ACS Publications. Most Trusted. Most Cited. Most Read
Lateral Interactions of Dynamic Adlayer Structures from Artificial Neural Networks
My Activity

Figure 1Loading Img
  • Open Access
C: Chemical and Catalytic Reactivity at Interfaces

Lateral Interactions of Dynamic Adlayer Structures from Artificial Neural Networks
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

The Journal of Physical Chemistry C

Cite this: J. Phys. Chem. C 2022, 126, 12, 5529–5540
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpcc.1c10401
Published March 17, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

Lateral interactions are a key factor in the correct description of adsorption isotherms relevant to heterogeneous catalytic reactions. To model these lateral interactions, a large number of monolayer structures have to be investigated, far exceeding the limitations of conventional techniques such as density functional theory. We have developed a new hybrid neural network model that can substitute the electronic structure calculations for these monolayer structures, without significant loss of accuracy. The low computational cost of this model allows the study of the adlayer structures close to industrial operating conditions. Lateral interactions are found to increase at elevated temperatures as a result of increased adsorbate mobility, and this contribution is found to be key in unifying theoretical and experimental observations. We show that the inclusion of dispersion interactions in stabilizing the adlayers is necessary to obtain correct predictions for both isotherms and adsorption site distributions.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2022 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Microkinetic modeling allows us to describe the reaction rates of complex networks of chemical reactions. It is widely used in heterogeneous catalysis to describe the kinetics of surface reactions based on elementary reaction steps involving surface-adsorbed intermediates. (1,2) Typically, the reaction energetics (adsorption energies and activation barriers) are obtained from density functional theory (DFT) calculations of single adsorbates. However, in practice, surface coverages can be high due to mild reaction temperatures or high reaction pressures. Lateral interactions between adsorbed species cannot be neglected when the aim is to accurately predict the kinetics of the overall catalytic reaction. (3,4) Describing such lateral interactions at the DFT level is preferred but challenging because of the inherently high cost of DFT combined with the large number of adlayer configurations that must be studied.
Several notable advances have been made in deriving coverage-dependent reaction energetics. Within the microkinetic approach, coverage-dependent adsorption energies can be obtained through interpolation of DFT data. (4−17) Self-consistent feedback approaches have also been developed, wherein the DFT coverages are updated based on the results from the microkinetic simulations. (18−21) Adsorption rates can alternatively be obtained through kinetic Monte Carlo simulations. (3,22−26) Interactions between adsorbates are then described using sets of multibody interactions, which can again be determined by DFT. We observe that these state-of-the-art methods, while successful in reproducing experimental results, still rely on approximations. Microkinetic models utilize fixed interpolation functions. Multibody models are typically truncated to two-body terms to avoid tabulation of a large number of configurations. (3,22,23) Even for self-consistent approaches, computational limitations restrict the DFT model to discrete coverages, thereby limiting the resolution of coverages in the microkinetics. (18,22) To improve on existing modeling efforts, we therefore seek a method that can accurately describe lateral interactions without reliance on approximations.
Computational limitations in the scale-up of quantum chemical (QC) methods are common in many branches of computational chemistry. (27−29) Traditional solutions use semi-empirical models to extrapolate energy-structure relationships from available QC data, as with (reactive) force-fields and tight-binding models. (30−32) These methods are based on potential models, enhancing computational efficiency at the cost of accuracy relative to full electronic structure calculations. In recent years, data-driven machine learning (ML) approaches have emerged as an alternative to traditional force fields. (29,33,34) While the semi-empirical models are based on a set of rigorous equations wherein the parameters are tuned to provide a good fit with a given training set, data-driven methods do not a priori assume any kind of rigorous mathematical form. Instead, they utilize a flexible set of matrix equations that can adopt any kind of known mathematical function. (35) The rigorous mathematical form of semi-empirical methods leads to forms of descriptive biasing, which are absent in data-driven methods. From the user perspective, the data-driven nature lends to a more methodological approach of model construction. This shifts the investment cost to the data generation stage, facilitating better scalability.
Early examples of the use of ML in chemistry include the interpretation of experimental and spectroscopic data (36,37) and the discovery of new materials and chemicals by extraction of correlations between chemical properties. (38−40) The attention then shifted toward application in atomistic modeling. The initial focus on construction of ML-assisted semi-empirical models (41) has now been largely replaced by full reconstruction of potential energy surfaces (PES) determined by QC. (42,43) A breakthrough in this development can be attributed to high-dimensional neural network potential (HDNNP) (33,44−46) models, which facilitate tackling problems with higher degrees of freedom. (27−30,47−53) Recently, Gaussian approximation potentials (34,54,55) are increasing in popularity, boasting high accuracy and robustness, although at higher computational cost than HDNNP.
In this work, we present a hybrid ML model for the description of adlayer compositions in heterogeneous catalysis, which combines features from both kernel-based atomic potentials (such as HDNNP) and quantitative structure-property relationship (QSPR) networks. (38−40,56−58) A methodology based on molecular degrees of freedom allows for automated construction of small, cost-efficient datasets. The effectiveness of our method is shown for a model system composed of flat (001) and stepped (211) cobalt surfaces, allowing comparison to available experimental and theoretical data. (4,59) By leveraging the low computational cost of this workflow, we can critically assess how the choice of exchange-correlation functional and dispersion corrections in the development of the dataset influences the description of the adsorption layer in connection to the experiment. Furthermore, we show that the lateral interactions depend on temperature. Here, increased adsorbate mobility can create dynamic adlayer structures, affecting both adsorption site distributions and lateral interactions. This mobility effect is found to be a crucial aspect for correctly describing experimentally observed adsorption trends.

Methods

Click to copy section linkSection link copied!

Basis Functions and Network Architecture

The ML approach employed here is that of the artificial neural network (ANN). (60) The general mathematical expression of an ANN can be written in recursive form
(1)
The ANN entails a series of sequential, alternating basis (W̿) and non-linear (F) transformations of a set of input parameters 0. The basis transformations map the input space onto a new space spanned by compound variables. The non-linear transforms and bias shifts () are then used to connect the transformed input to the desired output. By tuning the transformations, the overall mapping between the network input and output can be defined to represent a correlation of the data. During this process, the network has the freedom to define its own intermediate compound variables and is therefore free from any framework bias. Chaining multiple transforms allows defining correlations as a series of connections between different sets of compound variables, increasing the complexity of the model. Because of the periodic linear transforms, linearized optimization of the ANN variables is possible, resulting in computationally efficient training algorithms.
The purpose of the ANN in the present study will be to predict the energy of the atomic configurations, i.e., of the adlayer structures, as is typically done using QC. As the number of adsorbates may vary, the dimensionality of the input space must be invariant to the number of atoms in the system. The system is therefore decomposed into a set of atomic ANN each describing the contribution to the system energy from a particular atom. One atomic ANN will be constructed per element to allow transferability between systems.
(2)
While it is possible to use traditional decomposition schemes (61) to determine the atomic energies EA, here, the choice of decomposition was integrated as a linear transform in the ANN to limit model bias.
To distinguish between different atomic configurations, it is necessary to include structural descriptors to the input space. In the literature, many successful approaches have been reported for the geometric description of adsorbates using ML. (27,54,62−66) A common factor between these approaches is their use of multibody interaction kernels as a basis for describing the interactions between atoms. This is seen with, e.g., the atom-centered symmetry function (ACSF) class of methods, (27,45,46,54,62) which employ interaction potentials based on modified atomic distribution functions, or methods such as the smooth overlap of atomic positions (SOAP), which employ expansions of coordination environments into more traditional basis functions (Gaussians and spherical harmonics). (54,55,63,66) Many further variations exist, e.g., partial radial distribution functions, (64) Coulomb matrices, (65) and internal coordinate descriptors. (42,65,66)
In this work, we employ an approach inspired by ACSF/SOAP. Orientational invariance is achieved by defining a basis sampling radial and angular distribution functions, which are evaluated per atom. The signal intensity per function can be used to define the relative atom density distribution around the reference atom, characterizing the coordination environment. This process is illustrated in Figure 1.

Figure 1

Figure 1. Illustration of radial (a) and angular (b) coordination descriptors. Radial distributions are defined with respect to the atom marked as (i), angular distributions relative to the axis between atoms (i,j). Different descriptors are shown using various colors. The graphs show the signal values as a function of the descriptor parameters. (a) Left: cutoff function gc for various cutoff radii. Right: radial coordinate descriptors Ri for various decay rates. (b) Angular coordinate descriptors Ai for various angular decay parameters.

First, the distributions are limited to a local coordination environment by introducing a smooth cutoff function (34,62)
(3)
where r represents the interatomic distance and rc represents the radius of a cutoff sphere. For the radial sampling, this local environment is split into a set of subshells using a basis of Gaussians, defined by different decay rates η as given by
(4)
Within each radial shell, the distribution of angles between neighbors is sampled
(5)
with λ and ζ controlling the range of angles sampled by each descriptor. It is chosen here to omit the interaction of the neighbors gc(rjk) as it was found otherwise to result in low signal intensities for most angular functions. Compared to most ACSF methods, (45,46,62) these descriptors are not selective or weighted toward particular chemical elements.
A limitation of the described approach is the absence of explicit electronic information in the network input since the electronic effects must be learned implicitly as part of the atomic decomposition. To improve the network, the geometric description above is hybridized with a classical QSPR approach to allow better differentiation of local electronic character and to accelerate and improve the fitting procedure. The architecture of this hybrid network is shown in Figure 2. Within QSPR, the pattern-recognition capabilities of ML are used to find correlations directly between system properties without relying on atomistic models. The direct predictive quality of QSPR can now be leveraged to let the ANN make deductions on reactivity and stability based on physically relevant parameters, reminiscent of the force-field approaches. Several electronic parameters will be provided to supplement the structural descriptors. Two different forms for these electronic parameters were considered. First is the use of atomic parameters εi determined entirely by the nature of the atom, the common descriptors used in QSPR. Examples include atomic charge, magnetic moments, or electronegativity. Alongside these, effective local parameters are used that modify the atomic descriptors by accounting for the presence of the neighboring atoms in the local coordination environment. The latter therefore represents field- or density-averaged properties specific to the molecular environment. Different representations are used for scalar parameters
(6)
or vector properties
(7)
where H is the Heaviside function, which is used to limit the interaction range to the cutoff volume without introducing a radial dependence.
(8)

Figure 2

Figure 2. Schematic depiction of the hybrid network architecture. Separate geometric and electronic depictions of a molecular geometry are obtained, from which the atomic networks reconstruct the total energy. This compares to classical approaches in which only the geometric representation is used. Geometric and electronic information are communicated to the atomic networks using a set of basis functions, each represented as a single input node. One atomic network is constructed per chemically unique element (here shown in green, yellow, and red).

For determining the network parameters, a training dataset is constructed for which the distribution of descriptor values can be analyzed. As the dataset only provides a finite representation of the full input space, it yields an approximation of the true descriptor distributions. This allows us to determine the contributions of the individual basis functions to the representations of our samples. Basis functions that do not contribute significantly to the resolution of the network can be filtered prior to the network fitting, allowing for fully automated construction of an optimized basis set specific to the system of interest: (I) Basis functions with low signal intensity, thus poor numerical resolution, are filtered first by imposing a minimum intensity requirement, which was chosen here to a value of 0.01 (corresponding to a minimum of 10% of the maximum signal intensity for the normalized descriptors); (II) sample homogeneity is restricted by imposing a minimum deviation of 2% of samples at 0.01 dispersity to avoid signals with poor differentiability; (III) to resolve overlapping signals, a continuous similarity recurrence (CSR) metric is introduced.
(9)
with two descriptors denoted as signals Ga and Gb for a dataset of size ζ. A large CSR overlap between two signals would indicate that their output differs significantly for most samples. Signals that have low CSR values for all descriptor pairs can be characterized as redundant and are removed from the basis set, while signals with high CSR values highlight symmetry classes that may require extension of the basis set. CSR thresholds of 5% were used in this work. This approach allows for informed construction and optimization of a basis set, without the need to repeatedly retrain the network.

Automated Dataset Generation

The quality of an ANN depends strongly on the dataset used for its construction. If segments of the configurational space are omitted, then extrapolations into this domain may result in unpredictable behavior. As such, when constructing a dataset, it is critical that all domains that will be accessed in target studies for the ANN have been explored. On the other hand, it is desirable to make the dataset as small as possible. Each sample point requires a DFT calculation to be performed, making this the computationally most expensive stage of the model construction. Because the networks are system-specific, requiring construction of a new network and corresponding dataset for each system, it is critical for this process to be as cheap as possible if widespread application is to be achieved.
To balance the dataset cost-effectiveness, a uniform sampling in feature space is proposed. Herein, the internal degrees of freedom of the molecular ensemble will be used to define the configurations for the dataset. This process is illustrated in Figure 3 with algorithmic details provided in the Supporting Information. Our method is system-independent, facilitating fully automated generation of datasets once a system is defined. By sampling the chemical configurational space, this should limit the overlap within the dataset as much as possible by eliminating symmetry-equivalent configurations and considering equal sampling of the translational, rotational, and vibrational subspaces. The latter is especially important considering that both single-adsorbate states and higher monolayer coverages will be included in the same dataset. Additionally, we can readily quantify the extent to which our model is applicable for a particular system. For example, if it is chosen to restrict sampling of vibrations up to a (small) maximum displacement, then the model should not be used in simulations involving stronger distortions of the molecular geometry.

Figure 3

Figure 3. Different stages of dataset construction: I. Translational states along a geometrically bounding potential surface. II. At each translational site, different rotations are obtained by rotating the molecule along two internal coordinates. III. Each rotation is expanded into a set of vibrational states. IV. Combinations of these states are used to construct a set of adlayers. At this step, translations are extended over a set of support grids to alleviate symmetry constraints. Note the sparser sample density over these extended surfaces.

Results and Discussion

Click to copy section linkSection link copied!

We will first demonstrate the performance of our method for a well-defined system of CO adsorption on metallic hcp-cobalt (001) and (211) surfaces. Cobalt-catalyzed CO conversion is a crucial step for industrial processes such as FTS, and as such, both experimental and theoretical reference data are readily available. (1,4,59,67−69) Two different surface facets were considered to compare the effect of surface morphology and network performance. In the following segment, we will first validate the generated ANN models followed by a performance comparison against other methodologies. Afterward, we will apply our ANN to describe lateral interactions for CO adsorbed on Co(001) and Co(211) surfaces.

Machine Learning

Datasets were generated for both surfaces using the generation process described in the methods section, resulting in 551 samples for the (001) surface and 851 samples for the (211) surface. Herein, the larger size of the dataset of the (211) surface corresponds with the larger surface area for the primitive cell. DFT calculations were performed using a PBE exchange-correlation functional (70) to allow direct comparison to the literature. (4,5) The CSR was used to optimize the basis set followed by a hyperparameter screening to determine the network architecture. Details of these processes can be found in the Supporting Information.
The resulting networks reproduced the DFT data with root-mean-square-errors (RMSEs) of 0.22 meV/atom for Co(001) and 0.31 meV/atom for Co(211) and biases less than 0.01 meV/atom. This is well within the typical measure of chemical accuracy of 1 kcal/mol. This threshold is typically used in the context of comparing experimental and theoretical results, considering both numerical and experimental accuracy. As such, variations less than 4 kJ/mol will fall within the standard uncertainty of any theoretical method. For energy calculations, this corresponds with approximately 0.8 meV/atom for typical system sizes. Here, a per-atom accuracy metric is standard in ANN studies, given that the system sizes vary between samples. We see that our networks fall well within this margin.
Training, validation, and test subset error distributions have comparable mean and variance (shown in Table S1) indicating an absence of overfitting. The extrapolation behavior of our networks was tested by generating a secondary reference dataset with a sample density twice as large as the original set. The distribution of errors of this reference was found to be similar to that of the original dataset, even after filtering out samples shared between the datasets (0.24 meV/atom for Co(001) and 0.31 meV/atom for Co(211), see Table S2). This indicates correct extrapolation beyond the training set employed by the network. The RMSEs of the single-adsorbate and adlayer subsets were found to differ by less than 0.01 meV/atom, indicating that our models do not bias toward either set of samples.
The quality of an ANN is proportional to the amount of information contained in the training data. As both surface models contain the same chemical elements, we have explored the effect of combining both surfaces into a single network to benefit from an extended dataset. Similarly, we have explored addition of atomic carbon and oxygen data to observe if a similar trend holds for the adsorbate. The results are summarized in Tables S12–S14. A substantial improvement of the quality over the single-surface and single-adsorbate networks is observed. For the complete dataset, the overall RMSE decreases from 0.27 to only 0.18 meV/atom while subset biasing remains minimal. The skew is seen to decrease due to the smaller effective stepsize of the optimizer. This improvement in fit quality when adding additional adsorbates to the dataset follows as the distribution of coordination descriptors broadens. The effect levels off as the number of unique adsorbates increases, either due to the saturation of the distribution or numerical limitations in descriptor resolution. When adding different surfaces, the improvement in quality is more pronounced on account of the dissimilarity in descriptor distributions for either surface. The most prominent improvement is observed for the (211) surface. As the (211) dataset employs a sparser sampling than the (001) dataset (because of computational considerations), the former benefits the most from the extended number of samples. The combined datasets were also observed to improve the uniformity between training, validation, and test set error distributions owing to the more diverse array of chemical configurations in the input. This effect was most notable for CO as the addition of C and O data helps to establish atomic reference levels within the ANN framework.
Compared with the literature, other ANN methodologies have employed trajectory data from molecular dynamics (MD) as their training set, yielding on the order of 104 to 105 samples. (29,30,33) The approach taken in this work allows for much smaller datasets by exploiting chemical principles. The use of single-point calculations instead of local energy minima furthermore helps in capturing the rough features of the potential energy surfaces at lower sampling density by avoiding the bias induced by applying geometry optimizations. The more disperse sampling in our method avoids the existence of large extrapolation regimes, where the ANNs are known to perform poorly, (28,29,33) thereby obtaining solid predictive performance.

Effect of Network Hybridization

Key to supporting the small datasets is the proper mapping of the descriptor spaces as well as the resulting resolution for distinguishing samples. The hybridization of our networks was found to be the responsible factor in this regard. In Figure 4, we have compared the performance of our hybrid ANN to a selection of networks employing a purely structural basis set. Even after reoptimization of the network architecture, these models show notably larger errors. Increasing the size of the dataset used for training these networks does not significantly impact the observed error distributions, indicating that the network has reached a limit in its ability to learn additional details of the PES. Indeed, a more profound effect is observed when switching to an extended basis set (obtained by relaxing the CSR thresholds). Chemically accurate fit qualities are only obtained by combining both the extended basis set and extended dataset, in line with the other networks reported in the literature. (54,61) In the context of traditional HDNNP methods, increased architectural sizes and more narrow application foci of the datasets generated using MD suggest the presence of redundant parameters in the structural networks. (28−30,33,47) These networks require much larger datasets than the hybrid networks used here, partially on account of their larger size. The larger parameter set allows easier parametrization of the provided data but in turn increases the risk of overfitting that becomes critical once we exit the interpolation domain. This is supported by the trends observed here for the basis set size dependence and the observation of smaller extrapolation errors for the hybrid networks. The two network types therefore show different compatibilities with the small datasets.

Figure 4

Figure 4. Error distributions for a selection of CO/Co(001) networks. Model (II) is a purely structural network with the same architecture as the single-transform hybrid ANN detailed in the main text. Model (III) is a structural network with a reoptimized architecture with dimensionality 20 and 2 transforms. Model (IV) uses the reoptimized architecture but with the extended dataset (II) as the training dataset. Model (V) uses the reoptimized architecture along with an extended descriptor set (trained using dataset (I), parameters are provided in Figure S2). Model (VI) combines the extended dataset (II) with the extended descriptor set of model (V). Additional details are provided in Table S11.

Comparing the descriptor distributions across the samples in our datasets, the electronic basis functions appear to do well in distinguishing the surface from bulk states and empty sites from adsorption sites based on the electronic environment. The electronic basis functions are largely uncoupled from the geometric ones, reducing underlying interdependencies between network inputs. As interactions of Co–CO can be traced specifically to surface states, electronic augmentation helps achieve the good predictive behavior and high accuracy of the hybrid networks.

Lateral Interactions

Using the ANN developed in the previous section to describe the potential energy as function of the adlayer configuration, we here demonstrate its application in describing the average adsorption energy of CO as function of the surface coverage, temperature, and pressure. Such a description aids in advancing microkinetic modeling studies by incorporation of lateral interactions, allowing for more realistic simulations under typical working conditions. For example, reaction barriers are conventionally treated in a static fashion in such simulations but can be made dependent on the coverage to model the effect of lateral interactions.
To compute the interaction energy at elevated temperature, Monte Carlo (MC) simulations were performed as detailed in the Supporting Information. Since the ANN is orders of magnitude cheaper to evaluate than DFT, it can easily serve as the underlying potential in these simulations. The ANN uses a 15 × 15 unit cell for Co(001) and a 20 × 20 unit cell for Co(211), in contrast to the 3 × 3 and 4 × 4 unit cells for DFT. These larger unit cells support a higher resolution of coverages. For our analysis, we have used PBE-DFT data to train our networks in line with earlier literature studies. (4,5) The effect of the exchange-correlation functional in experimental comparison will be revisited in the next section.
Figure 5 shows the results for the average adsorption energy of CO as a function of the coverage for the Co(001) and Co(211) surfaces at 0, 300, and 500 K. A comparison is made against DFT results from geometry optimizations, which corresponds to the 0 K system. From this figure, it can be seen that the ANN closely matches the result found by DFT. Additionally, the ANN has now provided us with the interaction energy at the intermediate coverages without resorting to interpolation methods.

Figure 5

Figure 5. ANN potentials for the CO adsorption energy in the presence of lateral interactions over the left (Co(001)) and middle (Co(211)). Profiles at 300 and 500 K are derived from temperature-corrected adlayer energies obtained through MC. Hexagonal markers represent reference DFT minimum energy points. Right: onset coverage at which temperature-dependent lateral interaction potentials deviate from the 0 K potential by >1 kJ/mol.

At temperatures above 0 K, the adsorbates possess sufficient kinetic energy to traverse the PES, allowing them to access states above the energy minimum. This results in an increase in the average energy as readily seen in Figure 5. Below a fractional coverage of around 0.2 ML, the average energies at 0, 300, and 500 K are very similar. At low surface coverage, the distance between the neighboring adsorbates is too large for any noticeable lateral interactions. With increasing coverage, the distance between the adsorbates is smaller and an increased lateral interaction penalty is observed. This lateral interaction penalty increases with temperature as the adsorbates occupy a larger effective volume on the catalytic surface given their increased lateral mobility.
We can hence define a low and a high coverage domain, considering the temperature effect on the lateral interactions. In the low coverage domain, the distance between the adsorbates is sufficiently large that the increased occupational space of the adsorbates due to their enhanced mobility does not result in a noticeable difference in the lateral interactions as compared to the 0 K situation. The high coverage regime starts when this difference exceeds a set threshold. In Figure 5, this boundary surface coverage, marking the transition between the low and high coverage regimes, is shown as a function of temperature for a(n arbitrary) numerical threshold of 1.0 kJ/mol. The nonlinearity of these profiles relates to the exponential dependency of the hopping behavior of the adsorbates with temperature and the changes in the symmetry patterns of the adlayer with varying coverage.
Comparing the Co(001) and Co(211) surfaces, a few important differences can be observed, which can be directly attributed to the surface topology. Since the Co(001) surface is completely flat and rather homogeneous in terms of adsorption sites, we observe a gradual increase in the average adsorption energy as function of the coverage. For the Co(211) surface, which is more heterogeneous in terms of its adsorption site distribution, a more step-like curve is found. Below ∼0.35 ML coverage, the CO molecules can be distributed among a set of similar threefold adsorption sites (see Figure 6). Once these sites are all occupied, CO adsorbs on top sites in between the threefold sites, giving rise to a rapid decrease in the average adsorption energy. The next sharp increase is observed around 0.6 ML. Due to the large number of CO species adsorbed on the surface, the configuration where predominantly threefold sites are occupied becomes unfavorable as this configuration exhibits strong lateral repulsion. To mitigate this repulsion, the adlayer topology changes completely and predominantly bridge and on-top sites are occupied. A similar rapid change in the surface configuration is observed around 0.8 ML where the dominant adsorption mode changes from bridged to on-top.

Figure 6

Figure 6. Minimum energy structures for CO/Co(211) predicted with the ANN. From left to right: 0.2, 0.4, 0.7, and 0.9 ML.

The results as shown in Figure 5 can be directly utilized to describe lateral interactions in a mean-field manner, allowing these to be included in a microkinetic model. In Figure 7, the equilibrium CO coverage as function of temperature is shown. Herein, a comparison is made between our simulated results with and without the lateral correction potential and experimental data obtained from the literature both at ultrahigh vacuum and under atmospheric pressure. (59) Without lateral interactions at 1 bar, an unrealistic coverage of 1 ML is predicted below T = 700 K. Including the lateral interactions, the surface coverages modeled at p = 10–7 and 1 bar are in good agreement with experimental results. The deviation with the experiment at p = 10–7 bar in the temperature range near 200–350 K is highly sensitive to the calculated adsorption energy for CO. This deviation can be reduced by a small adjustment to the CO adsorption energy within the order of accuracy of MC.

Figure 7

Figure 7. Left: equilibrium CO coverage on the Co(001) surface, determined from a microkinetic model (MKM) composed of a single-reversible adsorption step with coverage-dependent adsorption energies. (Details are provided in the Supporting Information). The HK model omits lateral interactions. Markers show experimental data published by Weststrate et al. (59) Right: effect of the temperature correction applied to the lateral interaction model at 1 bar.

The sharp drop from 0.7 to 0.5 ML coverage at T = ∼250 K at p = 10–7 bar and T = ∼350 K at p = 1 bar is attributed to a sudden rearrangement of the CO molecules due to a rapid increase in the lateral interactions as the result of the increasing temperature. Multibody decomposition of the ANN energies (Table S21) revealed that the adequate modeling of this transition requires including interactions between CO molecules over multiple coordination shells (nearest, next-nearest, and even further neighbors). Pair interactions were found to be sufficient at describing the system only at low coverages <0.2 ML, when the surface is sparsely occupied and clustering of adsorbates is rare. Successive addition of triplet- and quartet-interactions converges the model toward the full ANN solution at higher coverages >0.6 ML (Figure S16). Here, densely packed structures are formed, for which short-range interactions are expected to dominate. For an intermediate regime of 0.2–0.6 ML, however, accurate decomposition of the complex surface structures would require incorporation of many higher-order terms. This task has been successfully delegated to the ANN.
The decrease in binding of the adsorbates due to their increased mobility on the surface was shown to be strongly non-linear in the coverage as a result of changes in the occupation of different adsorption sites. This is why a simple shift of the 0 K interaction profiles (Figure 5) is not observed. It is noted that the temperature effect up to 500 K is only in the order of 10 kJ/mol, primarily at higher coverages. The electronic interaction between the surface and adsorbate remains dominant in controlling the adlayer structure, and thereby the energy, on account of the relatively strong binding of CO to the metal. The temperature effect does become stronger when considering higher temperatures (>500 K), although under these conditions, the surface is predicted to be largely depopulated and the adsorbate–metal interaction again remains leading. Nevertheless, we have seen that the temperature-dependent lateral interaction potentials are critical for the correct prediction of the characteristics of the coverage profiles. As noted in Figure 7, the strongest deviations are observed precisely in the catalytically active regime, emphasizing the importance of this effect in modeling heterogeneous catalysis. This effect has not been investigated before. It is made possible here due to the low computational cost of the ANN method.

Effect of Exchange-Correlation Functionals and Dispersion Interactions

In the foregoing discussions, we have employed the PBE functional (70) for the investigation of the lateral interactions. For CO adsorption over transition metals, PBE is known to exhibit an overbinding effect and hence an incorrect prediction of the adsorption sites. (4,71) It is therefore somewhat surprising that the correct lateral interactions are predicted for this system as co-adsorbate interactions depend strongly on the site locality. To investigate this, several alternative exchange-correlation functionals were considered (Tables S18 and S19). Of these, the AM (72) and PBEsol (73) functionals were quickly discarded as they showed comparatively large errors in the prediction of the Co-HCP lattice parameters. The revPBE (74) and rPBE (71) functionals yielded similar predictions for the Co-HCP lattice parameters as PBE. As the minimum energy structures and the lateral interactions for revPBE and rPBE are nearly identical (Figure 8 and Figures S11–S14), we chose revPBE for further analysis as it provided the closest agreement to experimental values.

Figure 8

Figure 8. Difference in the lateral interaction potentials for Co(001) predicted for different XC functionals at T = 0, 300, and 500 K. Average adsorption energies are expressed relative to the adsorption energy of CO on the empty surface.

Using the revPBE functional, the experimentally observed top-adsorption below 0.4 ML is reproduced. Upon the increase in the monolayer coverage, a shift toward bridge states is found in contrast to experimental works, (59,68) which suggest a shift toward hollow sites first. However, it is noted that, depending on the degree of electron localization of the GGA, occupation of bridge-top and bridge-hollow intermediates can be observed, which matches the experimentally observed IR frequencies well (1820–1940 cm–1 compared to 1850–1910 cm–1). (59) At elevated coverages (0.5–0.75 ML), the non-temperature-corrected model predicts exclusively top sites, which are the dominant sites found in the experiment. With addition of finite temperature effects, CO is predicted to also adsorb on bridge sites, as is reflected by the peak broadening seen in the XPS signals at these coverages. (59,67,68)
While revPBE now produces an agreeable result for the monolayer geometries, the same cannot be said about the observed lateral interaction potentials. As seen previously in Figure 8, above 0.4 ML coverage, the lateral interaction penalty is now much larger than PBE. This is reflected in Figure 9, where the fractional monolayer coverage as function of temperature is shown, as lower surface coverages are obtained compared to the experiment. The difference between the revPBE and PBE functionals and thus between top–top and hollow–hollow interactions can be readily rationalized. For the hollow sites predicted with PBE, the surrounding cobalt atoms act to screen part of the lateral interactions. Hollow–hollow lateral interactions are much weaker than top–top lateral interactions (with two-body nearest-neighbor terms of ∼15 and ∼30 kJ/mol), yielding a stronger lateral effect for revPBE. The increased lateral interactions result in a shift toward lower coverages for the low-temperature regime, with the effect naturally leveling off as the surface depopulates.

Figure 9

Figure 9. Left: influence of different dispersion corrections on the revPBE lateral interaction potential for CO/Co(001). Integral adsorption energies have been expressed relative to the adsorption on the empty surface. Right: comparison of Grimme dispersion effects on the adsorption equilibrium for CO/Co(001) at 1 bar.

This discrepancy with the experiment suggests that there might be a secondary stabilizing effect present at higher coverages. Dispersion interactions were therefore considered as a possible source. Different dispersion models have been tested (Table S20), which show a good deal of variability in their predictions for the co-adsorbate interactions (Figure 9). The best match with the experiment is observed in this case for the Grimme D3 models, for which both D3(0) and D3(BJ) variants were considered. (75) The corresponding lateral interaction potentials show behavior that is intermediate to those of the non-dispersion-corrected PBE and revPBE potentials, mimicking addition of the PBE screening effect to the geometric models of revPBE. As shown in Figure 9, combining revPBE with dispersion now reproduces the adsorption equilibrium observed in the experiment.
We note that the same effect would not have been achieved for rerevision of the PBE exchange locality or PBE + rPBE hybridization as this is seen to coincide with a corresponding top-hollow shift in site occupation. Comparative predictions of the different dispersion models appear to follow similar trends as found in the literature. (75,76) The absence of the stabilizing effect for non-Grimme models could be attributed to density dependence of their dispersion coupled with the enhanced non-locality of revPBE. (74,76) While this does not provide a conclusive acknowledgement that the stabilization originates from dispersion interactions, given the strong sensitivity to their theoretical treatment, we note that, at the very least, an experimentally satisfying result could be obtained.

Conclusions

Click to copy section linkSection link copied!

A novel machine learning approach has been illustrated for the study of adlayer structures and lateral interactions. We have critically assessed the accuracy of our method and identified the factors that control this accuracy. The most important factor was found to be the inclusion of electronic basis functions, which were shown to enhance the basis set resolution, thereby allowing much smaller DFT datasets to be used. We demonstrated how these datasets can be efficiently generated using an algorithm based on uniform sampling of the molecular degrees of freedom. The models can easily be extended to include multiple surfaces and adsorbates, making them exceptionally suited for the study of catalytic systems. To facilitate this scale-up, the importance of a proper balance of the representation of the different species within the dataset was highlighted. As the dataset is structured in a chemically intuitive manner, it is straightforward to analyze the extent to which the ANN has been trained and to a priori identify missing features in the datasets.
The low cost of the ANN as compared to DFT makes it possible to use stochastic methods for tackling problems, which suffer from combinatorial complexity. This feature was demonstrated by constructing an ANN to describe CO adlayer compositions of Co(001) and Co(211) surfaces. Inclusion of finite temperature effects allowed for accurate reproduction of experimental temperature-coverage curves by means of mean-field microkinetics. When extending these studies toward systems containing multiple adsorbates, it may become possible to also include non-mean-field organization of the adsorbates using this methodology, which would allow us to significantly enhance the level of detail incorporated in microkinetic models. These microkinetic simulations would lend themselves to incorporation in macroscale simulations, fully bridging the multiscale domains.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcc.1c10401.

  • Overview of dataset generation methods, ANN training statistics, computational methods for DFT, MC, and MKM, monolayer geometries, and lateral interaction potentials (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Bart Klumpers - Laboratory of Inorganic Materials and Catalysis, Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, P. O. Box 513, 5600 MB Eindhoven, The NetherlandsOrcidhttps://orcid.org/0000-0001-9112-9069
    • Emiel J.M. Hensen - Laboratory of Inorganic Materials and Catalysis, Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, P. O. Box 513, 5600 MB Eindhoven, The NetherlandsOrcidhttps://orcid.org/0000-0002-9754-2417
  • Author Contributions

    The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This work was supported by the Netherlands Center for Multiscale Catalytic Energy Conversion (MCEC), an NWO Gravitation program funded by the Ministry of Education, Culture and Science of the government of the Netherlands. The Netherlands Organization for Scientific Research is acknowledged for providing access to computational resources.

References

Click to copy section linkSection link copied!

This article references 76 other publications.

  1. 1
    Chen, W.; Pestman, R.; Zijlstra, B.; Filot, I. A. W.; Hensen, E. J. M. Mechanism of Cobalt-Catalyzed CO Hydrogenation: 1 Methanation. ACS Catal. 2017, 7, 80508060,  DOI: 10.1021/acscatal.7b02757
  2. 2
    Huš, M.; Grilc, M.; Pavlišič, A.; Likozar, B.; Hellman, A. Multiscale Modelling from Quantum Level to Reactor Scale: An Example of Ethylene Epoxidation on Silver Catalysts. Catal. Today 2019, 338, 128140,  DOI: 10.1016/j.cattod.2019.05.022
  3. 3
    Prats, H.; Posada-Pérez, S.; Rodriguez, J. A.; Sayós, R.; Illas, F. Kinetic Monte Carlo Simulations Unveil Synergic Effects at Work on Bifunctional Catalysts. ACS Catal. 2019, 9, 91179126,  DOI: 10.1021/acscatal.9b02813
  4. 4
    Zijlstra, B.; Broos, R. J. P.; Chen, W.; Oosterbeek, H.; Filot, I. A. W.; Hensen, E. J. M. Coverage Effects in CO Dissociation on Metallic Cobalt Nanoparticles. ACS Catal. 2019, 9, 73657372,  DOI: 10.1021/acscatal.9b01967
  5. 5
    Zijlstra, B.; Broos, R. J. P.; Chen, W.; Bezemer, G. L.; Filot, I. A. W.; Hensen, E. J. M. The Vital Role of Step-Edge Sites for Both CO Activation and Chain Growth on Cobalt Fischer–Tropsch Catalysts Revealed through First-Principles-Based Microkinetic Modeling Including Lateral Interactions. ACS Catal. 2020, 10, 93769400,  DOI: 10.1021/acscatal.0c02420
  6. 6
    Grabow, L. C.; Hvolbæk, B.; Nørskov, J. K. Understanding Trends in Catalytic Activity: The Effect of Adsorbate–Adsorbate Interactions for CO Oxidation Over Transition Metals. Top. Catal. 2010, 53, 298310,  DOI: 10.1007/s11244-010-9455-2
  7. 7
    Lausche, A. C.; Medford, A. J.; Khan, T. S.; Xu, Y.; Bligaard, T.; Abild-Pedersen, F.; Nørskov, J. K.; Studt, F. On the Effect of Coverage-Dependent Adsorbate-Adsorbate Interactions for CO Methanation on Transition Metal Surfaces. J. Catal. 2013, 307, 275282,  DOI: 10.1016/j.jcat.2013.08.002
  8. 8
    Grabow, L. C.; Gokhale, A. A.; Evans, S. T.; Dumesic, J. A.; Mavrikakis, M. Mechanism of the Water Gas Shift Reaction on Pt: First Principles, Experiments, and Microkinetic Modeling. J. Phys. Chem. C 2008, 112, 46084617,  DOI: 10.1021/jp7099702
  9. 9
    Bajpai, A.; Frey, K.; Schneider, W. F. Comparison of Coverage-Dependent Binding Energy Models for Mean-Field Microkinetic Rate Predictions. Langmuir 2020, 36, 465474,  DOI: 10.1021/acs.langmuir.9b03563
  10. 10
    Mhadeshwar, A. B.; Kitchin, J. R.; Barteau, M. A.; Vlachos, D. G. The Role of Adsorbate–Adsorbate Interactions in the Rate Controlling Step and the Most Abundant Reaction Intermediate of NH3 Decomposition on Ru. Catal. Lett. 2004, 96, 1322,  DOI: 10.1023/B:CATL.0000029523.22277.e1
  11. 11
    Miller, S. D.; Pushkarev, V. V.; Gellman, A. J.; Kitchin, J. R. Simulating Temperature Programmed Desorption of Oxygen on Pt(111) Using DFT Derived Coverage Dependent Desorption Barriers. Top. Catal. 2013, 57, 106117
  12. 12
    Getman, R. B.; Schneider, W. F. DFT-Based Coverage-Dependent Model of Pt-Catalyzed NO-Oxidation. ChemCatChem 2010, 2, 14501460,  DOI: 10.1002/cctc.201000146
  13. 13
    Getman, R. B.; Xu, Y.; Schneider, W. F. Thermodynamics of Environment-Dependent Oxygen Chemisorption on Pt(111). J. Phys. Chem. C 2008, 112, 95599572,  DOI: 10.1021/jp800905a
  14. 14
    Bray, J. M.; Smith, J. L.; Schneider, W. F. Coverage-Dependent Adsorption at a Low Symmetry Surface: DFT and Statistical Analysis of Oxygen Chemistry on Kinked Pt(321). Top. Catal. 2014, 57, 89105,  DOI: 10.1007/s11244-013-0165-4
  15. 15
    Sutton, J. E.; Vlachos, D. G. Building Large Microkinetic Models with First-Principles’ Accuracy at Reduced Computational Cost. Chem. Eng. Sci. 2015, 121, 190199,  DOI: 10.1016/j.ces.2014.09.011
  16. 16
    Sabbe, M. K.; Canduela-Rodriguez, G.; Joly, J.-F.; Reyniers, M.-F.; Marin, G. B. Ab Initio Coverage-Dependent Microkinetic Modeling of Benzene Hydrogenation on Pd(111). Catal. Sci. Technol. 2017, 7, 52675283,  DOI: 10.1039/C7CY00962C
  17. 17
    Liu, H.; Liu, J.; Yang, B. Modeling the Effect of Surface CO Coverage on the Electrocatalytic Reduction of CO2 to CO on Pd Surfaces. Phys. Chem. Chem. Phys. 2019, 21, 98769882,  DOI: 10.1039/C8CP07427E
  18. 18
    Wu, P.; Zeffron, J.; Xu, D.; Yang, B. First-Principles-Based Microkinetic Simulations of CO2 Hydrogenation to Methanol over Intermetallic GaPd2: Method Development to Include Complex Interactions between Surface Adsorbates. J. Phys. Chem. C 2020, 124, 1597715987,  DOI: 10.1021/acs.jpcc.0c03975
  19. 19
    Yao, Z.; Zhao, J.; Bunting, R. J.; Zhao, C.; Hu, P.; Wang, J. Quantitative Insights into the Reaction Mechanism for the Direct Synthesis of H2O2 over Transition Metals: Coverage-Dependent Microkinetic Modeling. ACS Catal. 2021, 11, 12021221,  DOI: 10.1021/acscatal.0c04125
  20. 20
    Yao, Z.; Guo, C.; Mao, Y.; Hu, P. Quantitative Determination of C-C Coupling Mechanisms and Detailed Analyses on the Activity and Selectivity for Fischer-Tropsch Synthesis on Co(0001): Microkinetic Modeling with Coverage Effects. ACS Catal. 2019, 9, 59575973,  DOI: 10.1021/acscatal.9b01150
  21. 21
    Ding, Y.; Xu, Y.; Song, Y.; Guo, C.; Hu, P. Quantitative Studies of the Coverage Effects on Microkinetic Simulations for NO Oxidation on Pt(111). J. Phys. Chem. C 2019, 123, 2759427602,  DOI: 10.1021/acs.jpcc.9b08208
  22. 22
    Prats, H.; Illas, F.; Sayós, R. General Concepts, Assumptions, Drawbacks, and Misuses in Kinetic Monte Carlo and Microkinetic Modeling Simulations Applied to Computational Heterogeneous Catalysis. Int. J. Quantum Chem. 2018, 118, 25518
  23. 23
    Jørgenson, M.; Grönbeck, H. Selective Acetylene Hydrogenation over Single-Atom Alloy Nanoparticles by Kinetic Monte Carlo. J. Am. Chem. Soc. 2019, 141, 85418549,  DOI: 10.1021/jacs.9b02132
  24. 24
    Nagasaka, M.; Kondoh, H.; Nakai, I.; Ohta, T. CO Oxidation Reaction on Pt(111) Studied by the Dynamic Monte Carlo Method Including Lateral Interactions of Adsorbates. J. Chem. Phys. 2007, 126, 044704  DOI: 10.1063/1.2424705
  25. 25
    Wu, C.; Schmidt, D. J.; Wolverton, C.; Schneider, W. F. Accurate Coverage-Dependence Incorporated into First-Principles Kinetic Models: Catalytic NO oxidation on Pt (111). J. Catal. 2012, 286, 8894,  DOI: 10.1016/j.jcat.2011.10.020
  26. 26
    Yang, L.; Krim, A.; Muckerman, J. T. Density Functional Kinetic Monte Carlo Simulation of Water-Gas Shift Reaction on Cu/ZnO. J. Phys. Chem. C 2013, 117, 34143425,  DOI: 10.1021/jp3114286
  27. 27
    Smith, J. S.; Isayev, O.; Roitberg, A. E. ANI-1: An Extensible Neural Network Potential with DFT Accuracy at Force Field Computational Cost. Chem. Sci. 2017, 7, 31923203
  28. 28
    Meyer, J.; Bukas, V. J.; Mitra, S.; Reuter, K. Fingerprints of Energy Dissipation for Exothermic Surface Chemical Reactions: O2 on Pd(100). J. Chem. Phys. 2015, 143, 21312136,  DOI: 10.1063/1.4926989
  29. 29
    Boes, J. R.; Kitchin, J. R. Neural Network Predictions of Oxygen Interactions on a Dynamic Pd Surface. Mol. Simul. 2017, 43, 346354,  DOI: 10.1080/08927022.2016.1274984
  30. 30
    Boes, J. R.; Groenenboom, M. C.; Keith, J. A.; Kitchin, J. R. Neural Network and ReaxFF Comparison for Au Properties. Int. J. Quantum Chem. 2016, 116, 979987,  DOI: 10.1002/qua.25115
  31. 31
    Elstner, M.; Seifert, G. Density Functional Tight Binding. Philos. Trans. R. Soc., A 2014, 372, 20120483,  DOI: 10.1098/rsta.2012.0483
  32. 32
    van Duin, A. C. T.; Dasgupta, S.; Lorant, F.; Goddard, W. A. ReaxFF: A Reactive Force Field for Hydrocarbons. J. Phys. Chem. A 2001, 105, 93969409,  DOI: 10.1021/jp004368u
  33. 33
    Behler, J.; Parrinello, M. Generalized Neural-Network Representation of High-Dimensional Potential-Energy Surfaces. Phys. Rev. Lett. 2007, 98, 146401146404,  DOI: 10.1103/PhysRevLett.98.146401
  34. 34
    Bartók, A. P.; Csanyi, G. Gaussian Approximation Potentials: A Brief Tutorial Introduction. Int. J. Quantum Chem. 2015, 115, 10511057,  DOI: 10.1002/qua.24927
  35. 35
    Leshno, M.; Lin, V. Y.; Pinkus, A.; Schocken, S. Multilayer Feedforward Networks with a Nonpolynomial Activation Function Can Approximate Any Function. Neural Networks 1993, 6, 861867,  DOI: 10.1016/S0893-6080(05)80131-5
  36. 36
    Thomsen, J. U.; Meyer, B. Pattern Recognition of the 1H NMR Spectra of Sugar Alditols Using a Neural Network. J. Magn. Reson. 1989, 84, 212217
  37. 37
    Curry, B.; Rumelhart, D. E. MSnet: A Neural Network which Classifies Mass Spectra. Tetrahedron Comput. Methodol. 1990, 3, 213237,  DOI: 10.1016/0898-5529(90)90053-B
  38. 38
    Klopman, G. Artificial Intelligence Approach to Structure-Activity Studies. Computer Automated Structure Evaluation of Biological Activity of Organic Molecules. J. Am. Chem. Soc. 1984, 106, 73157321,  DOI: 10.1021/ja00336a004
  39. 39
    Hopfinger, A. J.; Burke, B. J.; Dunn, W. J., III A Generalized Formalism of Three-Dimensional Quantitative Structure-Property Relationship Analysis for Flexible Molecules Using Tensor Representation. J. Med. Chem. 1994, 37, 37683774,  DOI: 10.1021/jm00048a013
  40. 40
    Gakh, A. A.; Gakh, E. G.; Sumpter, B. G.; Noid, D. W. Neural Network-Graph Theory Approach to the Prediction of the Physical Properties of Organic Compounds. J. Chem. Inf. Comput. Sci. 1994, 34, 832839,  DOI: 10.1021/ci00020a017
  41. 41
    Sumpter, B. G.; Noid, D. W. Potential Energy Surfaces for Macromolecules. A Neural Network Technique. Chem. Phys. Lett. 1992, 192, 7986,  DOI: 10.1016/0009-2614(92)85498-Y
  42. 42
    Blank, T. B.; Brown, S. D.; Calhoun, A. W.; Doren, D. J. Neural Network Models of Potential Energy Surfaces. J. Chem. Phys. 1995, 103, 41294137,  DOI: 10.1063/1.469597
  43. 43
    Brown, D. F. R.; Gibbs, M. N.; Clary, D. C. Combining Ab Initio Computations, Neural Networks, and Diffusion Monte Carlo: An Efficient Method to Treat Weakly Bound Molecules. J. Chem. Phys. 1996, 105, 75977604,  DOI: 10.1063/1.472596
  44. 44
    Behler, J. Representing Potential Energy Surfaces by High-Dimensional Neural Network Potentials. Condens. Matter 2014, 26, 183001,  DOI: 10.1088/0953-8984/26/18/183001
  45. 45
    Behler, J. Atom-Centered Symmetry Functions for Constructing High-Dimensional Neural Network Potentials. J. Chem. Phys. 2011, 134, 074106  DOI: 10.1063/1.3553717
  46. 46
    Jose, K. V. J.; Artrith, A.; Behler, J. Construction of High-Dimensional Neural Network Potentials Using Environment-Dependent Atom Pairs. J. Chem. Phys. 2012, 136, 194111,  DOI: 10.1063/1.4712397
  47. 47
    Shakouri, K.; Behler, J.; Meyer, J.; Kroes, G.-J. Accurate Neural Network Description of Surface Phonons in Reactive Gas–Surface Dynamics: N2 + Ru(0001). J. Phys. Chem. Lett. 2017, 8, 21312136,  DOI: 10.1021/acs.jpclett.7b00784
  48. 48
    Gao, T.; Kitchin, J. R. Modeling Palladium Surfaces with Density Functional Theory, Neural Networks and Molecular Dynamics. Catal. Today 2018, 312, 132140,  DOI: 10.1016/j.cattod.2018.03.045
  49. 49
    Wang, C.; Tharval, A.; Kitchin, J. R. A Density Functional Theory Parameterised Neural Network Model of Zirconia. Mol. Simul. 2018, 44, 623630,  DOI: 10.1080/08927022.2017.1420185
  50. 50
    Eckhoff, M.; Behler, J. From Molecular Fragments to the Bulk: Development of a Neural Network Potential for MOF-5. J. Chem. Theory Comput. 2019, 15, 37933809,  DOI: 10.1021/acs.jctc.8b01288
  51. 51
    Schran, C.; Uhl, F.; Behler, J.; Marx, D. High-Dimensional Neural Network Potentials for Solvation: The Case of Protonated Water Clusters in Helium. J. Chem. Phys. 2018, 148, 102310,  DOI: 10.1063/1.4996819
  52. 52
    Lahey, S.-H. J.; Rowley, C. N. Simulating Protein–Ligand Binding with Neural Network Potentials. Chem. Sci. 2020, 11, 23622368,  DOI: 10.1039/C9SC06017K
  53. 53
    Sun, G.; Sautet, P. Metastable Structures in Cluster Catalysis from First-Principles: Structural Ensemble in Reaction Conditions and Metastability Triggered Reactivity. J. Am. Chem. Soc. 2018, 140, 28122820,  DOI: 10.1021/jacs.7b11239
  54. 54
    Zuo, Y.; Chen, C.; Li, X.; Deng, Z.; Chen, Y.; Behler, J.; Csányi, G.; Shapeev, A. V.; Thompson, A. P.; Wood, A. M.; Ong, S. P. A Performance and Cost Assessment of Machine Learning Interatomic Potentials. J. Phys. Chem. A 2020, 124, 731745,  DOI: 10.1021/acs.jpca.9b08723
  55. 55
    Bartók, A. P.; Gillan, M. J.; Manby, F. R.; Csányi, G. Machine-Learning Approach for One- and Two-Body Corrections to Density Functional Theory: Applications to Molecular and Condensed Water. Phys. Rev. B 2013, 88, 054104,  DOI: 10.1103/PhysRevB.87.184115
  56. 56
    Hanse, K.; Montavon, G.; Biegler, F.; Fazli, S.; Rupp, M.; Scheffler, M.; von Lilienfeld, O. A.; Tkatchenko, A.; Müller, K.-R. Assessment and Validation of Machine Learning Methods for Predicting Molecular Atomization Energies. J. Chem. Theory Comput. 2013, 9, 34043419,  DOI: 10.1021/ct400195d
  57. 57
    Montavon, G.; Rupp, M.; Gobre, V.; Vazquez-Mayagoitia, A.; Hansen, K.; Tkatchenko, A.; Müller, K.-R.; von Lilienfeld, O. A. Machine Learning of Molecular Electronic Properties in Chemical Compound Space. New J. Phys. 2013, 15, 095003  DOI: 10.1088/1367-2630/15/9/095003
  58. 58
    Rupp, M.; Ramakrishnan, R.; von Lilienfeld, O. A. Machine Learning for Quantum Mechanical Properties of Atoms in Molecules. J. Phys. Chem. Lett. 2015, 6, 33093313,  DOI: 10.1021/acs.jpclett.5b01456
  59. 59
    Weststrate, C. J.; van de Loosdrecht, J.; Niemantsverdriet, J. W. Spectroscopic Insights into Cobalt-Catalyzed Fischer-Tropsch Synthesis: A Review of the Carbon Monoxide Interaction with Single Crystalline Surfaces of Cobalt. J. Catal. 2016, 342, 116,  DOI: 10.1016/j.jcat.2016.07.010
  60. 60
    Werbos, P. J. Generalization of Backpropagation with Application to a Recurrent Gas Market Model. Neural Networks 1988, 1, 339356,  DOI: 10.1016/0893-6080(88)90007-X
  61. 61
    Huang, Y.; Kang, J.; Goddard, W. A.; Wang, L.-W. Density Functional Theory Based Neural Network Force Fields from Energy Decompositions. Phys. Rev. B 2019, 99, 064103  DOI: 10.1103/PhysRevB.99.064103
  62. 62
    Gastegger, M.; Schwiedrzik, L.; Bittermann, M.; Berzsenyi, F.; Marquetand, P. wACSF - Weighted Atom-Centered Symmetry Functions as Descriptors in Machine Learning Potentials. J. Chem. Phys. 2018, 148, 241709,  DOI: 10.1063/1.5019667
  63. 63
    Artrith, N.; Urban, A.; Ceder, G. Efficient and Accurate Machine-Learning Interpolation of Atomic Energies in Compositions with Many Species. Phys. Rev. B 2017, 96, 014112  DOI: 10.1103/PhysRevB.96.014112
  64. 64
    Seko, A.; Hayashi, H.; Nakayama, K.; Takahashi, A.; Tanaka, I. Representation of Compounds for Machine-Learning Prediction of Physical Properties. Phys. Rev. B 2017, 95, 144110,  DOI: 10.1103/PhysRevB.95.144110
  65. 65
    Huang, B.; von Lilienfeld, O. A. Communication: Understanding Molecular Representations in Machine Learning: The Role of Uniqueness and Target Similarity. J. Chem. Phys. 2016, 145, 161102,  DOI: 10.1063/1.4964627
  66. 66
    Handley, C. M.; Popelier, P. L. A. Potential Energy Surfaces Fitted by Artificial Neural Networks. J. Phys. Chem. A 2010, 114, 33713383,  DOI: 10.1021/jp9105585
  67. 67
    Ramsvik, T.; Borg, A.; Kildemo, M.; Raaen, S.; Matsuura, A.; Jaworowski, A. J.; Worren, T.; Leandersson, M. Molecular Vibrations in Core-Ionised CO Adsorbed on Co(0001) and Rh(100). Surf. Sci. 2001, 492, 152160,  DOI: 10.1016/S0039-6028(01)01446-7
  68. 68
    Lahtinen, J.; Vaari, J.; Kauraala, K. Adsorption and Structure Dependent Desorption of CO on Co(0001). Surf. Sci. 1998, 418, 502510,  DOI: 10.1016/S0039-6028(98)00711-0
  69. 69
    Joos, L.; Filot, I. A. W.; Cottenier, S.; Hensen, E. J. M.; Waroquier, M.; Van Speybroeck, V.; van Santen, R. A. Reactivity of CO on Carbon-Covered Cobalt Surfaces in Fischer-Tropsch Synthesis. J. Phys. Chem. C 2014, 118, 53175327,  DOI: 10.1021/jp4109706
  70. 70
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  71. 71
    Hammer, B.; Hansen, L. B.; Nørskov, J. K. Improved Adsorption Energetics within Density-Functional Theory using Revised Perdew-Burke-Ernzerhof Functionals. Phys. Rev. B 1999, 59, 74137421,  DOI: 10.1103/PhysRevB.59.7413
  72. 72
    Mattsson, A. E.; Armiento, R.; Paier, J.; Kresse, G.; Wills, J. M.; Mattsson, T. R. The AM05 Density Functional Applied to Solids. J. Chem. Phys. 2008, 128, 084714,  DOI: 10.1063/1.2835596
  73. 73
    Perdew, J. P.; Ruzsinszky, A.; Csonka, G. I.; Vydrov, O. A.; Scuseria, G. E.; Constantin, L. A.; Zhou, X.; Burke, K. Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces. Phys. Rev. Lett. 2008, 100, 136406,  DOI: 10.1103/PhysRevLett.100.136406
  74. 74
    Zhang, Y.; Yang, W. Comment on “Generalized Gradient Approximation Made Simple”. Phys. Rev. Lett. 1998, 80, 890,  DOI: 10.1103/PhysRevLett.80.890
  75. 75
    Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 14561465,  DOI: 10.1002/jcc.21759
  76. 76
    Grimme, S.; Hansen, A.; Brandenburg, J. G.; Bannwarth, C. Dispersion-Corrected Mean-Field Electronic Structure Methods. Chem. Rev. 2016, 116, 51055154,  DOI: 10.1021/acs.chemrev.5b00533

Cited By

Click to copy section linkSection link copied!

This article is cited by 8 publications.

  1. Andrey V. Tarasov, Sabine Wrabetz, Jutta Kröhnert, Frank Rosowski, Robert Schlögl, Annette Trunschke. Energy Maps of Complex Catalyst Surfaces. ACS Catalysis 2024, 14 (16) , 12581-12591. https://doi.org/10.1021/acscatal.4c01308
  2. Bart Klumpers, Emiel J. M. Hensen, Ivo A. W. Filot. Transferable, Living Data Sets for Predicting Global Minimum Energy Nanocluster Geometries. Journal of Chemical Theory and Computation 2024, 20 (15) , 6801-6812. https://doi.org/10.1021/acs.jctc.4c00572
  3. Shuqiao Wang, Alyssa J.R. Hensley. Rotational Symmetry Effects on Multibody Lateral Interactions between Co-Adsorbates at Heterogeneous Interfaces. ACS Physical Chemistry Au 2024, 4 (4) , 328-335. https://doi.org/10.1021/acsphyschemau.4c00019
  4. Konstantijn Tom Rommens, Mark Saeys. Molecular Views on Fischer–Tropsch Synthesis. Chemical Reviews 2023, 123 (9) , 5798-5858. https://doi.org/10.1021/acs.chemrev.2c00508
  5. Daniel Schwalbe-Koda, Nitish Govindarajan, Joel B. Varley. Comprehensive sampling of coverage effects in catalysis by leveraging generalization in neural network models. Digital Discovery 2025, 4 (1) , 234-251. https://doi.org/10.1039/D4DD00328D
  6. Bart Klumpers, Tim Luijten, Stijn Gerritse, Emiel Hensen, Ivo Filot. Direct coupling of microkinetic and reactor models using neural networks. Chemical Engineering Journal 2023, 475 , 145538. https://doi.org/10.1016/j.cej.2023.145538
  7. Pieter Janssens, Jeroen Poissonnier, Anoop Chakkingal, René Bos, Joris W. Thybaut. Recent advances in the use of Steady-State Isotopic Transient Kinetic Analysis data in (micro)kinetic modeling for catalyst and process design. Catalysis Communications 2023, 179 , 106688. https://doi.org/10.1016/j.catcom.2023.106688
  8. Bart Klumpers, Tim Luijten, Stijn Gerritse, Emiel Hensen, Ivo Filot. Direct Coupling of Microkinetic and Reactor Models Using Neural Networks. 2023https://doi.org/10.2139/ssrn.4510408
  9. Li Wang, Baojun Wang, Maohong Fan, Lixia Ling, Riguang Zhang. Unraveling the structure and composition sensitivity of transition metal phosphide toward catalytic performance of C2H2 semi-hydrogenation. Journal of Catalysis 2022, 416 , 112-128. https://doi.org/10.1016/j.jcat.2022.10.016

The Journal of Physical Chemistry C

Cite this: J. Phys. Chem. C 2022, 126, 12, 5529–5540
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpcc.1c10401
Published March 17, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

1750

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Illustration of radial (a) and angular (b) coordination descriptors. Radial distributions are defined with respect to the atom marked as (i), angular distributions relative to the axis between atoms (i,j). Different descriptors are shown using various colors. The graphs show the signal values as a function of the descriptor parameters. (a) Left: cutoff function gc for various cutoff radii. Right: radial coordinate descriptors Ri for various decay rates. (b) Angular coordinate descriptors Ai for various angular decay parameters.

    Figure 2

    Figure 2. Schematic depiction of the hybrid network architecture. Separate geometric and electronic depictions of a molecular geometry are obtained, from which the atomic networks reconstruct the total energy. This compares to classical approaches in which only the geometric representation is used. Geometric and electronic information are communicated to the atomic networks using a set of basis functions, each represented as a single input node. One atomic network is constructed per chemically unique element (here shown in green, yellow, and red).

    Figure 3

    Figure 3. Different stages of dataset construction: I. Translational states along a geometrically bounding potential surface. II. At each translational site, different rotations are obtained by rotating the molecule along two internal coordinates. III. Each rotation is expanded into a set of vibrational states. IV. Combinations of these states are used to construct a set of adlayers. At this step, translations are extended over a set of support grids to alleviate symmetry constraints. Note the sparser sample density over these extended surfaces.

    Figure 4

    Figure 4. Error distributions for a selection of CO/Co(001) networks. Model (II) is a purely structural network with the same architecture as the single-transform hybrid ANN detailed in the main text. Model (III) is a structural network with a reoptimized architecture with dimensionality 20 and 2 transforms. Model (IV) uses the reoptimized architecture but with the extended dataset (II) as the training dataset. Model (V) uses the reoptimized architecture along with an extended descriptor set (trained using dataset (I), parameters are provided in Figure S2). Model (VI) combines the extended dataset (II) with the extended descriptor set of model (V). Additional details are provided in Table S11.

    Figure 5

    Figure 5. ANN potentials for the CO adsorption energy in the presence of lateral interactions over the left (Co(001)) and middle (Co(211)). Profiles at 300 and 500 K are derived from temperature-corrected adlayer energies obtained through MC. Hexagonal markers represent reference DFT minimum energy points. Right: onset coverage at which temperature-dependent lateral interaction potentials deviate from the 0 K potential by >1 kJ/mol.

    Figure 6

    Figure 6. Minimum energy structures for CO/Co(211) predicted with the ANN. From left to right: 0.2, 0.4, 0.7, and 0.9 ML.

    Figure 7

    Figure 7. Left: equilibrium CO coverage on the Co(001) surface, determined from a microkinetic model (MKM) composed of a single-reversible adsorption step with coverage-dependent adsorption energies. (Details are provided in the Supporting Information). The HK model omits lateral interactions. Markers show experimental data published by Weststrate et al. (59) Right: effect of the temperature correction applied to the lateral interaction model at 1 bar.

    Figure 8

    Figure 8. Difference in the lateral interaction potentials for Co(001) predicted for different XC functionals at T = 0, 300, and 500 K. Average adsorption energies are expressed relative to the adsorption energy of CO on the empty surface.

    Figure 9

    Figure 9. Left: influence of different dispersion corrections on the revPBE lateral interaction potential for CO/Co(001). Integral adsorption energies have been expressed relative to the adsorption on the empty surface. Right: comparison of Grimme dispersion effects on the adsorption equilibrium for CO/Co(001) at 1 bar.

  • References


    This article references 76 other publications.

    1. 1
      Chen, W.; Pestman, R.; Zijlstra, B.; Filot, I. A. W.; Hensen, E. J. M. Mechanism of Cobalt-Catalyzed CO Hydrogenation: 1 Methanation. ACS Catal. 2017, 7, 80508060,  DOI: 10.1021/acscatal.7b02757
    2. 2
      Huš, M.; Grilc, M.; Pavlišič, A.; Likozar, B.; Hellman, A. Multiscale Modelling from Quantum Level to Reactor Scale: An Example of Ethylene Epoxidation on Silver Catalysts. Catal. Today 2019, 338, 128140,  DOI: 10.1016/j.cattod.2019.05.022
    3. 3
      Prats, H.; Posada-Pérez, S.; Rodriguez, J. A.; Sayós, R.; Illas, F. Kinetic Monte Carlo Simulations Unveil Synergic Effects at Work on Bifunctional Catalysts. ACS Catal. 2019, 9, 91179126,  DOI: 10.1021/acscatal.9b02813
    4. 4
      Zijlstra, B.; Broos, R. J. P.; Chen, W.; Oosterbeek, H.; Filot, I. A. W.; Hensen, E. J. M. Coverage Effects in CO Dissociation on Metallic Cobalt Nanoparticles. ACS Catal. 2019, 9, 73657372,  DOI: 10.1021/acscatal.9b01967
    5. 5
      Zijlstra, B.; Broos, R. J. P.; Chen, W.; Bezemer, G. L.; Filot, I. A. W.; Hensen, E. J. M. The Vital Role of Step-Edge Sites for Both CO Activation and Chain Growth on Cobalt Fischer–Tropsch Catalysts Revealed through First-Principles-Based Microkinetic Modeling Including Lateral Interactions. ACS Catal. 2020, 10, 93769400,  DOI: 10.1021/acscatal.0c02420
    6. 6
      Grabow, L. C.; Hvolbæk, B.; Nørskov, J. K. Understanding Trends in Catalytic Activity: The Effect of Adsorbate–Adsorbate Interactions for CO Oxidation Over Transition Metals. Top. Catal. 2010, 53, 298310,  DOI: 10.1007/s11244-010-9455-2
    7. 7
      Lausche, A. C.; Medford, A. J.; Khan, T. S.; Xu, Y.; Bligaard, T.; Abild-Pedersen, F.; Nørskov, J. K.; Studt, F. On the Effect of Coverage-Dependent Adsorbate-Adsorbate Interactions for CO Methanation on Transition Metal Surfaces. J. Catal. 2013, 307, 275282,  DOI: 10.1016/j.jcat.2013.08.002
    8. 8
      Grabow, L. C.; Gokhale, A. A.; Evans, S. T.; Dumesic, J. A.; Mavrikakis, M. Mechanism of the Water Gas Shift Reaction on Pt: First Principles, Experiments, and Microkinetic Modeling. J. Phys. Chem. C 2008, 112, 46084617,  DOI: 10.1021/jp7099702
    9. 9
      Bajpai, A.; Frey, K.; Schneider, W. F. Comparison of Coverage-Dependent Binding Energy Models for Mean-Field Microkinetic Rate Predictions. Langmuir 2020, 36, 465474,  DOI: 10.1021/acs.langmuir.9b03563
    10. 10
      Mhadeshwar, A. B.; Kitchin, J. R.; Barteau, M. A.; Vlachos, D. G. The Role of Adsorbate–Adsorbate Interactions in the Rate Controlling Step and the Most Abundant Reaction Intermediate of NH3 Decomposition on Ru. Catal. Lett. 2004, 96, 1322,  DOI: 10.1023/B:CATL.0000029523.22277.e1
    11. 11
      Miller, S. D.; Pushkarev, V. V.; Gellman, A. J.; Kitchin, J. R. Simulating Temperature Programmed Desorption of Oxygen on Pt(111) Using DFT Derived Coverage Dependent Desorption Barriers. Top. Catal. 2013, 57, 106117
    12. 12
      Getman, R. B.; Schneider, W. F. DFT-Based Coverage-Dependent Model of Pt-Catalyzed NO-Oxidation. ChemCatChem 2010, 2, 14501460,  DOI: 10.1002/cctc.201000146
    13. 13
      Getman, R. B.; Xu, Y.; Schneider, W. F. Thermodynamics of Environment-Dependent Oxygen Chemisorption on Pt(111). J. Phys. Chem. C 2008, 112, 95599572,  DOI: 10.1021/jp800905a
    14. 14
      Bray, J. M.; Smith, J. L.; Schneider, W. F. Coverage-Dependent Adsorption at a Low Symmetry Surface: DFT and Statistical Analysis of Oxygen Chemistry on Kinked Pt(321). Top. Catal. 2014, 57, 89105,  DOI: 10.1007/s11244-013-0165-4
    15. 15
      Sutton, J. E.; Vlachos, D. G. Building Large Microkinetic Models with First-Principles’ Accuracy at Reduced Computational Cost. Chem. Eng. Sci. 2015, 121, 190199,  DOI: 10.1016/j.ces.2014.09.011
    16. 16
      Sabbe, M. K.; Canduela-Rodriguez, G.; Joly, J.-F.; Reyniers, M.-F.; Marin, G. B. Ab Initio Coverage-Dependent Microkinetic Modeling of Benzene Hydrogenation on Pd(111). Catal. Sci. Technol. 2017, 7, 52675283,  DOI: 10.1039/C7CY00962C
    17. 17
      Liu, H.; Liu, J.; Yang, B. Modeling the Effect of Surface CO Coverage on the Electrocatalytic Reduction of CO2 to CO on Pd Surfaces. Phys. Chem. Chem. Phys. 2019, 21, 98769882,  DOI: 10.1039/C8CP07427E
    18. 18
      Wu, P.; Zeffron, J.; Xu, D.; Yang, B. First-Principles-Based Microkinetic Simulations of CO2 Hydrogenation to Methanol over Intermetallic GaPd2: Method Development to Include Complex Interactions between Surface Adsorbates. J. Phys. Chem. C 2020, 124, 1597715987,  DOI: 10.1021/acs.jpcc.0c03975
    19. 19
      Yao, Z.; Zhao, J.; Bunting, R. J.; Zhao, C.; Hu, P.; Wang, J. Quantitative Insights into the Reaction Mechanism for the Direct Synthesis of H2O2 over Transition Metals: Coverage-Dependent Microkinetic Modeling. ACS Catal. 2021, 11, 12021221,  DOI: 10.1021/acscatal.0c04125
    20. 20
      Yao, Z.; Guo, C.; Mao, Y.; Hu, P. Quantitative Determination of C-C Coupling Mechanisms and Detailed Analyses on the Activity and Selectivity for Fischer-Tropsch Synthesis on Co(0001): Microkinetic Modeling with Coverage Effects. ACS Catal. 2019, 9, 59575973,  DOI: 10.1021/acscatal.9b01150
    21. 21
      Ding, Y.; Xu, Y.; Song, Y.; Guo, C.; Hu, P. Quantitative Studies of the Coverage Effects on Microkinetic Simulations for NO Oxidation on Pt(111). J. Phys. Chem. C 2019, 123, 2759427602,  DOI: 10.1021/acs.jpcc.9b08208
    22. 22
      Prats, H.; Illas, F.; Sayós, R. General Concepts, Assumptions, Drawbacks, and Misuses in Kinetic Monte Carlo and Microkinetic Modeling Simulations Applied to Computational Heterogeneous Catalysis. Int. J. Quantum Chem. 2018, 118, 25518
    23. 23
      Jørgenson, M.; Grönbeck, H. Selective Acetylene Hydrogenation over Single-Atom Alloy Nanoparticles by Kinetic Monte Carlo. J. Am. Chem. Soc. 2019, 141, 85418549,  DOI: 10.1021/jacs.9b02132
    24. 24
      Nagasaka, M.; Kondoh, H.; Nakai, I.; Ohta, T. CO Oxidation Reaction on Pt(111) Studied by the Dynamic Monte Carlo Method Including Lateral Interactions of Adsorbates. J. Chem. Phys. 2007, 126, 044704  DOI: 10.1063/1.2424705
    25. 25
      Wu, C.; Schmidt, D. J.; Wolverton, C.; Schneider, W. F. Accurate Coverage-Dependence Incorporated into First-Principles Kinetic Models: Catalytic NO oxidation on Pt (111). J. Catal. 2012, 286, 8894,  DOI: 10.1016/j.jcat.2011.10.020
    26. 26
      Yang, L.; Krim, A.; Muckerman, J. T. Density Functional Kinetic Monte Carlo Simulation of Water-Gas Shift Reaction on Cu/ZnO. J. Phys. Chem. C 2013, 117, 34143425,  DOI: 10.1021/jp3114286
    27. 27
      Smith, J. S.; Isayev, O.; Roitberg, A. E. ANI-1: An Extensible Neural Network Potential with DFT Accuracy at Force Field Computational Cost. Chem. Sci. 2017, 7, 31923203
    28. 28
      Meyer, J.; Bukas, V. J.; Mitra, S.; Reuter, K. Fingerprints of Energy Dissipation for Exothermic Surface Chemical Reactions: O2 on Pd(100). J. Chem. Phys. 2015, 143, 21312136,  DOI: 10.1063/1.4926989
    29. 29
      Boes, J. R.; Kitchin, J. R. Neural Network Predictions of Oxygen Interactions on a Dynamic Pd Surface. Mol. Simul. 2017, 43, 346354,  DOI: 10.1080/08927022.2016.1274984
    30. 30
      Boes, J. R.; Groenenboom, M. C.; Keith, J. A.; Kitchin, J. R. Neural Network and ReaxFF Comparison for Au Properties. Int. J. Quantum Chem. 2016, 116, 979987,  DOI: 10.1002/qua.25115
    31. 31
      Elstner, M.; Seifert, G. Density Functional Tight Binding. Philos. Trans. R. Soc., A 2014, 372, 20120483,  DOI: 10.1098/rsta.2012.0483
    32. 32
      van Duin, A. C. T.; Dasgupta, S.; Lorant, F.; Goddard, W. A. ReaxFF: A Reactive Force Field for Hydrocarbons. J. Phys. Chem. A 2001, 105, 93969409,  DOI: 10.1021/jp004368u
    33. 33
      Behler, J.; Parrinello, M. Generalized Neural-Network Representation of High-Dimensional Potential-Energy Surfaces. Phys. Rev. Lett. 2007, 98, 146401146404,  DOI: 10.1103/PhysRevLett.98.146401
    34. 34
      Bartók, A. P.; Csanyi, G. Gaussian Approximation Potentials: A Brief Tutorial Introduction. Int. J. Quantum Chem. 2015, 115, 10511057,  DOI: 10.1002/qua.24927
    35. 35
      Leshno, M.; Lin, V. Y.; Pinkus, A.; Schocken, S. Multilayer Feedforward Networks with a Nonpolynomial Activation Function Can Approximate Any Function. Neural Networks 1993, 6, 861867,  DOI: 10.1016/S0893-6080(05)80131-5
    36. 36
      Thomsen, J. U.; Meyer, B. Pattern Recognition of the 1H NMR Spectra of Sugar Alditols Using a Neural Network. J. Magn. Reson. 1989, 84, 212217
    37. 37
      Curry, B.; Rumelhart, D. E. MSnet: A Neural Network which Classifies Mass Spectra. Tetrahedron Comput. Methodol. 1990, 3, 213237,  DOI: 10.1016/0898-5529(90)90053-B
    38. 38
      Klopman, G. Artificial Intelligence Approach to Structure-Activity Studies. Computer Automated Structure Evaluation of Biological Activity of Organic Molecules. J. Am. Chem. Soc. 1984, 106, 73157321,  DOI: 10.1021/ja00336a004
    39. 39
      Hopfinger, A. J.; Burke, B. J.; Dunn, W. J., III A Generalized Formalism of Three-Dimensional Quantitative Structure-Property Relationship Analysis for Flexible Molecules Using Tensor Representation. J. Med. Chem. 1994, 37, 37683774,  DOI: 10.1021/jm00048a013
    40. 40
      Gakh, A. A.; Gakh, E. G.; Sumpter, B. G.; Noid, D. W. Neural Network-Graph Theory Approach to the Prediction of the Physical Properties of Organic Compounds. J. Chem. Inf. Comput. Sci. 1994, 34, 832839,  DOI: 10.1021/ci00020a017
    41. 41
      Sumpter, B. G.; Noid, D. W. Potential Energy Surfaces for Macromolecules. A Neural Network Technique. Chem. Phys. Lett. 1992, 192, 7986,  DOI: 10.1016/0009-2614(92)85498-Y
    42. 42
      Blank, T. B.; Brown, S. D.; Calhoun, A. W.; Doren, D. J. Neural Network Models of Potential Energy Surfaces. J. Chem. Phys. 1995, 103, 41294137,  DOI: 10.1063/1.469597
    43. 43
      Brown, D. F. R.; Gibbs, M. N.; Clary, D. C. Combining Ab Initio Computations, Neural Networks, and Diffusion Monte Carlo: An Efficient Method to Treat Weakly Bound Molecules. J. Chem. Phys. 1996, 105, 75977604,  DOI: 10.1063/1.472596
    44. 44
      Behler, J. Representing Potential Energy Surfaces by High-Dimensional Neural Network Potentials. Condens. Matter 2014, 26, 183001,  DOI: 10.1088/0953-8984/26/18/183001
    45. 45
      Behler, J. Atom-Centered Symmetry Functions for Constructing High-Dimensional Neural Network Potentials. J. Chem. Phys. 2011, 134, 074106  DOI: 10.1063/1.3553717
    46. 46
      Jose, K. V. J.; Artrith, A.; Behler, J. Construction of High-Dimensional Neural Network Potentials Using Environment-Dependent Atom Pairs. J. Chem. Phys. 2012, 136, 194111,  DOI: 10.1063/1.4712397
    47. 47
      Shakouri, K.; Behler, J.; Meyer, J.; Kroes, G.-J. Accurate Neural Network Description of Surface Phonons in Reactive Gas–Surface Dynamics: N2 + Ru(0001). J. Phys. Chem. Lett. 2017, 8, 21312136,  DOI: 10.1021/acs.jpclett.7b00784
    48. 48
      Gao, T.; Kitchin, J. R. Modeling Palladium Surfaces with Density Functional Theory, Neural Networks and Molecular Dynamics. Catal. Today 2018, 312, 132140,  DOI: 10.1016/j.cattod.2018.03.045
    49. 49
      Wang, C.; Tharval, A.; Kitchin, J. R. A Density Functional Theory Parameterised Neural Network Model of Zirconia. Mol. Simul. 2018, 44, 623630,  DOI: 10.1080/08927022.2017.1420185
    50. 50
      Eckhoff, M.; Behler, J. From Molecular Fragments to the Bulk: Development of a Neural Network Potential for MOF-5. J. Chem. Theory Comput. 2019, 15, 37933809,  DOI: 10.1021/acs.jctc.8b01288
    51. 51
      Schran, C.; Uhl, F.; Behler, J.; Marx, D. High-Dimensional Neural Network Potentials for Solvation: The Case of Protonated Water Clusters in Helium. J. Chem. Phys. 2018, 148, 102310,  DOI: 10.1063/1.4996819
    52. 52
      Lahey, S.-H. J.; Rowley, C. N. Simulating Protein–Ligand Binding with Neural Network Potentials. Chem. Sci. 2020, 11, 23622368,  DOI: 10.1039/C9SC06017K
    53. 53
      Sun, G.; Sautet, P. Metastable Structures in Cluster Catalysis from First-Principles: Structural Ensemble in Reaction Conditions and Metastability Triggered Reactivity. J. Am. Chem. Soc. 2018, 140, 28122820,  DOI: 10.1021/jacs.7b11239
    54. 54
      Zuo, Y.; Chen, C.; Li, X.; Deng, Z.; Chen, Y.; Behler, J.; Csányi, G.; Shapeev, A. V.; Thompson, A. P.; Wood, A. M.; Ong, S. P. A Performance and Cost Assessment of Machine Learning Interatomic Potentials. J. Phys. Chem. A 2020, 124, 731745,  DOI: 10.1021/acs.jpca.9b08723
    55. 55
      Bartók, A. P.; Gillan, M. J.; Manby, F. R.; Csányi, G. Machine-Learning Approach for One- and Two-Body Corrections to Density Functional Theory: Applications to Molecular and Condensed Water. Phys. Rev. B 2013, 88, 054104,  DOI: 10.1103/PhysRevB.87.184115
    56. 56
      Hanse, K.; Montavon, G.; Biegler, F.; Fazli, S.; Rupp, M.; Scheffler, M.; von Lilienfeld, O. A.; Tkatchenko, A.; Müller, K.-R. Assessment and Validation of Machine Learning Methods for Predicting Molecular Atomization Energies. J. Chem. Theory Comput. 2013, 9, 34043419,  DOI: 10.1021/ct400195d
    57. 57
      Montavon, G.; Rupp, M.; Gobre, V.; Vazquez-Mayagoitia, A.; Hansen, K.; Tkatchenko, A.; Müller, K.-R.; von Lilienfeld, O. A. Machine Learning of Molecular Electronic Properties in Chemical Compound Space. New J. Phys. 2013, 15, 095003  DOI: 10.1088/1367-2630/15/9/095003
    58. 58
      Rupp, M.; Ramakrishnan, R.; von Lilienfeld, O. A. Machine Learning for Quantum Mechanical Properties of Atoms in Molecules. J. Phys. Chem. Lett. 2015, 6, 33093313,  DOI: 10.1021/acs.jpclett.5b01456
    59. 59
      Weststrate, C. J.; van de Loosdrecht, J.; Niemantsverdriet, J. W. Spectroscopic Insights into Cobalt-Catalyzed Fischer-Tropsch Synthesis: A Review of the Carbon Monoxide Interaction with Single Crystalline Surfaces of Cobalt. J. Catal. 2016, 342, 116,  DOI: 10.1016/j.jcat.2016.07.010
    60. 60
      Werbos, P. J. Generalization of Backpropagation with Application to a Recurrent Gas Market Model. Neural Networks 1988, 1, 339356,  DOI: 10.1016/0893-6080(88)90007-X
    61. 61
      Huang, Y.; Kang, J.; Goddard, W. A.; Wang, L.-W. Density Functional Theory Based Neural Network Force Fields from Energy Decompositions. Phys. Rev. B 2019, 99, 064103  DOI: 10.1103/PhysRevB.99.064103
    62. 62
      Gastegger, M.; Schwiedrzik, L.; Bittermann, M.; Berzsenyi, F.; Marquetand, P. wACSF - Weighted Atom-Centered Symmetry Functions as Descriptors in Machine Learning Potentials. J. Chem. Phys. 2018, 148, 241709,  DOI: 10.1063/1.5019667
    63. 63
      Artrith, N.; Urban, A.; Ceder, G. Efficient and Accurate Machine-Learning Interpolation of Atomic Energies in Compositions with Many Species. Phys. Rev. B 2017, 96, 014112  DOI: 10.1103/PhysRevB.96.014112
    64. 64
      Seko, A.; Hayashi, H.; Nakayama, K.; Takahashi, A.; Tanaka, I. Representation of Compounds for Machine-Learning Prediction of Physical Properties. Phys. Rev. B 2017, 95, 144110,  DOI: 10.1103/PhysRevB.95.144110
    65. 65
      Huang, B.; von Lilienfeld, O. A. Communication: Understanding Molecular Representations in Machine Learning: The Role of Uniqueness and Target Similarity. J. Chem. Phys. 2016, 145, 161102,  DOI: 10.1063/1.4964627
    66. 66
      Handley, C. M.; Popelier, P. L. A. Potential Energy Surfaces Fitted by Artificial Neural Networks. J. Phys. Chem. A 2010, 114, 33713383,  DOI: 10.1021/jp9105585
    67. 67
      Ramsvik, T.; Borg, A.; Kildemo, M.; Raaen, S.; Matsuura, A.; Jaworowski, A. J.; Worren, T.; Leandersson, M. Molecular Vibrations in Core-Ionised CO Adsorbed on Co(0001) and Rh(100). Surf. Sci. 2001, 492, 152160,  DOI: 10.1016/S0039-6028(01)01446-7
    68. 68
      Lahtinen, J.; Vaari, J.; Kauraala, K. Adsorption and Structure Dependent Desorption of CO on Co(0001). Surf. Sci. 1998, 418, 502510,  DOI: 10.1016/S0039-6028(98)00711-0
    69. 69
      Joos, L.; Filot, I. A. W.; Cottenier, S.; Hensen, E. J. M.; Waroquier, M.; Van Speybroeck, V.; van Santen, R. A. Reactivity of CO on Carbon-Covered Cobalt Surfaces in Fischer-Tropsch Synthesis. J. Phys. Chem. C 2014, 118, 53175327,  DOI: 10.1021/jp4109706
    70. 70
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    71. 71
      Hammer, B.; Hansen, L. B.; Nørskov, J. K. Improved Adsorption Energetics within Density-Functional Theory using Revised Perdew-Burke-Ernzerhof Functionals. Phys. Rev. B 1999, 59, 74137421,  DOI: 10.1103/PhysRevB.59.7413
    72. 72
      Mattsson, A. E.; Armiento, R.; Paier, J.; Kresse, G.; Wills, J. M.; Mattsson, T. R. The AM05 Density Functional Applied to Solids. J. Chem. Phys. 2008, 128, 084714,  DOI: 10.1063/1.2835596
    73. 73
      Perdew, J. P.; Ruzsinszky, A.; Csonka, G. I.; Vydrov, O. A.; Scuseria, G. E.; Constantin, L. A.; Zhou, X.; Burke, K. Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces. Phys. Rev. Lett. 2008, 100, 136406,  DOI: 10.1103/PhysRevLett.100.136406
    74. 74
      Zhang, Y.; Yang, W. Comment on “Generalized Gradient Approximation Made Simple”. Phys. Rev. Lett. 1998, 80, 890,  DOI: 10.1103/PhysRevLett.80.890
    75. 75
      Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 14561465,  DOI: 10.1002/jcc.21759
    76. 76
      Grimme, S.; Hansen, A.; Brandenburg, J. G.; Bannwarth, C. Dispersion-Corrected Mean-Field Electronic Structure Methods. Chem. Rev. 2016, 116, 51055154,  DOI: 10.1021/acs.chemrev.5b00533
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcc.1c10401.

    • Overview of dataset generation methods, ANN training statistics, computational methods for DFT, MC, and MKM, monolayer geometries, and lateral interaction potentials (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.