This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

Towards the reliable calculation of residence time for off-lattice kinetic Monte Carlo simulations

and

Published 5 August 2016 © 2016 IOP Publishing Ltd
, , Citation Kathleen C Alexander and Christopher A Schuh 2016 Modelling Simul. Mater. Sci. Eng. 24 065014 DOI 10.1088/0965-0393/24/6/065014

0965-0393/24/6/065014

Abstract

Kinetic Monte Carlo (KMC) methods have the potential to extend the accessible timescales of off-lattice atomistic simulations beyond the limits of molecular dynamics by making use of transition state theory and parallelization. However, it is a challenge to identify a complete catalog of events accessible to an off-lattice system in order to accurately calculate the residence time for KMC. Here we describe possible approaches to some of the key steps needed to address this problem. These include methods to compare and distinguish individual kinetic events, to deterministically search an energy landscape, and to define local atomic environments. When applied to the ground state  ∑5(2 1 0) grain boundary in copper, these methods achieve a converged residence time, accounting for the full set of kinetically relevant events for this off-lattice system, with calculable uncertainty.

Export citation and abstract BibTeX RIS

Introduction

Stochastic computational methods like kinetic Monte Carlo (KMC) provide a powerful alternative to molecular dynamics (MD) simulations since they offer a pathway to increasing simulation times by orders of magnitude. Additionally, the stochastic nature of KMC lends itself to efficient and straightforward parallelization, which allows for the optimized use of advanced computing resources. The reason for the timescale advantage of KMC methods derives from their use of a variable system residence time, rather than a timestep of pre-specified length [1]. Residence time, ${{\tau}_{\text{R}}}$ , can be calculated according to,

Equation (1)

where N is the total number of possible kinetic events accessible to a system, ${{\sigma}_{i}}$ is the degeneracy of event i, $\nu _{i}^{\ast}$ is the effective vibrational frequency associated with event i, ${{E}_{i}}$ is the activation energy required for event i, and kT is the thermal energy.

The challenges in implementing KMC in off-lattice systems include the need for methods to (a) determine the kinetic processes accessible to a configuration, i.e. find the ${{E}_{i}}$ terms in equation (1), and (b) ensure that the catalog of N events is sufficiently complete such that the sum in equation (1) does not miss critical contributing terms. Only when these criteria have been satisfied can the system kinetics calculated with a KMC method be considered meaningful.

A number of methods for finding the kinetic processes of off-lattice systems have been developed and implemented over the past several decades. In particular, the nudged-elastic band (NEB) method [2, 3] and the activation–relaxation technique (ART) [4, 5] have been shown to reliably map kinetic pathways on an energy landscape when adjacent minimum configurations are either known or unknown, respectively. However, there is no clear way to ensure the completeness of an energy landscape mapped using these methods, and we are not aware of quantitative efforts to estimate the degree of completeness of an energy landscape mapping. As long as the completeness of the energy landscape cannot be verified, the results of KMC approaches that use these methods for anything beyond the simplest systems could be met with skepticism.

For off-lattice KMC applications, methods not requiring a priori knowledge of adjacent minimum configurations are required to map an energy landscape. The ART algorithm is an example of such a method; it initiates a search for a kinetic event with a local, random perturbation (e.g. the displacement of a single random atom in a random direction) [4]. In this framework, mapping the energy landscape of a target configuration requires a series of these randomly initiated searches. However, in previous work there has been minimal focus on characterizing the total number of searches required before an energy landscape has been sufficiently mapped, and we suggest that random probing may not be the most efficient search method. We expect this to be particularly true for partially ordered systems like crystals containing defects. Conversely, a systematic and orderly search over the perturbation space could efficiently seek transitions and opens the door to a gradual refinement of the search in a way that decreases the uncertainty about the completeness of the search.

It is the purpose of this paper to examine such a deterministic search process for kinetic events in an off-lattice atomistic configuration. For an example system—a metal bicrystal—we use the ART algorithm to identify transition states on an energy landscape. We suggest methods for the unique specification of kinetic events, and a modification to the traditional ART approach that aims to deterministically probe perturbation space. With these tools, we are able to quantitatively estimate the relative completeness of the search. We then discuss our method of uniquely specifying local atomic environments (LAEs) and provide an example using this new technique to map the energy landscape of a grain boundary in copper.

An example system

In implementing the methods described in this document, we have studied a symmetric tilt grain boundary (GB) in copper, namely the  ∑5(2 1 0) grain boundary, as an example system. We consider grain boundaries as ideal target systems for off-lattice KMC, as they generally present a level of structural complexity that is beyond what can be analytically considered or treated with lattice KMC approaches [6], yet it is not so disordered, as in the case of a metallic glass [7], that the kinetic events of the system cannot be sufficiently tabulated.

The example GB was minimized with the LAMMPS [8, 9] molecular dynamics package using the static minimization technique described elsewhere [10] as a planar bicrystal with periodic boundary conditions in the plane of the GB and free surfaces in the direction perpendicular to the plane of the GB. This system is shown in figure 1, which was generated with the Ovito open visualization tool [11, 12] and shows atoms colored according to their centrosymmetry parameter [13]. The system contains 2040 atoms and is about 290 ${{{\mathring{\rm A}}}^{2}}$ in the plane of the GB. The embedded atom method (EAM) potential developed by Mishin et al [14, 15] was used both for the minimization of the grain boundary structure in LAMMPS and with the ART method in what follows.

Figure 1.

Figure 1. The  ∑5(2 1 0) grain boundary in copper which was used as an example system in this study; the image was generated with the Ovito open visualization tool [11, 12]. The exploded view shows only the atoms in the grain boundary plane, with all other atoms made invisible. Atoms are colored according to their centrosymmetry parameter [13] with the color scale shown at the right.

Standard image High-resolution image

Mapping the energy landscape

The activation–relaxation technique (ART) is a method for finding kinetic pathways on the potential energy surface associated with a given configuration and does not require that the structures of any adjacent configurations be known a priori [4, 16]. It is described in detail in a number of sources [5, 1618], so we will only briefly summarize it here. ART is an eigenvector following method that works by pushing a system that has been perturbed from its ground state configuration along the direction of its lowest curvature. It is initiated by a perturbation step in which the starting configuration is disturbed from the ground state via the displacement of a single atom. Next, in the convergence phase of the algorithm, the atoms are iteratively translated along the direction corresponding to the lowest eigenvalue of the configuration's Hessian matrix, and relaxed in the hyperplane of this direction. If a saddle point is successfully found during this process, the system can then be relaxed from this saddle point to an adjacent minimum configuration. A saddle point is identified when there is only one negative eigenvalue in the system's Hessian matrix and the forces on all atoms in the system are nominally zero, i.e. less than some specified cutoff value ${{f}_{\text{tol}}}$ . The difference between the system's energy at the saddle point and at the ground state configuration is the activation energy of the kinetic process, corresponding to the ${{E}_{i}}$ term in equation (1).

In most published work using ART, e.g. [18] this search process is repeated by attempting different random perturbations in the initiation phase until, given enough attempts, it is presumed that the kinetic pathways accessible to the initial configuration have been sufficiently explored. The development of a method for systematically identifying all events that are kinetically relevant to a system's evolution is the subject of the remainder of this manuscript.

Identifying distinct kinetic events

The first step towards the goal of systematically mapping a configuration's complete energy landscape is to define a means by which two independently identified kinetic pathways can be compared and identified as either equivalent or unique. Without such a comparison method, it is impossible to assess the completeness of an energy landscape and, in turn, ensure that we have identified the N possible kinetic events required to reliably calculate residence time in equation (1).

The difficulty associated with uniquely specifying a kinetic event is twofold: (1) the numerical nature of computer simulations complicates the comparison of atomistic structures, and (2) strain in a local configuration due to structural deformations far from the local area of interest may distort a configuration without affecting its kinetic behavior. In this paper, we focus on the former of these issues, numerical uncertainty, noting that determining the level of strain a local configuration can undergo whilst maintaining the same kinetic behavior is an area of future research that has the potential to greatly improve the efficiency of the proposed methods. An efficient saddle point search method, such as ART, must make use of numerical parameters to determine whether or not a saddle point has been identified in a given search and whether or not a minimum has been reached during the relaxation process. In particular, since transition states exist at points on a potential energy surface with a zero force vector, one must employ a cutoff force tolerance, ${{f}_{\text{tol}}}$ , below which the force is considered to be nominally zero. For our simulations that follow, we use a default value of ftot  =  0.005 eV Å−1 (unless otherwise noted).

If we perturb our ground state system in two slightly different ways that both result in convergence upon the same physical saddle point, the atomic configurations identified by the algorithm at the saddle point will not generally have numerically identical atomic positions when they reach the force cutoff. The stricter we make the force tolerance parameter, the closer the saddle point configurations will be to one another. However, there is a steep computational penalty associated with a tightening of the force tolerance. We have found that saddle point searches with a decreased force cutoff of ftot  =  0.0009 eV Å−1 can take 50% longer to converge than the same searches with a cutoff of ftot  =  0.005 eV Å−1. In most cases, the transition state configuration found with the larger force tolerance differed from that found with the tighter force tolerance by less than 0.1 $\,{\mathring{\rm A}}$ in all atomic positions. In a few rare cases, we observed the positional difference between configurations that were converged to each of these tolerances to be as large as 0.25 ${\mathring{\rm A}}$ for the most deviant atomic position. Similarly, the uncertainty in atomic position at a minimum energy configuration is generally less than 0.01 ${\mathring{\rm A}}$ , and occasionally as high as 0.025 ${\mathring{\rm A}}$ (noting that, in this case, it is a much stricter set of cutoff parameters that terminate relaxation to a minimum). Thus, though it is computationally worthwhile to use the largest force tolerance that a method can accept (based on the required certainty for the activation energy), choosing to do so requires that the method used to compare transition states must be able to account for positional uncertainty of atoms that is generally on the order of 0.2 ${\mathring{\rm A}}$ at the saddle point and 0.02 ${\mathring{\rm A}}$ at a minimum configuration. In cases where even larger force tolerances have been used to converge to saddle points and adjacent minima, e.g. [19] the method of comparing saddle point configurations must be able to account for even greater positional uncertainty. Bearing in mind this challenge, we next review several approaches that have been employed in prior studies characterizing off-lattice kinetic events.

Classifying and comparing kinetic events

The simplest approaches to classifying kinetic events make use of individual macroscopic metrics such as the activation energy of an event and the relative energy of its adjacent minimum configuration [10], for example. However, these macroscopic values do not uniquely specify a given process, so researchers have turned to the use of more direct topological approaches.

Béland et al used a graphical isomorphism approach, specifically making using of a program called NAUTY [20, 21], to specify unique kinetic processes [22]. In this approach, a center-point for each event must first be identified. A sphere with radius, R, is then drawn around this center and the set of atoms that lie within this sphere then make up the nodes of a truncated graph. The edges of the graph are defined to connect any two nodes (atoms) that are closer than some edge cutoff distance, a, in the physical configuration. An example of how a graph is constructed in this fashion is shown in figure 2. A kinetic event, using this approach, is then completely specified by three graphs, one for each of the initial configuration, the transition state configuration, and the adjacent minimum configuration. NAUTY, or a similar tool, can be used to analyze these graphs and assign a set of hex keys corresponding to the canonical representation of each one. From this, the set of three hex keys that identify the graphs at the initial, transition, and final states uniquely specify a kinetic process, and anytime an event is identified, its uniqueness can be determined by comparing the hex keys associated with its configurations to those that have been found previously.

Figure 2.

Figure 2. 2D schematic of how a truncated graph is generated from a configuration of atoms given parameters R and a. A hex key identifying the canonical labeling of the graph is generated by NAUTY. In this approach, minor perturbations in atomic position, such as the slight displacement of the atom outlined in black in the exploded view, results in the presence of a new edge, colored light blue. The hex keys generated for the graph with this additional edge share no resemblance to those corresponding to the unperturbed configuration.

Standard image High-resolution image

This approach is powerful as it minimizes the amount of information that must be stored to fully specify a kinetic event and has a very straightforward comparison rule. However, it has a number of issues that make it ultimately unfavorable as a tool for uniquely specifying kinetic events. For example, the method is unable to account for the positional irregularities associated with the numerically converged configurations used in off-lattice methods. This results in either the false classification of two events as being distinct when they are, in fact, the same, or vice versa. In figure 2 an example is shown demonstrating how a small, numerically-based irregularity in an atom's position results in a graph with one additional edge and completely different graph hex keys. Additionally, the use of hex keys for transition state identification, though offering a simple means of defining kinetic processes, leaves the user no information with which to compare the similarity between two events that were found to be distinct, making it difficult to validate and optimize the method.

An alternate topological approach developed by Nandipati et al [23] that suffers from similar issues stores the transition state configuration as a binary number by overlaying a fine grid in the region central to the transition and assigning ones to the voxels of the grid that contain the center of mass of an atom and zeros to all other voxels. The grid is then unwrapped into a string of bits that make up a large binary number. This number can be conveniently stored and referenced. However, uncertainty of atomic position can, once again, confound this method, making it much too easy to store the same kinetic event multiple times due to minor perturbations in the atomic structure, complicating any attempt to determine the completeness of the energy landscape.

Present approach

Due to concerns with the existing methods for uniquely specifying a kinetic process, we have pursued an alternate approach. Since our goal is to uniquely label atomic topologies without abstraction that would lose the physical features of the structure, we have chosen to work directly with atomic coordinates and trajectories. Furthermore, knowing the importance of uncertainty in atomic positions associated with numerical methods in general and with ART in particular, we hope to leverage that knowledge in event labeling. Specifically, we store atomic motion vectors associated with the transition. For each kinetic event found with ART, the x, y, and z displacements of each atom are stored between both the initial state and the saddle point, and the initial and final states. A schematic illustrating an event in which 4 atoms are displaced is shown in figure 3. In our method, the full vector description of each ${{t}_{i}}$ displacement is stored to uniquely specify the kinetic event.

Figure 3.

Figure 3. Schematic of a kinetic event in an FCC system. In this example, 4 atoms, originally located at the dotted circles, participate in the process. The kinetic event is defined according to the x, y, and z components of the displacement vectors of these atoms (${{t}_{1}},~{{t}_{2}},~{{t}_{3}},\,\text{and}\,{{t}_{4}}$ ).

Standard image High-resolution image

The comparison of two kinetic processes can be conducted by comparing the x, y, and z displacements of each atom in each of the events. If the difference between the displacements of any atoms in any direction is greater than a cutoff value $\epsilon $ , which is chosen based on the known uncertainty associated with atomic position for the force tolerance being employed, then the events are taken to be unique. With a properly chosen value of $\epsilon $ , atomic motions that are literally indistinguishable to within numerical uncertainty associated with the potentials and ART can be clearly identified; provided the uncertainty in atomic positions is deemed acceptable in the first place, the uncertainty in comparing and labelling events is guaranteed to be no worse.

At first glance, this approach might seem unnecessarily cumbersome compared with some of the memory-leaner alternatives previously discussed, but it has several crucial advantages. To begin with, it uses the most basic definition of a kinetic event, i.e. the movements of atoms in the system, to define and distinguish kinetic events. Secondly, because it makes use of the known numerical uncertainty of atomic positions in distinguishing between events, it allows the user to choose convergence criteria that are logically consistent with an application without sacrificing the ability to compare events. A comparison of the present method with previous approaches is presented later in the Implementation section of this paper.

If the information about each kinetic process is stored in a database, we can easily minimize the search space by performing comparisons with only the subset of events that have, e.g. activation energies within some tolerance of that found in the current search, or some other rapid screening criterion. This has the added advantage of allowing us to consider only the displacements between the initial and final configurations of a process, which halves the number of required comparisons by excluding the saddle point, which is associated with greater uncertainty in atomic position. Though it is possible for multiple activated states to lead to the same final state, we have not found these activated states to have similar enough activation energies so as to be confused with one another, at least in the system of our analysis. It is possible that some systems may be able to access symmetrically related transition states, in which case, both the saddle point and final configurations would need to be compared to identify a unique event.

Completeness of the search

Assuming that we can reliably distinguish between two kinetic events, it is possible to determine whether an event has been previously identified or is new as the search of the energy landscape progresses. From this, it is possible to develop metrics for determining the completeness of an energy landscape search.

Search termination criteria

The simplest approach to terminating an ART search is to qualitatively deem the search to be complete after some arbitrary number of search attempts or successful searches [17, 18]. For example one could deem their search complete when n unique saddle points have been identified [17, 24] but n remains somewhat arbitrary. More sophisticated stopping criteria have been introduced; for example Fan et al in [24] terminate a search if a newly found event has a kinetic probability that is less than 0.1% of the current residence time estimate. However, this approach still relies on arbitrary factors such as the order in which events are randomly identified. An improvement over arbitrary termination is to choose some property of the search or the energy landscape around which a statistical test can be built to determine the probability that the chosen property is sufficiently close to the true value. For example, in [10] we used the frequency of the most commonly found kinetic event as a metric for terminating the search. Specifically, we binned all the kinetic events found with random ART searches and tracked how the average frequency across the bins of the most commonly found saddle point, ${{f}_{\max}}$ , changed with additional successful searches. We then used the sample mean of ${{f}_{\max}}$ and its standard deviation, s, in a Student's t-test to calculate the minimum sample size, ${{n}_{\min}}$ , that would be required to achieve the desired confidence that our calculated value was sufficiently close to the true value, and once our sample size exceeded ${{n}_{\min}}$ , we deemed our search complete.

The above proposal was presented in the context of a static energy landscape search, i.e. without connection to a KMC context or, indeed, any structural evolution. Since system kinetics are of prime interest to a KMC application, we presently suggest that a further improved approach would be to define a metric based on a kinetic property of the system, rather than a numerical one, as has been previously suggested [24]. Specifically, since an accurate calculation of the residence time, ${{\tau}_{\text{R}}}$ , is our goal, we propose to use the convergence of ${{\tau}_{\text{R}}}$ to assess the completeness of an energy landscape search. The metric used to establish convergence in our present approach is discussed below.

A deterministic search method

Whether one uses an arbitrary or quantitative approach to search termination, the random nature of the perturbations typically used in the search method are a common limitation. In either case, we cannot be sure that we have effectively searched the energy landscape, and, in either case, we must perform a large number of searches that, by definition, contain no new information in order to sufficiently characterize the energy landscape; this all results in an inefficient use of computational resources.

As an alternative to the stochastic search method, we employ a deterministic search method by applying a set of systematic perturbations (each of which separately initiates the ART algorithm) to one atom in the system at a time. The perturbations make up the points of a 3D mesh, with points that are uniformly spaced on concentric spheres with radii corresponding to the magnitude of the perturbation, ${{d}_{i}}$ , and centered on the atom being investigated. The product of the number of directions on a perturbation sphere, $\rho $ , and the number of shells in the radial direction of the mesh, $d$ , is the total number of perturbations in the search of each atom. In figure 4, perturbation meshes with several values for $\rho $ and a representative set of concentric shells (with $d=4$ ) that are separated in the radial direction by a distance δ, are shown to illustrate the nature of the perturbation space used in the deterministic search.

Figure 4.

Figure 4. Perturbation meshes with increasing $\rho $ can be used to calculate a system's residence time and determine the minimum required value of $\rho.$ Each perturbation mesh is projected on a set of $d$ concentric shells, yielding the full set of perturbation vectors required for the systematic search. The $\rho $ points are placed approximately uniformly on the sphere according to the method described by Rusin [25].

Standard image High-resolution image

When $\rho $ and d are made sufficiently large and $\delta ~$ is made sufficiently small, the systematic search can explore all the relevant kinetic pathways in a system. However, determining the critical values for $\rho $ , d, and $\delta $ requires some optimization. By calculating the residence time for a system as a function of these parameters, a critical value—beyond which residence time is independent of $\rho $ , d, and $\delta $ respectively—can be determined for each of them. One caveat important to the validity of this approach is that meshes of increasing $\rho $ or d must not inherit all the same perturbation directions as the coarser meshes. This requirement removes the bias on $\rho $ and d that is introduced if particular directions or radii in the coarse mesh are very advantageously aligned for searching certain atomic environments.

It should be noted that future approaches may achieve efficiency improvements in this search process by exploiting the symmetry in a LAE to minimize the set of perturbations required to search its energy landscape. Additionally, as with other methods that rely on the perturbation of individual atoms to initiate the search for kinetic events, there is the necessary concern that concerted events involving large numbers of atoms will not be identified by the search method. In our prior work, we have found that eigenvector-following methods such as ART allow for collective mechanisms to be identified due to the inherently non-local nature of the eigenmodes of a system [10]. This does not guarantee that all collective mechanisms will be found during a local perturbation search process, as described in this paper. As such, future studies may pursue a more exhaustive investigation into this topic.

Describing atomic environments uniquely

The efficiency of our deterministic search method hinges on the need to only search for the kinetic events of atoms in unique local atomic environments (LAEs); atomic environments that are sufficiently similar will produce the same set of kinetic events, and thus need not be searched again. If we have a means to specify a LAE, then we only need to perform this systematic search on one representative atom with each unique LAE. Particularly for the case of periodic defects such as grain boundaries or defects that appear with many high-fidelity duplicates in the same system, like point defects, the consideration of only unique LAEs leverages the order in the structure to simplify the computational problem.

Throughout the literature, a number of parameters have been used to specify LAEs in off-lattice systems including the centrosymmetry parameter [13], the common neighbor analysis [26], and hex keys from graphical or topological approaches such as those presented in our discussion above [19, 23]. However, each of these approaches is limited in either its ability to uniquely specify LAEs (as in the first two cases), or in its ability to accommodate small perturbations in atomic structure (as in the latter cases), as described earlier. We therefore suggest a more exhaustive approach to labelling LAEs. Namely, we define an LAE by the set of vectors, ${{r}_{i}}$ , denoting the M closest neighbor positions relative to our target atom. An appropriate size for M needs yet to be determined, and will depend on the complexity of the system being studied. However, we envision that at a minimum one would require the nearest-neighbor positions, which is illustrated as an example only for an FCC system in figure 5, i.e. $M=12$ . Two atomic environments are deemed to be equivalent if the vectors to all the M closest neighbors are the same to within $\epsilon $ , where $\epsilon $ is again the uncertainty in atomic position, 0.025 ${\mathring{\rm A}}$ at minima on the potential energy landscape in our method. In an efficient application of this approach, a means to quickly identify if LAEs are rotationally or symmetrically equivalent should be implemented.

Figure 5.

Figure 5. The vectors, ${{r}_{i}}$ , locating the M closest neighbors of a central atom, are used to define its unique atomic environment, illustrated here for the case of $M=12$ .

Standard image High-resolution image

In addition to storing the M closest neighbor vectors, the magnitude of the square of the vector sum of these vectors, ${{S}^{2}},$ can also be stored to facilitate screening the set of known LAEs (where ${{r}_{ij}}$ indicates the jth component of vector ${{r}_{i}}$ ):

Equation (2)

This permits the fewest possible comparisons to determine the uniqueness of a target LAE. In this scenario, only candidates with ${{S}^{2}}$ values that are within ${{\epsilon}_{{{S}^{2}}}}$ of the target LAE's ${{S}^{2}}$ are compared, where ${{\epsilon}_{{{S}^{2}}}}$ is based on the propagation of uncertainty of $\epsilon ~$ [27]:

Equation (3)

Given this approach, the challenge of specifying an atom's LAE becomes the challenge of determining what value of M is sufficient to uniquely specify a LAE for kinetic simulations, which is discussed in the Implementation section below. The kinetic transitions that result from the perturbation of an atom with a given LAE can then be systematically searched according to the method described above. This also improves the performance of the method we have described for comparing kinetic events, because once an LAE has been searched, we no longer need to store all the information about its kinetic events. Instead, we can simply store the information about a perturbation that led to the event as well as the energetic information required for the residence time calculation. When implemented in a KMC algorithm, a kinetic event is selected from the catalog of possible events, and the system can be evolved by executing the perturbation and ART search that led to finding the event initially. This prevents the method from unintentionally forcing an event that is not physical and can accommodate elastic strains present in dynamic processes, as has been discussed in other KMC cataloging approaches [18].

Implementation

We have implemented these methods in a study of the previously introduced example system, the  ∑5(2 1 0) grain boundary in copper, and here we evaluate how well these methods allow us to reliably calculate residence time and achieve quantifiable uncertainty. This process requires the optimization of several parameters including the best value of $\epsilon $ (the uncertainty in atomic position) to use to distinguish between kinetic events, the critical values of $\rho $ , $d$ , and $\delta $ (the number of perturbation directions on a perturbation sphere, the number of concentric perturbation spheres, and the radial distance between perturbation spheres, respectively) required to specify the deterministic search, and the number of atoms, M, necessary to sufficiently define a LAE.

In the following analysis, we have calculated residence time at 773 K and approximated the vibrational frequency at each saddle point to be constant at $8\centerdot {{10}^{12}}$ Hz, which we have found to be a good approximation for these kinetic processes after conducting a more exhaustive study of its range of values.

Sensitivity to the choice of $\epsilon $

We consider values for $\epsilon \,$ in the range $0.010-0.050\,{\mathring{\rm A}}$ based on our knowledge that atomic position has an uncertainty on the order of $0.01\,{\mathring{\rm A}}$ , but can be as much as $0.025\,{\mathring{\rm A}}$ in minimum energy configurations. To assess how the residence time is affected by $\epsilon $ , we have calculated the residence time, according to equation (1), using our systematic search method for various choices of $\epsilon $ in this target range. Since the critical value of $\rho $ is initially unknown, we perform the systematic search for a range of $\rho $ values (with d  =  4, $\delta ~$   =  0.5 Å, and ${{d}_{1}}=0.75\,{\mathring{\rm A}}$ held constant), the results of which are shown in figure 6.

Figure 6.

Figure 6. Residence time of the ground state $\Sigma\text{5}(2\,1\,0)$ GB in copper versus perturbation directions used in a deterministic search when different values of $\epsilon $ are used to distinguish between kinetic events. Error bars are centered on data with $\epsilon =0.025~\,{\mathring{\rm A}}$ and are defined by the maximum difference between the residence time calculated with $\epsilon =0.010\,{\mathring{\rm A}}$ and $\epsilon =0.050~\,{\mathring{\rm A}}$ with $\rho d~>~256$ .

Standard image High-resolution image

From figure 6, we note that the calculated residence time seems to be relatively converged for about 500 perturbation directions; a more quantitative assessment of this is discussed below. As expected, larger values of $\epsilon $ result in larger residence times due to cases where distinct events are incorrectly identified as being the same, while the smallest values of $\epsilon $ result in the opposite effect, where the same kinetic event is incorrectly counted multiple times. The optimal value of $\epsilon $ is different for each event. Some kinetic events are somewhat sloppy, with the highest uncertainty in their atomic positions ($\pm 0.025\,{\mathring{\rm A}}$ ), for which a large $\epsilon $ is preferable, while many other events are much cleaner and a smaller $\epsilon $ could be used. However, for meshes with greater than 250 perturbation directions, the difference between the residence time calculated with the different choices of $\epsilon $ is, at most $7.9\centerdot {{10}^{-11}}\,~\text{s}$ . Thus, we take half this value, $4.0\centerdot {{10}^{-11}}\,\text{s}$ , to be the uncertainty associated with the residence time we obtain with this method (shown as error bars in figure 6) and choose an intermediate $\epsilon =0.025\,{\mathring{\rm A}}$ in future calculations.

At this point, it is important to recall that none of the exact points on the sparser meshes are inherently included in the denser meshes. Thus, it is possible to find certain kinetic events with a sparser mesh that are then missed with the denser mesh. This prevents the determination of the minimal mesh size from being biased by a 'lucky' sparse mesh population that is inherited by all denser meshes and accounts for the non-monotonic behavior seen in figure 6.

Complete search

Using $\epsilon =0.025\,{\mathring{\rm A}}$ , we can now optimize the mesh parameters required for a complete search of the energy landscape for our target system. From figure 6 it appears that the critical value of ρ may be around 150; however, it is possible that with different choices of d and δ a more minimal perturbation sphere may be used. We proceed by optimizing the radial parameters of the mesh. After finding that no search converged to a saddle point when the initial perturbation was less than 0.75 Å in our implementation of ART, we define the space of possible radial perturbation distances as ranging from 0.75–2.25 Å. We consider shell spacing values of $\delta $ ranging between 0.1–1.0 ${\mathring{\rm A}}$ . Setting $\rho =328$ , we investigate residence time as a function of d and $\delta $ , as shown in figure 7.

Figure 7.

Figure 7. Calculated residence time for the ground state configuration of the  ∑5(2 1 0) GB in copper as a function of radial shells, d, and shell spacing, $\delta $ , for a perturbation mesh with $\rho $   =  328. If $\delta $ is set to $~0.15\,{\mathring{\rm A}}$ , then d  =  4 is the minimal number of required shells in this system.

Standard image High-resolution image

From figure 7, we note that the residence time is independent of d so long as $d~>~4$ and $\delta =0.15~\,{\mathring{\rm A}}$ . As such, we have found that including perturbations $>1.20~\,{\mathring{\rm A}}$ did not affect the residence time calculation. Decreasing $\delta $ to 0.10 Å did not affect the residence time either. However, we note that the residence time did show sensitivity to the choice of the lowest ${{d}_{i}}$ value, ${{d}_{1}}$ . Holding d  =  5 and $\delta =0.15\,{\mathring{\rm A}}$ , we increased ${{d}_{1}}$ from 0.75 ${\mathring{\rm A}}$ to 0.85 ${\mathring{\rm A}}$ and observed an increase in residence time.

Having a sense of the optimal set of perturbation shells with ρ held constant, we next need to determine the codependence of our mesh parameters. In figure 8, we show the residence time as a function of d for $\rho ~=~156$ and $\rho ~=~328$ with $\delta ~=~0.15$ ${\mathring{\rm A}}$ and ${{d}_{1}}=0.75\,{\mathring{\rm A}}$ and error bars centered on data with $\rho ~=~328$ . In figure 8, it can be seen that the residence time versus d behavior is independent of the mesh density ρ, with both datasets converging to d  =  4 as the minimal number of radial shells in the perturbation mesh.

Figure 8.

Figure 8. Calculated residence time for the ground state configuration of the  ∑5(2 1 0) GB in copper as a function of radial shells, d, in the perturbation mesh with $\rho ~=~156$ and $\rho ~=~328$ and $\delta =~0.15~\,{\mathring{\rm A}}$ .

Standard image High-resolution image

Having established that the mesh parameters are independent, we can now revisit figure 6 and determine the minimal mesh density, $\rho $ , for this system. Noting that choosing a constant value for $\epsilon $ lead to an uncertainty in the residence time of $4.0\centerdot {{10}^{-11}}\,\,\text{s}$ , we can calculate the standard deviation of the residence time for the data shown in figure 6 with $\epsilon =0.025\,{\mathring{\rm A}}$ . By successively neglecting the data from the coarsest meshes, we obtain a series of standard deviation values, ${{s}_{\overline{{{\tau}_{\text{R}}}},i}}$ , where ${{\overline{{{\tau}_{\text{R}}}}}_{,i}}$ is defined as the average residence time calculated using the set of meshes with $\rho \geqslant ~{{\rho}_{i}}$ and ${{\rho}_{P}}$ being the finest mesh:

Equation (4)

We use the requirement that a minimally sufficient mesh must satisfy ${{s}_{\overline{{{\tau}_{\text{R}}}},i}}<4.0\centerdot {{10}^{-11}}\,\text{s}$ . In figure 9 we plot ${{s}_{\overline{{{\tau}_{\text{R}}}},i}}$ as a function of $\rho $ (along with ${{\tau}_{\text{R}}}$ for illustrative purposes). The uncertainty associated with ${{\tau}_{\text{R}}}$ is shown as a horizontal black line; the point at which ${{s}_{\overline{{{\tau}_{\text{R}}}},i}}$ crosses below this line, at $\rho =46$ , is marked with a vertical dotted line, and we can take $\rho =46$ to be minimally sufficient for the present system of study.

Figure 9.

Figure 9. Residence time, ${{\tau}_{\text{R}}}$ , and standard deviation of residence time, ${{s}_{\overline{{{\tau}_{\text{R}}}},i}}$ , as described in equation (4), for the ground state configuration of the  ∑5(2 1 0) GB in copper as a function of perturbation mesh, $\rho $ . The black solid line indicates the uncertainty associated with ${{\tau}_{\text{R}}}$ as determined from figure 6, and the vertical dotted line at $\rho =46$ indicates the threshold where the standard deviation of the residence time crosses below this uncertainty.

Standard image High-resolution image

Labeling kinetic events

In figure 10, we show several examples in which we have used NAUTY to identify unique kinetic events and LAEs for the energy landscape search of the  ∑5(2 1 0) grain boundary for several example combinations of the R and a parameters (recall that R is the radius of the truncated graph and a is the cutoff distance for an edge). For the purposes of this comparison, we have only used meshes with ρ in the most relevant range: 46–562. To allow for a direct comparison with the results of figure 6, we have held d  =  4, $\delta =~0.5$ ${\mathring{\rm A}}$ , and ${{d}_{1}}=0.75\,{\mathring{\rm A}}$ constant.

Figure 10.

Figure 10. Residence time of the ground state  ∑ 5(2 1 0) grain boundary in copper with increasing perturbation mesh refinement is plotted for different choices of R and a using NAUTY to identify unique kinetic events and LAEs. The color of the points indicate the edge cutoff value a and the line color indicates the radius of the truncated graph R.

Standard image High-resolution image

From this figure we see, as was observed with our method above, that the residence time appears to converge for perturbation meshes with greater than about 500 directions. The sensitivity of the residence time to the choices of R and a are notable, as the spread of calculated residence times in the converged region, ~$1\centerdot {{10}^{-10}}$ s, is about the same as the residence time itself, and there is no obvious way of determining which combination of R and a lead to the true residence time of this system.

In figure 11, we plot the same data as shown in figure 10 and compare it with the data from figure 6 with $\epsilon =0.025\,\,{\mathring{\rm A}}$ . From figure 11, we note that residence time calculated using the present method lies within the same range as that calculated using NAUTY, but with a tighter uncertainty (under half). In all cases, the meshes with greater than 500 directions show a converged residence time behavior.

Figure 11.

Figure 11. Comparison for calculated residence times for the ground state configuration of the  ∑5(2 1 0) GB in copper for increasing perturbation mesh densities where kinetic events are determined using NAUTY (see color bar) and our present approach (black). Error bars based on uncertainty calculated from figure 6 are overlain.

Standard image High-resolution image

In addition to having a tighter uncertainty, our approach has the advantage of not depending on the arbitrary choice of parameters R and a. Because it relies instead on the physical definition of a kinetic event (the displacement of atoms) in distinguishing between events, we believe it provides a reasonable means of determining the residence time in an off-lattice atomic configuration.

Specification of a local atomic environment

To determine the minimum number of closest neighbor atoms, M, required to specify an LAE for our system, we have calculated the residence time of our example system using values of M from 12 to 64. Only the atoms in or near to the grain boundary were of interest in the kinetic analysis, thus atoms more than one nearest neighbor spacing from the GB were not searched for kinetic events. In this highly ordered GB, the number of unique LAEs in the GB region ranged from 8 to 22 for the investigated values of M. The results are shown in figure 12. The error bars denote the uncertainty in the residence time calculation. Due to the minimal fluctuation observed in the residence time across the full range of LAE sizes considered, we conclude that 12 closest neighbors is sufficient for the definition of the LAE in this case.

Figure 12.

Figure 12. Residence times for the ground state configuration of the  ∑5(2 1 0) GB in copper for increasing LAE sizes. Error bars based on uncertainty calculated from figure 6 are overlain.

Standard image High-resolution image

Illustrating the deterministic search

In figure 13 we show the perturbation directions that resulted in successfully finding a saddle point for three different LAEs for perturbation meshes with ρ  =  128 and ρ  =  562. Perturbations that led to a successful ART search are plotted on a unit sphere, or 'transition globe', and colored according to the activation energy of the corresponding kinetic event. For clarity in visualization, perturbations of all magnitudes are collapsed onto a single globe.

Figure 13.

Figure 13. Perturbation directions that resulted in the identification of a saddle point are plotted on a unit sphere, colored according to the activation energy of their corresponding transition state, creating a 'transition globe'. The transition globes of atoms in three different LAEs are shown for perturbation meshes with $\rho =128$ and $\rho =562$ .

Standard image High-resolution image

From figure 13, we see that in the cases where a denser mesh was used, there are 'continents' of perturbation directions that result in what appear to be the same kinetic event. In some cases, these continents are quite large, indicating that a broad range of perturbations will lead to convergence to the same saddle point. In other cases, only a narrow set of directions lead to a given event. It is these latter kinetic events, if sufficiently low in activation energy, that dictate the required mesh density of a system and which highlight the risks of declaring any search 'complete': there is always some chance of missing the critical perturbation that might find a low energy (highly kinetically significant) transition.

To investigate the potential relationship between the activation energy of a kinetic event and the ease with which it can be converged upon, in figure 14 we plot the solid angle, in steradians, that a 'continent' on a transition globe occupies as a function of its activation energy. In general, we see that across the whole range of activation energies there are many events that will only be observed if triggered by a relatively localized perturbation. Conversely, there is a subset of transitions for which a broad range of perturbations will converge to what appears to be the same saddle point, as can be observed in figure 13. There is a possible correlation between higher energy saddle points with larger continent areas. One explanation of this trend might be that higher energy saddle points are less likely to fall back down into the initial minimum state during convergence than lower energy activated states, so a wider set of initial perturbations can converge to the same saddle point in these cases.

Figure 14.

Figure 14. The solid angle in perturbation space on the transition globe that leads to a given kinetic event is plotted (in steradians) with respect to activation energy.

Standard image High-resolution image

Conclusions

Towards the goal of performing reliable kinetic Monte Carlo simulations on off-lattice systems, we have shown that it is possible to calculate the residence time of a system by deterministically searching the events accessible to atoms in unique local atomic environments (LAEs). If a sufficiently dense set of perturbations are used in the search, a full set of kinetically relevant activated pathways accessible to a system can be cataloged. In order to achieve this goal, this paper presented discussion of several key ingredients:

  • Distinguishing kinetic events: we presented an exhaustive approach to defining kinetic events by the displacements of atoms and used the numerical uncertainty in atomic position to differentiate between events. We compared this method to a previously developed topological approach.
  • Performing complete searches on an energy landscape: we presented a deterministic search method for achieving a converged residence time calculation that made use of an optimizable 3D mesh of perturbation directions, rather than random perturbations, to initiate energy landscape searches.
  • Specifying local atomic environments (LAEs): we defined a LAE by the positions of its neighbors and showed that, for the relatively simple system being considered here, the positions of the 12 nearest neighbors were sufficient to define a unique LAE.

In demonstrating the effectiveness of this method, we have mapped the energy landscape of the ground state  ∑5(2 1 0) GB in copper, a simple, symmetric tilt GB. We have shown that residence time calculated for this system converges independently with each of the mesh parameters of the deterministic search. We have shown that the previously implemented topological method used to identify unique kinetic events and the present method result in similar residence time calculations and similar convergence behavior with the perturbation mesh parameters. However, the present method's uncertainty in the residence time is less than half that of the topological approach. Additionally, since our method makes use of the physical definition of a kinetic process (the displacement of atoms) in comparing kinetic events, we have improved confidence in our calculation of residence time over the previous approaches, which rely on the selection of arbitrary parameters. In future work, this approach can be implemented in a KMC method to study the kinetic behavior of ordered defects such as the simple grain boundary considered here.

Acknowledgments

This study was supported by the US National Science Foundation under contract CMMI-1332789. KCA acknowledges support from a DOE Computational Science Graduate Fellowship under Grant No. DE-FG02-97ER25308 and support from the Fannie and John Hertz Foundation.

Please wait… references are loading.
10.1088/0965-0393/24/6/065014