics: equilibrium phase-space distributions. Phys Rev A 31(3):1695–1697CrossRef Hoover WG (1985) Canonical dynamics: equilibrium phase-space distributions. Phys Rev A 31(3):1695–1697CrossRef
43.
Zurück zum Zitat Mishin Y, Farkas D, Mehl M, Papaconstantopoulos D (1999) Interatomic potentials for monoatomic metals from experimental data and ab initio calculations. Phys Rev B 59(5):3393–3407CrossRef Mishin Y, Farkas D, Mehl M, Papaconstantopoulos D (1999) Interatomic potentials for monoatomic metals from experimental data and ab initio calculations. Phys Rev B 59(5):3393–3407CrossRef
44.
Zurück zum Zitat Mishin Y, Mehl M, Papaconstantopoulos D, Voter A, Kress J (2001) Structural stability and lattice defects in copper: ab initio, tight-binding, and embedded-atom calculations. Phy Rev B 63(22):224106CrossRef Mishin Y, Mehl M, Papaconstantopoulos D, Voter A, Kress J (2001) Structural stability and lattice defects in copper: ab initio, tight-binding, and embedded-atom calculations. Phy Rev B 63(22):224106CrossRef
45.
Zurück zum Zitat Plimpton S (1995) Fast parallel algorithms for short-range molecular dynamics. J Comput Phys 117(1):1–19CrossRef Plimpton S (1995) Fast parallel algorithms for short-range molecular dynamics. J Comput Phys 117(1):1–19CrossRef
47.
Zurück zum Zitat Honeycutt JD, Andersen HC (1987) Molecular dynamics study of melting and freezing of small Lennard-Jones clusters. J Chem Phys 91:4950–4963CrossRef Honeycutt JD, Andersen HC (1987) Molecular dynamics study of melting and freezing of small Lennard-Jones clusters. J Chem Phys 91:4950–4963CrossRef
48.
Zurück zum Zitat Larsen P.M, Schmidt S, Schiøtz J (2016) Robust structural identification via polyhedral template matching. Model Simul Mater Sci Eng 24(5):055007CrossRef Larsen P.M, Schmidt S, Schiøtz J (2016) Robust structural identification via polyhedral template matching. Model Simul Mater Sci Eng 24(5):055007CrossRef
50.
Zurück zum Zitat Anderson PM, Hirth JP, Lothe J (2017) Theory of dislocations. Cambridge University Press, Cambridge Anderson PM, Hirth JP, Lothe J (2017) Theory of dislocations. Cambridge University Press, Cambridge
51.
Zurück zum Zitat Vimal M, Sandfeld S, Prakash A (2022) Grain segmentation in atomistic simulations using orientation-based iterative self-organizing data analysis. Materialia 21:101–314CrossRef Vimal M, Sandfeld S, Prakash A (2022) Grain segmentation in atomistic simulations using orientation-based iterative self-organizing data analysis. Materialia 21:101–314CrossRef
52.
Zurück zum Zitat Law R, Schmid S, Wheeler J (1990) Simple shear deformation and quartz crystallographic fabrics: a possible natural example from the Torridon area of NW Scotland. J Struct Geol 12(1):29–45CrossRef Law R, Schmid S, Wheeler J (1990) Simple shear deformation and quartz crystallographic fabrics: a possible natural example from the Torridon area of NW Scotland. J Struct Geol 12(1):29–45CrossRef
53.
Zurück zum Zitat Van Swygenhoven H, Derlet PM, Frøseth A (2004) Stacking fault energies and slip in nanocrystalline metals. Nat Mater 3:399–403PubMedCrossRef Van Swygenhoven H, Derlet PM, Frøseth A (2004) Stacking fault energies and slip in nanocrystalline metals. Nat Mater 3:399–403PubMedCrossRef
54.
Zurück zum Zitat Prakash A, Weygand D, Bitzek E (2017) Influence of grain boundary structure and topology on the plastic deformation of nanocrystalline aluminum as studied by atomistic simulations. Int J Plast 97:107–125CrossRef Prakash A, Weygand D, Bitzek E (2017) Influence of grain boundary structure and topology on the plastic deformation of nanocrystalline aluminum as studied by atomistic simulations. Int J Plast 97:107–125CrossRef
55.
Zurück zum Zitat Chassagne M, Legros M, Rodney D (2011) Atomic-scale simulation of screw dislocation/coherent twin boundary interaction in AL, AU, CU and NI. Acta Materialia 59(4):1456–1463CrossRef Chassagne M, Legros M, Rodney D (2011) Atomic-scale simulation of screw dislocation/coherent twin boundary interaction in AL, AU, CU and NI. Acta Materialia 59(4):1456–1463CrossRef
56.
Zurück zum Zitat Jin ZH, Gumbsch P, Ma E, Albe K, Lu K, Hahn H, Gleiter H (2006) The interaction mechanism of screw dislocations with coherent twin boundaries in different face-centred cubic metals. Scripta Materialia 54(6):1163–1168CrossRef Jin ZH, Gumbsch P, Ma E, Albe K, Lu K, Hahn H, Gleiter H (2006) The interaction mechanism of screw dislocations with coherent twin boundaries in different face-centred cubic metals. Scripta Materialia 54(6):1163–1168CrossRef
57.
Zurück zum Zitat Jin ZH, Gumbsch P, Albe K, Ma E, Lu K, Gleiter H, Hahn H (2008) Interactions between non-screw lattice dislocations and coherent twin boundaries in face-centered cubic metals. Acta Materialia 56(5):1126–1135CrossRef Jin ZH, Gumbsch P, Albe K, Ma E, Lu K, Gleiter H, Hahn H (2008) Interactions between non-screw lattice dislocations and coherent twin boundaries in face-centered cubic metals. Acta Materialia 56(5):1126–1135CrossRef
60.
Zurück zum Zitat Tucker GJ, Zimmerman JA, McDowell DL (2011) Continuum metrics for deformation and microrotation from atomistic simulations: application to grain boundaries. Int J Eng Sci 49(12):1424–1434CrossRef Tucker GJ, Zimmerman JA, McDowell DL (2011) Continuum metrics for deformation and microrotation from atomistic simulations: application to grain boundaries. Int J Eng Sci 49(12):1424–1434CrossRef
61.
Zurück zum Zitat Tucker GJ, Tiwari S, Zimmerman JA, McDowell DL (2012) Investigating the deformation of nanocrystalline copper with microscale kinematic metrics and molecular dynamics. J Mech Phys Solids 60(3):471–486CrossRef Tucker GJ, Tiwari S, Zimmerman JA, McDowell DL (2012) Investigating the deformation of nanocrystalline copper with microscale kinematic metrics and molecular dynamics. J Mech Phys Solids 60(3):471–486CrossRef
62.
Zurück zum Zitat Prakash A, Sandfeld S (2022) Automated analysis of continuum fields from atomistic simulations using statistical machine learning. Adv Eng Mater 24(12):2200–574CrossRef Prakash A, Sandfeld S (2022) Automated analysis of continuum fields from atomistic simulations using statistical machine learning. Adv Eng Mater 24(12):2200–574CrossRef
63.
Zurück zum Zitat Silcox J, Hirsch P (1959) Direct observations of defects in quenched gold. Philos Mag 4(37):72–89CrossRef Silcox J, Hirsch P (1959) Direct observations of defects in quenched gold. Philos Mag 4(37):72–89CrossRef
64.
Zurück zum Zitat Loretto M, Phillips P, Mills M (2015) Stacking fault tetrahedra in metals. Scripta Materialia 94:1–4CrossRef Loretto M, Phillips P, Mills M (2015) Stacking fault tetrahedra in metals. Scripta Materialia 94:1–4CrossRef
65.
Zurück zum Zitat Kiritani M, Satoh Y, Kizuka Y, Arakawa K, Ogasawara Y, Arai S, Shimomura Y (1999) Anomalous production of vacancy clusters and the possibility of plastic deformation of crystalline metals without dislocations. Philos Mag Lett 79(10):797–804CrossRef Kiritani M, Satoh Y, Kizuka Y, Arakawa K, Ogasawara Y, Arai S, Shimomura Y (1999) Anomalous production of vacancy clusters and the possibility of plastic deformation of crystalline metals without dislocations. Philos Mag Lett 79(10):797–804CrossRef
66.
Zurück zum Zitat Wang H, Xu D, Yang R, Veyssière P (2008) The transformation of edge dislocation dipoles in aluminium. Acta Materialia 56(17):4608–4620CrossRef Wang H, Xu D, Yang R, Veyssière P (2008) The transformation of edge dislocation dipoles in aluminium. Acta Materialia 56(17):4608–4620CrossRef
67.
Zurück zum Zitat Wang H, Xu D, Yang R, Veyssière P (2011) The formation of stacking fault tetrahedra in Al and Cu I: dipole annihilation and the nucleation stage. Acta Materialia 59(1):1–9CrossRef Wang H, Xu D, Yang R, Veyssière P (2011) The formation of stacking fault tetrahedra in Al and Cu I: dipole annihilation and the nucleation stage. Acta Materialia 59(1):1–9CrossRef
68.
Zurück zum Zitat Phillips P, Mills M (2013) Fine-scale structure of dislocations and debris in deformed NI-based superalloy r104. Philos Mag 93(1–3):82–95CrossRef Phillips P, Mills M (2013) Fine-scale structure of dislocations and debris in deformed NI-based superalloy r104. Philos Mag 93(1–3):82–95CrossRef
69.
Zurück zum Zitat Wei Wang J, Narayanan S, Yu Huang J, Zhang Z, Zhu T, Mao S.X (2013) Atomic-scale dynamic process of deformation-induced stacking fault tetrahedra in gold nanocrystals. Nat Commun 4(1):2340CrossRef Wei Wang J, Narayanan S, Yu Huang J, Zhang Z, Zhu T, Mao S.X (2013) Atomic-scale dynamic process of deformation-induced stacking fault tetrahedra in gold nanocrystals. Nat Commun 4(1):2340CrossRef
70.
Zurück zum Zitat Osetsky YN, Stoller RE, Rodney D, Bacon DJ (2005) Atomic-scale details of dislocation-stacking fault tetrahedra interaction. Mater Sci Eng A 400:370–373CrossRef Osetsky YN, Stoller RE, Rodney D, Bacon DJ (2005) Atomic-scale details of dislocation-stacking fault tetrahedra interaction. Mater Sci Eng A 400:370–373CrossRef
71.
Zurück zum Zitat Osetsky Y.N., Rodney D, Bacon D.J (2006) Atomic-scale study of dislocation–stacking fault tetrahedron interactions, Part I: mechanisms. Philos Mag 86(16):2295–2313CrossRef Osetsky Y.N., Rodney D, Bacon D.J (2006) Atomic-scale study of dislocation–stacking fault tetrahedron interactions, Part I: mechanisms. Philos Mag 86(16):2295–2313CrossRef
72.
Zurück zum Zitat Kraft O, Gruber PA, Mönig R, Weygand D (2010) Plasticity in confined dimensions. Ann Rev Mater Res 40:293–317CrossRef Kraft O, Gruber PA, Mönig R, Weygand D (2010) Plasticity in confined dimensions. Ann Rev Mater Res 40:293–317CrossRef
73.
Zurück zum Zitat Van Swygenhoven H (2008) Footprints of plastic deformation in nanocrystalline metals. Mater Sci Eng A 483:33–39CrossRef Van Swygenhoven H (2008) Footprints of plastic deformation in nanocrystalline metals. Mater Sci Eng A 483:33–39CrossRef
75.
Zurück zum Zitat Lee S, Vaid A, Im J, Kim B, Prakash A, Guénolé J, Kiener D, Bitzek E, Oh SH (2020) In-situ observation of the initiation of plasticity by nucleation of prismatic dislocation loops. Nat Commun 11(1):1–11 Lee S, Vaid A, Im J, Kim B, Prakash A, Guénolé J, Kiener D, Bitzek E, Oh SH (2020) In-situ observation of the initiation of plasticity by nucleation of prismatic dislocation loops. Nat Commun 11(1):1–11
76.
Zurück zum Zitat Prakash A, Sandfeld S (2018) Chances and challenges in fusing data science with materials science. Pract Metall 55(8):493–514CrossRef Prakash A, Sandfeld S (2018) Chances and challenges in fusing data science with materials science. Pract Metall 55(8):493–514CrossRef
77.
Zurück zum Zitat Prakash A, Bitzek E (2017) Idealized vs. realistic microstructures: an atomistic simulation case study on \(\gamma /\gamma ^{\prime }\) microstructures. Materials 10:88PubMedPubMedCentralCrossRef Prakash A, Bitzek E (2017) Idealized vs. realistic microstructures: an atomistic simulation case study on \(\gamma /\gamma ^{\prime }\) microstructures. Materials 10:88PubMedPubMedCentralCrossRef
79.
Zurück zum Zitat Rosochowski A, Olejnik L (2007) Finite element simulation of severe plastic deformation processes. Proc Inst Mech Eng Part L: J Mater Des Appl 221(4):187–196 Rosochowski A, Olejnik L (2007) Finite element simulation of severe plastic deformation processes. Proc Inst Mech Eng Part L: J Mater Des Appl 221(4):187–196
80.
Zurück zum Zitat Prakash A, Nöhring W, Lebensohn R, Höppel H, Bitzek E (2015) A multiscale simulation framework of the accumulative roll bonding process accounting for texture evolution. Mater Sci Eng A 631:104–119CrossRef Prakash A, Nöhring W, Lebensohn R, Höppel H, Bitzek E (2015) A multiscale simulation framework of the accumulative roll bonding process accounting for texture evolution. Mater Sci Eng A 631:104–119CrossRef
81.
Zurück zum Zitat Beyerlein I, Lebensohn R, Tome C (2003) Modeling texture and microstructural evolution in the equal channel angular extrusion process. Mater Sci Eng A 345(1–2):122–138CrossRef Beyerlein I, Lebensohn R, Tome C (2003) Modeling texture and microstructural evolution in the equal channel angular extrusion process. Mater Sci Eng A 345(1–2):122–138CrossRef
82.
Zurück zum Zitat Kalidindi SR, Donohue BR, Li S (2009) Modeling texture evolution in equal channel angular extrusion using crystal plasticity finite element models. Int J Plast 25(5):768–779CrossRef Kalidindi SR, Donohue BR, Li S (2009) Modeling texture evolution in equal channel angular extrusion using crystal plasticity finite element models. Int J Plast 25(5):768–779CrossRef
83.
Zurück zum Zitat Lu C, Deng G, Tieu A, Su L, Zhu H, Liu X (2011) Crystal plasticity modeling of texture evolution and heterogeneity in equal channel angular pressing of aluminum single crystal. Acta Materialia 59(9):3581–3592CrossRef Lu C, Deng G, Tieu A, Su L, Zhu H, Liu X (2011) Crystal plasticity modeling of texture evolution and heterogeneity in equal channel angular pressing of aluminum single crystal. Acta Materialia 59(9):3581–3592CrossRef
84.
Zurück zum Zitat Tóth LS, Estrin Y, Lapovok R, Gu C (2010) A model of grain fragmentation based on lattice curvature. Acta Materialia 58(5):1782–1794CrossRef Tóth LS, Estrin Y, Lapovok R, Gu C (2010) A model of grain fragmentation based on lattice curvature. Acta Materialia 58(5):1782–1794CrossRef
86.
Zurück zum Zitat Kobaissy AAH, Ayoub G, Nasim W, Malik J, Karaman I, Shehadeh M (2020) Modeling of the ECAP induced strain hardening behavior in FCC metals. Metall Mater Trans A 51(10):5453–5474CrossRef Kobaissy AAH, Ayoub G, Nasim W, Malik J, Karaman I, Shehadeh M (2020) Modeling of the ECAP induced strain hardening behavior in FCC metals. Metall Mater Trans A 51(10):5453–5474CrossRef
87.
Zurück zum Zitat . Möller J.J, Prakash A, Bitzek E (2013) FE2AT—finite element informed atomistic simulations. Modell Simul Mater Sci Eng 21(5):05501CrossRef . Möller J.J, Prakash A, Bitzek E (2013) FE2AT—finite element informed atomistic simulations. Modell Simul Mater Sci Eng 21(5):05501CrossRef
Metadaten
Titel
Atomistic modeling of idealized equal channel angular pressing process
verfasst von
Aruna Prakash
Publikationsdatum
27.02.2024
Verlag
Springer US
Erschienen in
Journal of Materials Science / Ausgabe 14/2024
Print ISSN: 0022-2461
Elektronische ISSN: 1573-4803
DOI
https://doi.org/10.1007/s10853-024-09414-w

Weitere Artikel der Ausgabe 14/2024

Journal of Materials Science 14/2024 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.